You are on page 1of 491

Biology of Metabolism

in Growing Animals
Edited by

D.G. Burrin
USDA /ARS Children’s Nutrition Research Center, Department of Pediatrics,
Baylor College of Medicine, Houston, Texas, USA

H.J. Mersmann
USDA/ARS Children’s Nutrition Research Center, Department of Pediatrics,
Baylor College of Medicine, Houston, Texas, USA

Technical Editor

E. Salek
The Kielanowski Institute of Animal Physiology and Nutrition,
Polish Academy of Sciences, Jablonna n/Warsaw, Poland

Edinburg London New York Oxford Philadelphia


St. Louis Sydney Toronto 2005
Elsevier Limited
© 2005 Elsevier Limited. All rights reserved.
No part of this publication may be reproduced, stored in a retrieval system, or transmitted in any form or
by any means, electronic, mechanical, photocopying, recording or otherwise, without either the prior per-
mission of the publishers or a licence permitting restricted copying in the United Kingdom issued by the
Copyright Licensing Agency, 90 Tottenham Court Road, London W1T 4LP. Permissions may be sought
directly from Elsevier’s Health Sciences Rights Department in Philadelphia, USA: phone: (+1) 215 238
7869, fax: (+1) 215 238 2239, e-mail: healthpermissions@elsevier.com. You may also complete your
request on-line via the Elsevier homepage (http://www.elsevier.com), by selecting ‘Customer Support’
and then ‘Obtaining Permissions’.
First published 2005
ISBN 0 444 510133
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library
Library of Congress Cataloging in Publication Data
A catalog record for this book is available from the Library of Congress
Notice
Veterinary knowledge and best practice in this field are constantly changing. As new research and experi-
ence broaden our knowledge, changes in practice, treatment and drug therapy may become necessary or
appropriate. Readers are advised to check the most current information provided (i) on procedures fea-
tured or (ii) by the manufacturer of each product to be administered, to verify the recommended dose or
formula, the method and duration of administration, and contraindications. It is the responsiblity of the
practitioner, relying on their own experience and knowldege of the patient, to make diagnoses, to deter-
mine dosages and the best treatment for each individual patient, and to take all appropriate safety precau-
tions. To the fullest extent of the law, neither the publisher nor the editors assumes any liability for any
injury and/or damage.
The Publisher

Printed in China The


Publisher’s
policy is to use
paper manufactured
from sustainable forests
Keynotes

Progress in life sciences is unbelievably quick and usually unpredictable. The amount of
research results communicated each minute, every day of the week makes it impossible to be
up-to-date even in a very narrow scientific field. The situation as regards the transfer of these
achievements to lecture halls and their integration with current “practical” scientific knowl-
edge is even worse. The gap between the latest developments in life sciences announced by the
world’s leading labs and the possibilities of their verification in medicine, biomedicine, and
animal production seems to be expanding at a geometrical rate. At the same time “more and
less” is known. It appears that the professional scientific world has run into difficulties in inte-
grating what the scientific world knows. Soon, the old Scandinavian adage “the top consultants
know everything about nothing” will be a truism.
This series of books prepared by leading professionals will try to fill the gap between practi-
cal and basic knowledge in life sciences. We believe that the authors and their selections of the
information presented in their chapters will still leave room for young animals to grow.

Stepan Pierzynowski, Prof


Series Editor

v
INSTITUTIONS PROVIDING PATRONAGE AND FINANCIAL SUPPORT

USDA/ARS Children’s Nutrition


Research Center, Department of
Pediatrics, Baylor College of Medicine,
Houston, Texas, USA

MS Milk Specialties Company

CIL Cambridge Isotope Laboratories, Inc.

ISOTEC Member of the SIGMA-


ALDRICH Family

Lund University, Sweden

The Kielanowski Institute of Animal


Physiology and Nutrition, Polish
Academy of Sciences, Poland

SGP Consulting, Lund Sweden

Gramineer International AB, Lund,


Sweden
Preface

This book Biology of Metabolism in Growing Animals is the third volume in the Elsevier
book series entitled Biology of Growing Animals. This book is intended to provide in-depth
reviews of the major areas of metabolism in growing domestic animals. The authors are
leading, internationally recognized experts in the fields of nutrition, metabolism, and physi-
ology and highlight some of the most recent advances in the field of metabolism. The chap-
ters cover important new developments in interorgan, tissue-specific, and cell-specific
metabolism of protein and amino acids, lipids and fatty acids and carbohydrates in mono-
gastric and ruminant species, including humans. The study of metabolism represents a nexus
of biological phenomena that integrates the nutrition, physiology, endocrinology, immunol-
ogy, biochemistry and cell biology in an organism. The development of new methodological
techniques and experimental approaches has provided scientists with a greater understanding
of how key nutrients or substrates are metabolized at the cellular, organ and whole animal
level. The book describes the impact of specific biochemical pathways and expression of
critical enzymes, routes of nutrient or substrate input and anatomical or structural influences
on the rates of metabolism in a given tissue or cell type. Major substrates/fuels for oxidative
metabolism, key endocrine signaling pathways and intracellular molecules that regulate the
major metabolic processes are described. Also discussed is the influence of ontogeny, stage
of differentiation and major changes in diet, or the environment, on metabolism of growing
animals. The concepts and specific findings in each area are discussed in the context of their
impact on the nutrient requirements, growth, environmental impact, health and well-being
of animals.

Acknowledgements

The editors wish to thank all of the authors for their outstanding contributions to the book.
We also thank Ewa Salek for her assistance with technical editing and Jane Schoppe for
administrative support. Thanks also go to the Series Editors, Stefan Pierzynowski and
Romuald Zabielski, for the invitation and opportunity to put together this book. We sin-
cerely thank the sponsors for their financial support, including USDA/ARS, Milk Specialties
Company, Cambridge Isotope Laboratories, and Sigma-Aldrich-Isotec Inc.

D.G. Burrin and H.J. Mersmann


Editors

vii
Dedication

Peter J. Reeds

The editors and many contributing authors of the book wish to dedicate this book to the mem-
ory of Dr. Peter Reeds. Peter Reeds was a close colleague, friend and mentor to many of the
contributing authors of this book. Peter Reeds was born in England in 1945 and completed his
Ph.D. in nutritional biochemistry at the University of Southampton, in 1971. His doctoral
research focused on the interactions between insulin and growth hormone in the regulation of
muscle protein synthesis and demonstrated the synergy between their separate mechanisms of
action. Peter Reeds went on to complete postdoctoral training at the Tropical Metabolism
Research Unit in Jamaica under the mentorship of Professor John Waterlow. His early years
of training provided a foundation in key areas that would be central themes in his career,
namely protein metabolism, isotope kinetics and growth regulation. In 1976, Peter Reeds
moved to the Rowett Research Institute in Aberdeen, Scotland, to work under the guidance of
the Director, Sir Kenneth Blaxter. During his years at the Rowett, Peter Reeds established
himself as a leader in the science of growth regulation, protein metabolism and the nutrient
requirements of farm livestock. In 1987, Peter Reeds moved to the Children’s Nutrition
Research Center in the Department of Pediatrics at Baylor College of Medicine, where he
resumed his longstanding interests in human pediatric nutrition and developmental aspects of
growth. In 2001, Peter Reeds left the Children’s Nutrition Research Center to assume a posi-
tion as Professor of Animal Sciences in the Faculty Excellence Program at the University of
Illinois at Urbana-Champaign.
During his career, Peter Reeds made many seminal contributions to our understanding of
protein and amino acid metabolism and the biology of growth regulation. His intellectual
brilliance was evident in the breadth and volume of his work. More importantly, however,

ix
x Dedication

Peter Reeds was a wonderful human being with an irrepressible wit and sense of humor. His
sense of humor was reflected in his exuberance and excitement for science, which was infec-
tious to those with whom he worked. Peter Reeds died on August 13, 2002, from complica-
tions of Legionnaire’s disease. His legacy to the science of nutrition and metabolism will be
long remembered by his countless friends, colleagues and members of the nutrition science
community.
Contributors

Ball R.O. – Department of Agricultural, Food and Nutritional Sciences, University


of Alberta, Edmonton, Alberta, Canada T6G 2P5; The Research Institute, The
Hospital for Sick Children, Toronto, Department of Nutritional Sciences, University
of Toronto, Toronto, Ontoria, Canada
Baracos V.E. – Department of Oncology, University of Alberta, Edmonton, Alberta,
Canada T6G1Z2
Bell A.W. – Department of Animal Science, Cornell University, Ithaca, NY
14853–4801, USA
Bertolo R.F.P. – Department of Biochemistry, Memorial University of Newfoundland,
St. John’s, Newfoundland, Canada, A1B 3X9
Burrin D.G. – USDA/ARS Children’s Nutrition Research Center, Department of
Pediatrics, Baylor College of Medicine, Houston, TX 77030, USA
Carstens G.E. – Department of Animal Science, Texas A&M University, College
Station, TX 77483–2471, USA
Damon M. – INRA, Joint Research Unit for Calf and Pig Production, 35590 Saint
Gilles, France
Davis T.A. – United States Department of Agriculture/Agricultural Research
Service, Children's Nutrition Research Center, Department of Pediatrics, Baylor
College of Medicine, Houston, TX 77030, USA
Donkin S.S. – Department of Animal Sciences, Purdue University, West Lafayette,
IN 47907, USA
Drackley J.K. – Department of Animal Sciences, University of Illinois, Urbana,
IL 61801, USA
Ehrhardt R.A. – Department of Animal Science, Cornell University, Ithaca, NY
14853–4801, USA
Escobar J. – Department of Animal Sciences, University of Illinois, Urbana,
IL61801, USA
Fiorotto M.L. – United States Department of Agriculture/Agricultural Research
Service, Children's Nutrition Research Center, Department of Pediatrics, Baylor
College of Medicine, Houston, TX 77030, USA
Flynn N.E. – Department of Chemistry and Biochemistry, Angelo State University,
San Angelo, TX 76909, USA
Greenwood P.L. – NSW Agriculture Beef Industry Centre, University of New
England, Armidale, NSW 2351, Australia

xi
xii Contributors

Guan X. – USDA/ARS Children’s Nutrition Research Center, Department of


Pediatrics, Baylor College of Medicine, Houston, TX 77030, USA
Hammon H. – Research Institute for Biology of Farm Animals (Oskar Kellner
Institute), 18196 Dummerstorf, Germany
Harmon D.L. – Department of Animal Sciences, University of Kentucky,
Lexington, KY 40546-0215, USA
Herpin P. – INRA, Joint Research Unit for Calf and Pig Production, 35590, Saint-
Gilles, France
Huntington G.B. – Department of Animal Science, North Carolina State
University, Raleigh, NC 27695-7621, USA
Innis S.M. – Department of Paediatrics, University of British Columbia, Vancouver,
British Columbia, Canada, V5Z 4H4
Jesse B.W. – Department of Animal Science, Rutgers, The State University of
New Jersey, New Brunswick, NJ 08901-8525, USA
Johnson R.W. – Department of Animal Sciences, University of Illinois, Urbana, IL
61801, USA
Knabe D.A. – Department of Animal Science and Faculty of Nutrition, Texas A & M
University, College Station, TX 77843-2471, USA
Kristensen N.B. – Department of Animal Nutrition and Physiology, Danish
Institute of Agricultural Sciences, DK-8830 Tjele, Denmark
Le Dividich J. – INRA, Joint Research Unit for Calf and Pig Production, 35590
Saint-Gilles, France
Lin X. – Department of Animal Science, North Carolina State University, Raleigh,
NC 27695-7621, USA
Louveau I. – INRA, Joint Research Unit for Calf and Pig Production, 35590 Saint-
Gilles, France
Lyvers-Peffer P. – Department of Animal Science, North Carolina State University,
Raleigh, NC 27695-7621, USA
Mersmann H.J. – USDA/ARS Children’s Nutrition Research Center, Department
of Pediatrics, Baylor College of Medicine, Houston, TX 77030, USA.
Odle J. – Department of Animal Science, North Carolina State University, Raleigh,
NC 27695-7621, USA
Pencharz P.B. – Department of Paediatrics, University of Toronto, Toronto, Ontario,
Canada M5G 1X8; The Research Institute, The Hospital for Sick Children, Toronto,
Department of Nutritional Sciences, University of Toronto, Toronto, Ontaria, Canada
Reynolds C.K. – Department of Animal Sciences, The Ohio State University,
OARDC, 1680 Madison Avenue, Wooster, OH 44691-4096, USA
Smith S.B. – Department of Animal Science, Texas A & M University, College
Station, TX 77843-2471, USA
Stoll B. – USDA/ARS Children’s Nutrition Research Center, Department of
Pediatrics, Baylor College of Medicine, Houston, TX 77030, USA
Wu G. – Department of Animal Science and Faculty of Nutrition, Texas A & M
University, College Station, TX 77843-2471, USA
1 Regulation of metabolism and growth
during prenatal life

A. W. Bella, P. L. Greenwoodb, and R. A. Ehrhardt a

aDepartment
of Animal Science, Cornell University, Ithaca, NY 14853-4801, USA
bNSW Agriculture Beef Industry Centre, University of New England, Armidale,
NSW 2351, Australia

Fetal energy and nitrogen requirements are met mostly by placental transfer of glucose and
amino acids; fatty acids may contribute additional energy in some species. Placental metab-
olism accounts for much of the total net consumption of oxygen and macronutrients by the
conceptus, and alters the composition of nutrients delivered to the fetus. The molecular basis
for the facilitated transport of glucose by the placenta is well described; molecular character-
ization of the more complex systems for the active transport of most amino acids is under
way. Maternal and placental macronutrient supply is a powerful regulator of fetal metabolism
and growth, especially in late gestation. Endocrine mediation of these responses matures
as gestation advances, adding to the influences of locally expressed regulators throughout
gestation. Insulin, thyroid hormones, and, near term, corticosteroids, are especially influential
in the direct and indirect control of fetal nutrient disposal and tissue growth. Prenatal growth
retardation does not necessarily constrain the rate of neonatal growth, but at any given post-
natal body weight, low-birth-weight lambs are fatter and have smaller muscles. Experimental
evidence is accumulating for longer-term influences of prenatal nutrition through fetal pro-
gramming of propensity for mature-onset diseases such as hypertension and type II diabetes.

1. INTRODUCTION
The coordination of nutrient supply with tissue metabolism and growth during prenatal life in
placental mammals is complex due to the varying influences of maternal nutrition and meta-
bolic adaptations to the state of pregnancy, placental function, and gestational maturation
of fetal endocrine and local regulatory systems. It is important to understand the separate
and interdependent mechanisms by which these factors exert their effects on fetal growth
and development, for several reasons. Increased neonatal mortality and morbidity in low-birth-
weight offspring remain major problems in some human and livestock populations, despite
decades of study on the multifaceted etiology of intrauterine growth retardation (IUGR).

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
3 © 2005 Elsevier Limited. All rights reserved.
4 A. W. Bell et al.

Fetal overgrowth due to maternal nutrition or diseases, such as diabetes, also increases perina-
tal mortality and incidence of postnatal problems. More intriguing and, possibly, with major
ramifications for long-term health and productivity of humans and other animals, is the
emerging evidence that fetal metabolic disturbance can lead to “programming” of increased
predisposition to various disease syndromes during later postnatal life.
This chapter will summarize briefly the quite detailed state of knowledge of quantitative
metabolism of macronutrients in individual tissues and whole body of the fetus, and in the
placenta, with emphasis on data obtained in vivo. The current understanding of placental
transport of macronutrients and its implications for fetal nutrition and growth will be treated
similarly. These topics will be a prelude to the major theme of regulation and coordination of
metabolism and growth in the conceptus. Finally, the influence of prenatal experience
on postnatal performance will be considered, with brief reference to recent experimental
evidence for the concept of “fetal programming”.

2. MAJOR FEATURES OF CONCEPTUS METABOLISM AND GROWTH


2.1. Patterns of prenatal growth

Early embryonic development, including organogenesis and initiation of placentation, is beyond


the scope of this review. The morphology of embryo development in domestic animal species
has been described by Noden and deLahunta (1985). Patterns of fetal and placental growth in
the normal and growth-retarded sheep conceptus are illustrated in fig. 1. In this species, as in

Fig. 1. Patterns of fetal and placental growth in the normal (——) and growth-retarded (---) sheep conceptus.
Adapted from the data of Ehrhardt and Bell (1995) and Greenwood et al. (2000).
Regulation of metabolism and growth during prenatal life 5

other placental mammals, postembryonic growth becomes quantitatively significant only


after mid-gestation. However, this is preceded by rapid hyperplastic growth of the placenta,
which attains all or most of its mass of dry tissue, protein, and DNA by mid-gestation
(Ehrhardt and Bell, 1995). Fetal growth then follows its familiar, flattened sigmoid pattern
during the latter half of gestation as it proceeds from an early exponential phase through a
rapid, linear phase, and then, as term approaches, begins to diminish in rate. In most species,
there is little or no increase in placental weight during this period; the ovine placenta actually
diminishes in weight, mostly due to loss of extracellular water (Ehrhardt and Bell, 1995).
However, the placenta undergoes extensive tissue remodeling after mid-gestation, including
major proliferative growth of the umbilical vasculature (Teasdale, 1976), which is associated
with a progressive increase in its functional capacity. Relations between placental size and
function, and implications for fetal growth, are discussed in the next section.

2.2. Fetal requirements and metabolism of macronutrients

Numerous studies on pregnant ewes have described fetal macronutrient requirements and
metabolism in terms of umbilical exchanges of oxygen, nutrients, and metabolites, and of
rates of net accretion of nutrients in growing tissues (see Battaglia and Meschia, 1988; Bell,
1993). These and similar data from pregnant cows (Comline and Silver, 1976; Reynolds et al.,
1986; Ferrell, 1991) are summarized in table 1.
During late pregnancy in these species, 35–40% of fetal energy is taken up as glucose and its
fetal-placental metabolite, lactate, and a further 55% is taken up as free amino acids. In contrast
to its importance as an energy source in the maternal ruminant, umbilical uptake of acetate could
account for only 5–10% of fetal energy consumption. Placental capacity for transfer of long-
chain, nonesterified fatty acids (NEFA) and keto-acids is even more limited (see Bell, 1993),
making these maternal substrates trivial contributors to fetal metabolism. Almost all of the
nitrogen acquired by the fetus is in the form of amino acids, but a small net umbilical uptake
of ammonia is derived from placental deamination of amino acids during the latter half of

Table 1
Fetal sources and disposal of energy and nitrogen in ewes and cows during late pregnancy

Energy (kJ/kg·d) Nitrogen (g/kg·d)

Ewe Cow Ewe Cow

Sources
Glucose + lactate 217a 114f — —
Amino acids 177a 156g 1.19a 1.09g
Acetate 20b 30h — —
NH3 — — 0.05e ND
Total 414 300 1.24 1.09
Disposal
Accretion 133c 72i 0.79c 0.34i
Heat 240a 192g — —
Urea 16d 15g 0.36d 0.66g
Glutamate + serine efflux 16a ND 0.11a ND
Total 405 279 1.26 1.00
a Chung et al. (1998), b Char and Creasy (1976), c McNeill et al. (1997), d Lemons and Schreiner (1983),
e Holzman et al. (1977), f Reynolds et al. (1986), g Ferrell (1991), h Comline and Silver (1976), i Ferrell et al. (1976).
6 A. W. Bell et al.

gestation (Holzman et al., 1977; Bell et al., 1989). About 60% of these amino acids are used for
tissue protein synthesis, which accounts for ~18% of fetal energy expenditure (Kennaugh et al.,
1987). The remaining 40% are rapidly catabolized, accounting for at least 30% of the oxidative
requirements in the well-nourished sheep fetus (Faichney and White, 1987), or, in the case of
glutamate and serine, taken up and metabolized by the placenta (Battaglia and Regnault, 2001).
Less comprehensive studies of the fetal pig (Fowden et al., 1997) and horse (Fowden and
Silver, 1995) suggest that in these species during late pregnancy, glucose is an even more
important energy substrate than in fetal ruminants. The fetal horse, at least, appears to
make less extensive use of amino acids as a source of energy (Silver et al., 1994; Fowden
et al., 2000a).
In all species studied, the fetal liver and, to a lesser extent, kidneys, develop the enzymatic
capacity for gluconeogenesis during late gestation (see Fowden, 1997). In the well-fed,
unstressed sheep fetus, endogenous glucose synthesis is negligible (Hay et al., 1984; Leury
et al., 1990a). However, significant endogenous synthesis of glucose can be induced by
maternal starvation or chronic undernutrition, presumably due to hepatic gluconeogenesis
from amino acids (Hay et al., 1984; Leury et al., 1990a). Acute hypoxia and other stressors
also increase net hepatic release of glucose due to increased rates of gluconeogenesis and/or
glycogenolysis in fetal sheep (Rudolph et al., 1989; Townsend et al., 1991).

2.3. Metabolism of nonfetal conceptus tissues

2.3.1. Glucose metabolism


The major contribution of the nonfetal components of the gravid uterus, especially the placenta,
to oxygen and nutrient requirements of the conceptus is sometimes ignored. However, these
requirements greatly affect the partitioning of nutrients within the gravid uterus and add sub-
stantially to the nutrient demands upon the dam. In late-pregnant ewes and cows, the aggregate
weight of placentomes, consisting of fetal (cotyledonary) and maternal (caruncular) tissues, is
less than 15% that of the attached fetus. However, the weight-specific metabolic rate of the pla-
centa is so great that the uteroplacental tissues (placentomes, endometrium, myometrium)
consume 35–50% of the oxygen and 60–70% of the glucose taken up by the uterus in ewes
(Meschia et al., 1980) and cows (Reynolds et al., 1986). The weight-specific consumption of
glucose by the diffuse placental tissues of the horse and pig is even greater than that of the
epitheliochorial ruminant placenta, accounting for 80–90% of uterine glucose uptake during late
gestation (Fowden, 1997).
In all species, a considerable fraction of the glucose consumed by uteroplacental tissues is
converted to lactate. Rates of lactate production and disposal into maternal and fetal circula-
tions vary with species and gestational age. For example, production is relatively high and
distributed mostly into the uterine circulation during late pregnancy in the mare, whereas the
lower production in ruminants is mostly released into the umbilical circulation (Fowden,
1997). In ruminants, horse, and pig, a further, smaller fraction of glucose consumed by utero-
placental tissues is converted to fructose which is released into the fetal circulation and slowly
metabolized by fetal tissues (Meznarich et al., 1987).

2.3.2 Amino acid metabolism

Net uteroplacental consumption of amino acids, as a fraction of uterine uptake, is lower


than that of glucose, presumably related to the negligible or small growth of the placenta
Regulation of metabolism and growth during prenatal life 7

and uterine tissues in sheep (Ehrhardt and Bell, 1995) and cattle (Bell et al., 1995) during
late pregnancy. Nevertheless, net removal by the uteroplacental tissues has been estimated to
account for 24% of uterine uptake of amino acid nitrogen in well-fed ewes during late preg-
nancy (Chung et al., 1998).

2.4. Gestational development of conceptus metabolism

The many-fold increase in fetal mass from mid- to late gestation is, not unexpectedly, accom-
panied by increased absolute rates of uterine and umbilical uptake of oxygen and nutrients
and of urea export by conceptus tissues, and of fetal whole-body protein synthesis in sheep
and cattle (Bell et al., 1986, 1989; Reynolds et al., 1986; Kennaugh et al., 1987; Ferrell, 1991).
However, when expressed on a weight-specific basis these rates are considerably greater in
mid than in late gestation, concomitant with greater rates of relative growth in the immature
fetus. More recent studies of fetal and uteroplacental metabolic ontogeny in the horse have
shown a qualitatively similar pattern (Fowden et al., 2000a,b). The apparent absence of a
decrease in weight-specific fetal oxygen consumption between mid- and late gestation in this
species (Fowden et al., 2000a) may be related to its slower relative rates of fetal growth and
the failure to account for the greater tissue hydration of the mid-gestation fetus. In sheep, the
gestational decline in weight-specific fetal whole-body metabolic rates is associated with
changes in the allometric growth of metabolically active vital organs, such as the liver, versus
that of less active skeletal tissues (Bell et al., 1987a), as well as a decline in the weight-specific
rate of fetal hepatic oxygen consumption (Vatnick and Bell, 1992).

3. PLACENTAL TRANSPORT OF MACRONUTRIENTS


3.1. Molecular and physiological mechanisms

3.1.1. Glucose

Glucose is transported from the maternal to the fetal circulation by carrier-mediated, facilitated
diffusion (Widdas, 1952; Simmons et al., 1979). This process is strongly dependent on the
maternal–fetal plasma glucose concentration gradient (Simmons et al., 1979; DiGiacomo and
Hay, 1990a). The predominant glucose transporter protein isoforms in the sheep placenta are
GLUT-1 and GLUT-3 (Ehrhardt and Bell, 1997; Das et al., 1998), the mRNA and protein
abundance of which increase with gestational age, especially for GLUT-3 (Currie et al., 1997;
Ehrhardt and Bell, 1997). This, together with its low Km and localization at the apical, maternal-
facing layer of the trophoblastic cell layer (Das et al., 2000), suggests that ontogenic changes
in GLUT-3 expression and activity may account for much of the 5-fold increase in glucose
transport capacity of the sheep placenta in vivo between mid- and late gestation (Molina
et al., 1991). Other factors must include remodeling and expansion of the placenta’s effec-
tive exchange surface (Stegeman, 1974) and the increasing maternal–fetal plasma
concentration gradient (Molina et al., 1991). Similar developmental patterns in placental
GLUT expression have been observed in the rat (Zhou and Bondy, 1993) but not in the human
(see Illsley, 2000) or horse placenta (Wooding et al., 2000), in which gestational changes
were small or absent. These species differences may be due to the considerably slower rates
of fetal growth and glucose demand in humans and horses, and, possibly, their greater
dependence on changes in placental morphology to permit increased fetal access to glucose
during late pregnancy.
8 A. W. Bell et al.

3.1.2. Amino acids

Most amino acids taken up by the placenta are transported against a fetal–maternal concen-
tration gradient, implying the use of energy-dependent, active transport processes (Young and
McFadyen, 1973). Studies of isolated human and rodent placental vesicles have confirmed that
the transport systems in the placenta are similar to those described for plasma membranes of
other tissues (see Battaglia and Regnault, 2001). These include at least six sodium-dependent
and five sodium-independent systems that have been classified systematically on the basis of
their affinity for neutral, acidic, or basic amino acids, and their intracellular location (Battaglia
and Regnault, 2001). Recent results from in vivo studies on sheep suggest that rapid placental
transport of neutral amino acids requires both sodium-dependent transport at the maternal
epithelial surface and affinity for highly reversible, sodium-independent transporters located
at the fetal surface (Jozwik et al., 1998; Paolini et al., 2001). These researchers also demon-
strated major differences in placental clearance among the essential amino acids, with the
more rapidly transported branched-chain acids, plus methionine and phenylalanine, apparently
sharing the same rate-limiting transport system (Paolini et al., 2001).

3.1.3. Fatty acids

Placental capacity for maternal–fetal transport of short- and long-chain fatty acids and their
keto-acid derivatives varies widely among species, associated with variations in placental
structure (see Bell and Ehrhardt, 2002). Thus, the epitheliochorial placentae of ruminants
and the diffuse placentae of pigs and horses allow only meager fetal access to maternal
fatty acids and ketones, whereas the hemochorial placentae of rodents, lagomorphs, and, by
inference, humans, are more permeable to these substrates. Molecular mechanisms for pla-
cental transport of fatty acids have yet to be defined but may involve a placenta-specific
fatty-acid binding protein that has been identified in sheep (Campbell et al., 1994) and
humans (Campbell et al., 1995).

3.2. Influence of placental metabolism on maternal–fetal nutrient transfer

3.2.1. Glucose metabolism

Glucose entry into the gravid uterus and its component tissues is determined by maternal arte-
rial glucose concentration (Hay and Meznarich, 1988; Leury et al., 1990b), while glucose
transport to the fetus is determined by the transplacental (maternal–fetal) concentration gra-
dient (Hay et al., 1984). In turn, the transplacental gradient is directly related to both placental
and fetal glucose consumption, which are dependent on fetal arterial glucose concentration
(Hay et al., 1990). Thus, as fetal glucose concentration changes relative to that of the mother,
thereby changing the transplacental gradient, placental transfer of glucose to the fetus varies
reciprocally with placental glucose consumption.
In addition to its quantitative impact on placental transfer of glucose, placental glucose
metabolism has a major qualitative influence on the pattern of carbohydrate metabolites
delivered to the fetus. Rapid metabolism to lactate (~35%), fructose (~4%), and CO2 (~17%)
accounted for about 56% of uteroplacental glucose consumption in late-pregnant ewes, and
was directly related to placental glucose supply (Aldoretta and Hay, 1999). The fate of the
remaining 44% of glucose metabolized by the placenta must include synthesis of alanine and other
nonessential amino acids (Timmerman et al., 1998), directly or via lactate (Carter et al., 1995).
Regulation of metabolism and growth during prenatal life 9

Umbilical uptake and fetal oxidation of placentally derived lactate (Sparks et al., 1982; Hay
et al., 1983) and fructose (Meznarich et al., 1987) are estimated to contribute approximately
20% and 5%, respectively, to fetal CO2 production, in addition to the 30% contributed by the
rapid oxidation of glucose (Hay et al., 1983).

3.2.2. Amino acid metabolism

Placental metabolism also affects the quantity and composition of amino acids delivered to
the fetus. The significant net consumption by uteroplacental tissues of glutamate, serine, and
the branched-chain amino acids (Liechty et al., 1991b; Chung et al., 1998) implies catabolism
or transamination of these acids. An additional, small fraction of this net loss of amino acids
will be in the form of secreted peptides.
The ovine placenta has very little enzymatic capacity for urea synthesis, but produces con-
siderable amounts of ammonia, much of which is released into maternal and, to a lesser
extent, fetal circulations (Holzman et al., 1977; Bell et al., 1989). This is consistent with
extensive placental deamination of branched-chain amino acids to their respective keto-acids,
which are released into fetal and maternal bloodstreams (Smeaton et al., 1989; Loy et al.,
1990), and with rapid rates of glutamate oxidation in the placenta (Moores et al., 1994).
Transamination of branched-chain amino acids accounts for some of the net glutamate acqui-
sition by the placenta, the remainder of which is taken up from the umbilical circulation
(Moores et al., 1994). That which is not quickly oxidized combines with ammonia to synthe-
size glutamine, which is then released back into the umbilical bloodstream (Chung et al.,
1998). Some of this glutamine is converted back to glutamate by the fetal liver, which pro-
duces most of the glutamate consumed by the placenta (Vaughn et al., 1995). This establishes
a glutamate–glutamine shuttle which promotes placental oxidation of glutamate and fetal
hepatic utilization of the amide group of glutamine.
Similarly, the placenta almost quantitatively converts serine, mostly taken up from maternal
blood, to glycine (Chung et al., 1998), reconciling the discrepancy between the negligible net
uptake of glycine by the uterus and substantial net release of this amino acid into the umbilical
circulation (see Hay, 1998).
The complexity of interrelations among placental uptake, metabolism, and transport of
amino acids was further illustrated by a study of alanine metabolism in ewes during late
pregnancy (Timmerman et al., 1998). Application of tracer methodology showed that negli-
gible net placental consumption or production of alanine masks an appreciable metabolism of
maternal alanine entering the placenta which exchanges with endogenously produced alanine.
Thus, most of the alanine delivered to the fetus is of placental origin, derived from placental
protein turnover and transamination.

3.2.3. Fatty acid metabolism

The relative abundance of polyunsaturated C20 and C22 derivatives of the essential C18 fatty
acids in fetal tissues has been attributed largely to the placenta’s capacity for hydrolyzing
esterified lipids (Clegg, 1981) and for desaturation and chain elongation of the resulting free
polyunsaturated C18 acids (Noble et al., 1985). Thus, placental metabolism ensures an
adequate fetal supply of the longer-chain ω6 and ω3 metabolites of the C18 essential fatty
acids, which are the forms ultimately required by tissues, despite the poor placental transport
of the parent molecules
10 A. W. Bell et al.

3.3. Factors affecting placental transport capacity

3.3.1. Placental size

Placental capacity for glucose transport was substantially reduced, as were uteroplacental
glucose consumption rate and fetal glycemia, in carunclectomized (Owens et al., 1987a) and
heat-treated ewes (Bell et al., 1987b; Thureen et al., 1992). At least part of the absolute reduc-
tion in glucose transport capacity is presumed to be due to reduction in exchange surface area
of the trophoblastic membrane, as shown in carunclectomized ewes (Robinson et al., 1995).
In previously heat-treated ewes (Thureen et al., 1992), placental weight-specific glucose
transport capacity was also reduced. This implies that chronic heat stress, which reduces aver-
age weight but not total number of placentomes, additionally reduces the number and/or
activity of specific glucose transport proteins at maternal and/or fetal exchange surfaces. In
contrast, carunclectomy, which reduces placentome number but may stimulate a compensa-
tory increase in average weight of individual placentomes, caused a modest increase in the
placental weight-specific clearance of the nonmetabolizable glucose analog, 3-O-methyl
glucose (Owens et al., 1987b). This implies that glucose transporter expression was preserved
or increased in the remaining placentomes.
Placental insufficiency in heat-treated ewes also extends to impaired capacity for amino
acid transport, including major reductions in placental uptake and fetal transfer of leucine
(Ross et al., 1996) and threonine (Anderson et al., 1997). The normally extensive placental
catabolism of leucine was also greatly reduced (Ross et al., 1996).

3.3.2. Maternal nutrition

Recent evidence indicates that the activity of placental transport mechanisms can be modu-
lated by maternal nutrition, independent of more chronic effects on placental size. For
example, moderate undernutrition of ditocous ewes during late pregnancy caused a 50%
increase in capacity for maternal–fetal glucose transport in vivo (Ehrhardt et al., 1996) which
was at least partly explained by a 20% increase in total GLUT abundance, associated with a
similar increase in GLUT-3 protein abundance (Ehrhardt et al., 1998). These responses help
explain how placental glucose transfer remained sufficient to sustain normal fetal growth,
despite chronic maternal hypoglycemia and a 26% decrease in the maternal–fetal gradient in
arterial plasma glucose concentration (Bell et al., 1999).
During more severe, chronic undernutrition or starvation for several days, the development
of profound fetal hypoglycemia helps to sustain the maternal–fetal gradient in glucose con-
centration by restricting the reverse transfer of glucose to the placenta, and reducing placental
glucose consumption (see Hay, 1995). Under these more stringent conditions, fetal gluco-
neogenesis is induced (Leury et al., 1990a), with amino acids being the presumed major
substrate, consistent with increased fetal urea synthesis (Lemons and Schreiner, 1983;
Faichney and White, 1987). The ultimate consequence is reduced fetal tissue protein synthesis
(Krishnamurti and Schaefer, 1984) and slowing of fetal growth to a rate that can be sustained
by the reduced placental nutrient supply.
Effects of energy and/or protein supply on placental capacity for amino acid transport have been
little studied. Fasting late-pregnant ewes for 5 days had an insignificant effect on umbilical net
uptake of amino acids despite appreciable decreases in maternal arterial plasma concentrations of
many amino acids (Lemons and Schreiner, 1983). This suggests that during short-term energy/
protein deprivation, placental mechanisms for active transport of amino acids are unimpaired and
may even be upregulated. Under similar fasting conditions, the uteroplacental deamination of
Regulation of metabolism and growth during prenatal life 11

branched-chain amino acids appeared to be increased, judging from a 3-fold increase in the efflux
of α-ketoisocaproate, the keto-acid derivative of leucine, into uterine and umbilical circulations
(Liechty et al., 1991a). This suggests that increased amino acid catabolism may partly compen-
sate for the likely reduction in placental glucose oxidation under these conditions.
Placental transport and metabolism of amino acids have not been studied during more pro-
longed restriction of energy or protein. However, in ewes fed adequate energy but insufficient
protein during the last month of pregnancy, fetal growth and protein deposition over this
period were decreased by 18% (McNeill et al., 1997). It is also notable that in chronically
hyperglycemic ewes with secondary hyperinsulinemia and hypoaminoacidemia, placental
and fetal uptakes of several amino acids were reduced, and fetal total nitrogen uptake declined
by 60% (Thureen et al., 2001).

3.4. Consequences of placental insufficiency

Placental weight and associated capacity for maternal–fetal nutrient transfer are powerful
determinants of fetal growth during late gestation in all species studied. This has been most
persuasively demonstrated by controlled manipulation of placental size and/or functional
capacity using premating carunclectomy (Alexander, 1964), heat-induced placental stunting
(Alexander and Williams, 1971), or uteroplacental vascular embolization (Creasy et al., 1972).
Natural variations in fetal weight due to varying litter size in prolific ewes are strongly corre-
lated with placental mass per fetus (Rhind et al., 1980; Greenwood et al., 2000). Recently,
the quite profound growth retardation of fetuses in overfed, primiparous ewes also has been
attributed to a primary reduction in placental growth (Wallace et al., 2000).
The probably common etiology of IUGR in experimentally induced and natural cases of
placental insufficiency is illustrated by the similar patterns of association between fetal
and placental weights in pregnant ewes with varying conceptus weights due to carunclectomy,
heat stress, litter size, and overfeeding of primiparous dams (fig. 2). In each case, severe
growth retardation was associated with chronic fetal hypoxemia and hypoglycemia during late
gestation (Creasy et al., 1972; Harding et al., 1985; Bell et al., 1987b; Wallace et al., 2002).

4. REGULATION OF CONCEPTUS METABOLISM


4.1. General features

The extracellular and local regulation of fetal metabolism and its relation to tissue growth has
several distinctive characteristics. First, placental nutrient supply has a powerful, limiting
influence on nutrient disposal, especially in late gestation when fetal demands are greatest.
Second, the fetal endocrine system is largely independent from the direct influence of mater-
nal hormones because the placenta is impermeable to most of the important metabolic
regulatory peptide and steroid hormones. Thus, reported effects of maternal hormones on
fetal growth must be mediated indirectly by changes in maternal metabolism and/or utero-
placental tissue growth and resulting changes in fetal nutrient supply. Examples include the
effects of maternal treatment with growth hormone (GH) during early pregnancy on fetal
growth in pigs (Sterle et al., 1995; Rehfeldt et al., 2001) and of maternal immunization
against placental lactogen (PL) on fetal growth in sheep (Leibovich et al., 2000). Third, while
most fetal endocrine organs develop the capacity to synthesize and secrete hormones early in
gestation, target tissue receptor and neuroendocrine feedback systems are variably immature
until late pregnancy. As a result, there is a much greater reliance on paracrine and autocrine
regulation by locally expressed factors, especially in early and mid-pregnancy.
12 A. W. Bell et al.

Fig. 2. Relation between fetal and placental weights in ewes representing different models of placental
insufficiency during late pregnancy. Variation in placental weight was achieved by premating carunclectomy
(●; Owens et al., 1986), chronic heat treatment (䊊; Bell et al., 1987b), natural variation in litter size
(▲; Greenwood et al., 2000), and overfeeding of adolescent ewes (䉭; Wallace et al., 2000). Reproduced with
permission from the Society for Reproduction and Fertility (Greenwood and Bell, 2003).

4.2. Nutrient supply

4.2.1. Glucose

The Km for saturable glucose transport by the sheep placenta is ~3.9 mM (Simmons et al.,
1979), which is within the physiological range of glycemia in well-fed, adult sheep. Thus,
uterine uptake, placental metabolism and transfer, and fetal metabolism of glucose are very
sensitive to maternal arterial glucose concentration in sheep (fig. 3; Hay and Meznarich,
1988). In sheep, cows, and horses fetal utilization of glucose is highly correlated with fetal
plasma glucose concentration, which, in turn is correlated with maternal glycemia (see
Fowden, 1997). In contrast, fetal glucose metabolism was not related to fetal glycemia in
pigs, possibly because in this species, fetal glycemia is influenced by individual relative to
total fetal mass, as well as maternal nutrition (Fowden et al., 1997).
It is well established that in sheep, the maternal and fetal hypoglycemia caused by starva-
tion or chronic undernutrition is associated with increased fetal urea synthesis (Hodgson et al.,
1982; Lemons and Schreiner, 1983; Faichney and White, 1987) due to increased amino acid
deamination (Krishnamurti and Schaefer, 1984; Van Veen et al., 1987). Conversely, fetal
hyperglycemia appears to cause substitution of glucose for amino acids as an oxidative fuel
because under these conditions, increased glucose oxidation (Hay and Meznarich, 1986) is
associated with decreased leucine oxidation (Liechty et al., 1991a). Interestingly, the latter
response occurred independently of glucose-induced changes in fetal insulin concentration
(Liechty et al., 1993).
Fetal glucose supply also influences fetal endogenous glucose production, presumably due
to hepatic gluconeogenesis. In addition to the association of increased endogenous production
Regulation of metabolism and growth during prenatal life 13

Fig. 3. Relations between maternal arterial blood glucose concentration and (A) uterine, (B) fetal, and
(C) uteroplacental net uptakes of glucose in ewes during late pregnancy. Reproduced with permission from
the Society for Experimental Biology and Medicine (Hay and Meznarich, 1988).
14 A. W. Bell et al.

with fetal hypoglycemia in starved or undernourished ewes (Hay et al., 1984; Leury et al., 1990a),
progressive fetal hypoglycemia induced by different levels of maternal insulin infusion
caused fetal endogenous glucose production to increase linearly (DiGiacomo and Hay, 1990b).
A mediating role for fetal insulin was suggested by the incomplete suppression of endoge-
nous glucogenesis by fetal infusion with insulin while maintaining basal fetal glycemia
(DiGiacomo and Hay, 1990b).

4.2.2. Amino acids

Effects of amino acid supply on fetal metabolism have not been studied systematically.
Decreased maternal plasma concentrations of essential amino acids in fasted ewes were not
associated with a significant decrease in umbilical uptake of these acids (Lemons and
Schreiner, 1983). In contrast, maternal hyperglycemia with secondary hyperinsulinemia and
hypoaminoacidemia caused substantial reductions in uterine, uteroplacental, and fetal uptakes
of several amino acids, particularly the branched-chain acids, and a 60% reduction in total
fetal uptake of nitrogen (Thureen et al., 2000, 2001). Correction of maternal amino acid con-
centrations by appropriate exogenous infusion restored uterine and umbilical exchanges to
normal levels (Thureen et al., 2000). Maternal hyperaminoacidemia, caused by infusion of
amino acids, had little effect on the umbilical uptake of most amino acids, except for
increased uptake of the branched-chain acids, and did not affect fetal total nitrogen supply
(Jozwik et al., 1999). However, uteroplacental production and fetal concentrations of ammo-
nia increased moderately, implying some increase in placental deamination of amino acids.

4.3. Fetal hormones and growth factors

4.3.1. Pancreatic hormones

Insulin protein and preproinsulin mRNA are detectable from early gestation in the fetal pan-
creas of all species studied (D’Agostino et al., 1985). In the sheep fetus, gestational increases
in pancreatic and basal plasma concentrations of insulin (Alexander et al., 1968) are accom-
panied by a steady increase in glucose- and arginine-stimulated insulin secretion during the
latter half of gestation (Aldoretta et al., 1998). Euglycemic, hyperinsulinemic clamp studies
have demonstrated that fetal insulin and glucose have independent, positive effects on fetal
whole-body glucose utilization (Hay et al., 1988). These observations are consistent with
tissue-specific responses that vary between insulin-responsive tissues, such as skeletal muscle
(Wilkening et al., 1987; Anderson et al., 2001b), and tissues unresponsive to insulin, such as
the brain (Anderson et al., 2001a).
Neither fetal (Jodarski et al., 1985) nor maternal (Rankin et al., 1986) plasma insulin
concentration has a direct effect on placental transport of glucose, consistent with our failure
to detect significant concentrations of the insulin-responsive glucose transport protein,
GLUT-4, in the ovine placenta (Ehrhardt and Bell, 1997). However, fetal hyperinsulinemia
indirectly promotes placental transfer and umbilical uptake of glucose through its influence
on fetal glycemia and the maternal–fetal glucose concentration gradient (see Hay, 1995).
In addition to its promotion of glucose uptake and metabolism in fetal tissues, a physio-
logical increase in fetal plasma insulin stimulated umbilical uptake and whole-body
utilization of amino acids when fetal glycemia and aminoacidemia were carefully controlled
(Thureen et al., 2000). The specific metabolic fates of amino acids were not measured, but it
is likely that protein anabolism was increased by both reduction of proteolysis (Milley, 1994)
Regulation of metabolism and growth during prenatal life 15

and stimulation of protein synthesis (Horn et al., 1983). This anabolic effect may have been
reinforced indirectly by the effect of increased glucose utilization in reducing amino acid
deamination (Liechty et al., 1993).
Independently of its metabolic effects, insulin may influence fetal tissue growth through
modulation of the expression and activity of other growth regulators such as the insulin-like
growth factor (IGF) system. For example, when fetal plasma glucose and amino acid concen-
trations were clamped, fetal insulin infusion caused an increase in plasma concentration of IGF
binding protein (BP)-3 and a decrease in hepatic expression of IGFBP-1 (Shen et al., 2001).
The latter response is consistent with the opposite effects of hypoinsulinemia in the under-
nourished sheep fetus (Osborn et al., 1992). Ovine fetal hyperinsulinemia also increased the
farnesylation of p21 Ras in ovine fetal liver, skeletal muscle, adipose tissue, and white blood
cells (Stephens et al., 2001). This is significant because the Ras pathway is an important intra-
cellular signaling element in the mitogenic actions of insulin and other growth factors,
including the IGFs, and greater availability of farnesylated Ras augments mitogenic cellular
responsiveness to IGF-1 and other growth factors in isolated systems (Goalstone et al., 1998).
The fetal pancreas synthesizes glucagon from early in gestation, but the regulation and
metabolic role of this peptide in fetal life remain unclear. Secretory responses to hypo-
glycemia and other metabolic stimuli in fetal sheep are small and sluggish during late
gestation (Alexander et al., 1976), but birth is accompanied by a major surge in secretion of
glucagon (Grajwer et al., 1977). Exogenous administration of glucagon to mimic fetal plasma
concentrations observed during maternal fasting (Schreiner et al., 1980) caused hyper-
glycemia in the fetal sheep (Philipps et al., 1983), implying a possible role in regulation of
hepatic glycogenolysis and/or gluconeogenesis.

4.3.2. Growth hormone and the IGF system

During postnatal life, growth hormone (GH) is a powerful homeorhetic regulator of metabolism
and growth through its direct actions on some tissues, such as adipose tissue, and its indirect
actions on most lean tissues, mediated by the IGF system (see Etherton and Bauman, 1998).
Notable among its pleiotropic effects are inhibition of lipogenesis and enhancement of
responses to lipolytic stimuli in adipose tissue, and potent effects on cell cycle activity and
protein turnover in muscle and other tissues via regulation of the expression of IGF-1 and its
binding proteins in multiple tissues, including the liver. In general, these effects are greatly
muted during fetal life, which is characterized by persistently high plasma levels of GH (Bassett
et al., 1970; Gluckman et al., 1979) and low plasma levels of IGF-1 (Van Vliet et al., 1983). The
apparent uncoupling of the GH/IGF-1 axis is consistent with low hepatic expression of the GH
receptor, IGF-1, IGFBP-3, and the acid-labile subunit (ALS) (Klempt et al., 1993; Rhoads et al.,
2000a). Thus, although pituitary secretion of GH is active through much of gestation (de Zegher
et al., 1989), maturation of the endocrine IGF-1 system is retarded by hepatic unresponsiveness
to GH, which, in postnatal life, strongly regulates expression of all three components of the
ternary binding complex (IGF-1, IGFBP-3, ALS) that accounts for most circulating IGF-1
(Boisclair et al., 2001). Therefore, it is not surprising that infusion of normal sheep fetuses with
GH for 10 days had no effect on fetal plasma IGF-1 levels (Bauer et al., 2000).
It is possible that some direct metabolic effects of GH develop before engagement of the
GH/IGF-1 system. For example, Bauer et al. (2000) reported a decrease in glucose uptake
and, presumably, utilization, with no change in plasma insulin in GH-infused fetal sheep,
consistent with an earlier report of apparent insulin resistance in GH-treated fetuses (Parkes
and Bassett, 1985). Also, hypophysectomy of fetal lambs causes increased fat deposition that
16 A. W. Bell et al.

can be reversed by GH administration (Stevens and Alexander, 1986), implying the existence
of functional GH receptors in adipose tissue during late gestation. This could account for the
substantial decline in capacity for adipose tissue lipogenesis in fetal sheep during the last
month of gestation (Vernon et al., 1981).
Immaturity of the fetal GH/IGF-1 system raises the possibility that fetal protein anabolism
and tissue growth may be limited by low levels of circulating IGF-1, despite the generally
accepted notion that, during fetal life, the metabolic and trophic influences of locally
expressed IGF are more important than those of systemic IGF (see Jones and Clemmons,
1995). It is therefore notable that infusion of IGF-1 into fetal sheep decreased proteolysis and
amino acid catabolism (Harding et al., 1994; Liechty et al., 1996). Conversely, increased
amino acid catabolism in the undernourished sheep fetus (Hodgson et al., 1982; Lemons and
Schreiner, 1983) is associated with decreased plasma IGF-1 levels, whether due to maternal
nutrient deprivation (Bassett et al., 1990) or placental insufficiency (Owens et al., 1994).
In all species studied, fetal tissue expression and plasma levels of IGF-2 exceed those of
IGF-1 (Han et al., 1988; Mesiano et al., 1989; Lee et al., 1991; Delhanty and Han, 1993).
A special role for IGF-2 in the regulation of prenatal growth was demonstrated by initial gene
knockout studies in the mouse (DeChiara et al., 1991). Recently, tissue-specific gene inacti-
vation has been used to show that the IGF-2 gene is paternally imprinted in the placenta and
acts to promote placental growth and functional capacity, thereby influencing fetal nutrient
supply and growth in late gestation (fig. 4; Constancia et al., 2002). Lack of IGF-2 also
reduced fetal hepatic glycogen storage and glycemia, associated with decreased activity but
not mRNA abundance of glycogen synthase, and impaired the ability of newborn IGF-2
knockout mice to survive fasting for 12h (Lopez et al., 1999).

4.3.3. Placental lactogen

Placental lactogen (PL; also known as chorionic somatomammotropin) is a major, unique pro-
tein product of the placentae of ruminants, humans, rodents, and some other species. The
molecular identity and interspecies homology of these molecules, as well as their lactogenic and
somatogenic effects through their ability to bind to both GH and prolactin receptors, has been
reviewed recently (Gertler and Djiane, 2002). Ovine and bovine fetal plasma contains PL
throughout gestation (Anthony et al., 1995) and the effective half-life of circulating PL in fetal
sheep is similar to that of GH (Schoknecht et al., 1992). The physiological roles of this putative
regulator of fetal metabolism and growth remain to be established definitively. Glycogen
synthesis in isolated fetal hepatocytes was promoted by PL treatment in sheep (Freemark and
Handwerger, 1986) and rats (Freemark and Handwerger, 1984), and we observed a 56% increase
in hepatic glycogen accumulation in fetal sheep infused i.v. with native ovine PL for 14 days
(table 2; Schoknecht et al., 1996). In the latter study, PL treatment caused modest increases in
fetal plasma IGF-1 concentration and the relative weights of some visceral organs but did not
significantly affect fetal weight.

4.3.4. Glucocorticoids

In all species studied, there is a major increase in the circulating glucocorticoid concentration
in the fetus toward term, mostly due to a pronounced surge in fetal adrenal cortisol secretion.
The vital, pleiotropic influences of fetal cortisol on the structural and biochemical maturation
of multiple fetal tissues to prepare them for postnatal functions have been reviewed by
Fowden et al. (1998). Less is known about the effects of glucocorticoids on the regulation
Regulation of metabolism and growth during prenatal life 17

Fig. 4. Placental and fetal growth in mutant mice lacking paternal expression of the IGF-2 gene in
labyrinthine trophoblastic tissue of the placenta, and in their wild-type littermates. Significant differences
between wild-type and mutant mice are indicated: * P< 0.05; *** P<0.001. Adapted from the data of
Constancia et al. (2002).

of fetal metabolism in relation to growth and development. In general, fetal cortisol appears
to promote the availability of glucose to the neonatal animal by stimulating both hepatic
glycogen synthesis (Barnes et al., 1978) and maturation of the capacity for hepatic glucose
production (Townsend et al., 1991; Barbera et al., 1997) in the near-term sheep fetus. During
late gestation, treatment with glucocorticoids reduced umbilical glucose uptake (Milley, 1996;

Table 2
Effect of i.v. infusion of ovine placental lactogen for 14 days on liver glycogen concentration,
content in fetal sheep at day 136 of gestation (from Schoknecht et al., 1996)a

Parameter Controlb Placental lactogenc

Liver weight, g 115.8 ± 9.2 124.8 ± 9.9


Glycogen concentration, mg/g 79.3 ± 6.9 105.0 ± 5.6*
Glycogen content, g 8.4 ± 0.7 13.1 ± 1.7*
a Values are means ± SEM, n = 5.
b Infused with saline containing ovine plasma (15 ml/l), days 122 to 136 of gestation.
c Infused with ovine placental lactogen (1.2 mg/d), days 122 to 136 of gestation; caused a 4-fold increase in

fetal plasma concentration of placental lactogen.


* Treatment effect was significant at P<0.05.
18 A. W. Bell et al.

Barbera et al., 1997) and placental uptake of fetal glutamate (Barbera et al., 1997; Timmerman
et al., 2000). The latter response was associated with decreased hepatic output of glutamate
apparently due to decreased fetal hepatic uptake of glucogenic amino acids, including
glutamine, and diversion of hepatic glutamine to metabolism in the TCA cycle rather than
glutamate synthesis (Timmerman et al., 2000).
Other growth-related effects of the prepartum increase in fetal cortisol include induction
of the hepatic GH receptor and hepatic synthesis of IGF-1 (Li et al., 1996) but suppression
of IGF-1 expression in muscle independently of any change in GH receptor gene expression
(Li et al., 2002). Cortisol also suppresses IGF-2 expression in liver, muscle, and adrenal
glands (Li et al., 1993), stimulates the deiodination of thyroxine (T4) to triiodothyronine (T3)
(Fowden et al., 1998), and appears to downregulate the production of PL by binucleate cells
in the ovine placenta (Ward et al., 2002).

4.3.5. Thyroid hormones

The fetal thyroid secretes T4 from early in gestation, and thyroidectomy of fetal sheep in mid-
gestation causes generalized growth retardation and delayed maturation of the skin, skeleton,
and pulmonary and neuromuscular systems (Hopkins and Thorburn, 1972). Fetal sheep made
hypothyroid by thyroidectomy or hypophysectomy suffered a 20–30% decrease in umbilical
oxygen uptake that was restored to normal by exogenous T4 administration (Fowden and
Silver, 1995). This reduction in oxygen consumption was accompanied by abnormal blood-gas
status and reductions in rate of glucose oxidation and the fraction of oxygen consumption
used for glucose oxidation, all of which also were normalized by T4 replacement (Fowden and
Silver, 1995). Interestingly, plasma T3 levels remained low and were unchanged by thyroid or
pituitary ablation or exogenous T4, suggesting that at least before maturation of the enzymatic
capacity for T4 deiodination near term (Polk et al., 1988), thyroid hormone effects may be
mediated directly by T4. However, it should be noted that administration of T3 alone, albeit
in supraphysiological doses, caused an increase in oxygen consumption of fetal sheep (Lorijn
et al., 1980).
In addition to its negative effects on glucose utilization, thyroid deficiency impaired the
ability of fetal sheep to increase hepatic glucogenesis in response to fasting (Fowden et al.,
2001). Recent evidence also suggests that the cortisol-induced increase in deiodination of
T4 to T3, and the consequent prenatal surge in fetal plasma T3, at least partly mediates the
maturational effects of cortisol on the hepatic somatotropic axis (Forhead et al., 2000).

4.3.6. Catecholamines
Prenatal maturation of the sympathoadrenomedullary system is vital to enable the perinatal
animal to respond to the stresses of parturition and adaptation to the extrauterine environment.
In precocial species such as the sheep, central nervous (splanchnic) control of the adrenal
medulla develops relatively early; in other species, functional innervation is not apparent until
after birth (see Slotkin and Seidler, 1988). During late gestation, the fetal sheep responds to
acute hypoxia (Cohen et al., 1982) and hypoglycemia (Harwell et al., 1990) with pronounced
increases in adrenomedullary secretion of epinephrine and norepinephrine. Metabolic conse-
quences include rapid stimulation of hepatic glucose production, presumably through increased
glycogenolysis (Jones et al., 1983), and mobilization of NEFA (Harwell et al., 1990), associated
with reduced pancreatic secretion and plasma concentrations of insulin (Bassett and Hanson,
1998), and attenuated action of IGF-1 (Hooper et al., 1994). Restoration of normal insulinemia
Regulation of metabolism and growth during prenatal life 19

by insulin infusion abolished most of the metabolic and growth-inhibitory effects of prolonged
catecholamine infusion in the sheep fetus (Bassett and Hanson, 2000). This suggests that
establishment and maintenance of hypoinsulinemia is a necessary mediating factor for the
adverse effects of elevated circulating catecholamines on fetal growth.

4.3.7. Leptin

The peptide hormone leptin is synthesized and secreted primarily by adipose tissue in postnatal
animals and is considered to play an important role in the regulation of energy balance (Ahima
and Flier, 2000). Leptin has been detected in ovine fetal plasma as early as day 40 of gestation
(Ehrhardt et al., 2002) and its concentration increases moderately throughout gestation, espe-
cially during the last 2 weeks (Ehrhardt et al., 2002; Forhead et al., 2002). The gestational
increase in plasma concentration is accompanied by increased abundance of leptin mRNA in
perirenal adipose tissue in late pregnancy (Yuen et al., 1999). Our results indicate that before
100 days of gestation, tissues other than adipose tissue, such as brain and liver, are the primary
source of circulating leptin, and that this role is assumed by brown adipose tissue only after this
tissue develops appreciably during the last one-third of gestation (fig. 5; Ehrhardt et al., 2002).
Regulation of tissue expression and biological actions of leptin during fetal life have yet to
be studied systematically. The increase in plasma leptin during late gestation in fetal sheep was
associated with the prepartum surge in fetal cortisol and abolished by fetal adrenalectomy
(Forhead et al., 2002). It also appears that expression of leptin mRNA in perirenal brown adi-
pose tissue in the sheep fetus responds positively to hyperinsulinemia but not hyperglycemia
(Devaskar et al., 2002). The functional significance of fetal leptin is unclear. Leptin signaling
apparently is not essential during prenatal life because leptin-deficient ob/ob mice are born
relatively normal (Mounzih et al., 1998). Also, fetal plasma leptin was unaffected by changes in
maternal nutrition sufficient to change fetal glycemia and insulinemia in late-pregnant ewes
(Ehrhardt et al., 2002; Mühlhäusler et al., 2002; Yuen et al., 2002). Fetal plasma leptin was
correlated with body fatness as represented by the relative mass of unilocular cells in perirenal
and interscapular brown adipose tissue (Mühlhäusler et al., 2002). Infusion of leptin for several
days into the sheep fetus caused decreases in relative abundance of leptin mRNA and the pro-
portion of unilocular cells in perirenal adipose tissue, suggesting a feedback effect on adipose
tissue function (Yuen et al., 2003). However, the relevance of this observation is unclear because
of the unphysiologically high levels of plasma leptin in treated fetuses. A potential role for fetal
and/or maternal leptin in the regulation of placental function is suggested by the abundant
expression of the physiologically relevant long (Ob-Rb) form of the leptin receptor by the ovine
placenta (Ehrhardt et al., 1999; Thomas et al., 2001).

4.4. Coordination of fetal metabolism and growth

The mechanisms relating nutrient supply to expression of endocrine and local regulatory
factors and, thence, tissue metabolism and growth, can be illustrated by synthesis of the present
knowledge on IUGR, whether caused by placental insufficiency, maternal undernutrition, or
insulin-induced maternal hypoglycemia. Effects on the local expression of trophic factors and
the cellular growth of skeletal muscle will serve as an example of tissue responses to
an altered extracellular milieu. The putative relationships discussed below are schematically
represented in fig. 6.
Placental insufficiency during late gestation is generally characterized by fetal hypoxemia
and hypoglycemia, whether caused by surgical reduction (carunclectomy; Harding et al., 1985),
20 A. W. Bell et al.

Fig. 5. Relative abundance of leptin mRNA in ovine fetal tissues at different stages of gestation. Adapted
from the data of Ehrhardt et al. (2002).

placental embolization (Creasy et al., 1972), maternal heat stress (Bell et al., 1987b), or over-
feeding of adolescent ewes (Wallace et al., 2002). Associated endocrine changes include
decreased fetal plasma concentrations of insulin (Robinson et al., 1980) and IGF-1 and -2
(Owens et al., 1994), and increased concentrations of cortisol (Phillips et al., 1996). All of
these changes can be elicited by maternal undernutrition or insulin-induced hypoglycemia,
implicating fetal glycemia as an important primary signal (Mellor et al., 1977; Osgerby et al.,
2002). However, it must be recognized that hypoxemia may reinforce these responses through
its stimulation of fetal adrenal secretion of cortisol and catecholamines, and the inhibitory
influence of the latter on fetal insulin secretion.
It seems likely that hypoinsulinemia is a primary, coordinating mediator of the numerous
metabolic and trophic consequences of reduced fetal nutrient supply. Disruption of fetal pan-
creatic insulin secretion has a potent, negative effect on fetal growth (Fowden et al., 1995),
Regulation of metabolism and growth during prenatal life 21

Fig. 6. Schematic outline of some important factors linking maternal undernutrition and placental
insufficiency to intrauterine growth retardation.

associated with decreased fetal tissue uptake and metabolism of glucose (Fowden and Hay,
1988), decreased uptake of amino acids and increased proteolysis (Carver et al., 1997), and
reduced circulating levels of IGF-1 (Gluckman et al., 1987).
However, although circulating IGF-1 may be of increasing importance during late gesta-
tion, it is likely that local tissue expression and actions of this and other growth factors are
more significant mediators of tissue growth responses to altered nutrient supply. For exam-
ple, fetal muscle strongly expresses IGF-1 throughout gestation (Dickson et al., 1991; Lee
et al., 1993) and disruption of the IGF-1 gene causes lethal abnormalities in muscle develop-
ment (Liu et al., 1993), consistent with the extensive evidence for the role of IGF-1 in
regulation of myogenesis (Florini et al., 1996). It therefore seems likely that the reduced
mitotic activity of myosatellite cells and growth of skeletal muscle in acutely undernourished
or placentally growth-retarded sheep fetuses (Greenwood et al., 1999) was mediated, at least
partly, by reduced local expression of IGF-1, possibly caused by elevated plasma levels of
cortisol (Li et al., 2002).
Finally, although this section has focused on IUGR to illustrate aspects of the coordination
of nutrient supply with growth in the fetus, it should be recognized that even in optimally fed,
healthy animals, fetal growth is constrained by placental capacity for nutrient transfer during
late pregnancy. This phenomenon ensures that the unborn animal’s demands upon its
dam’s nutrient reserves are not excessive, and reduces the possibility of birth injury to itself
22 A. W. Bell et al.

and its mother. The capacity for increased growth in response to increased nutrient supply was
demonstrated by the almost 20% increase in birth weight of singleton lambs that had been
infused directly with glucose for the last 30 days of gestation, in ewes that were extremely
well fed (Stevens et al., 1990).

5. INFLUENCE OF PRENATAL METABOLISM AND GROWTH


ON POSTNATAL PERFORMANCE AND HEALTH
5.1. Postpartum metabolism and growth

We recently have reported some of the metabolic characteristics of naturally growth-retarded


lambs from prolific ewes immediately after birth and during neonatal growth to a nominal live
weight (LW) of 20 kg (Greenwood et al., 2002; Greenwood and Bell, 2003). At birth, these
lambs tended to be hypoglycemic and had elevated plasma urea nitrogen levels. More strik-
ing was the apparent immaturity of their hepatic GH/IGF system as represented by greatly
elevated plasma concentrations of GH and low concentrations of IGF-1. This blood picture
was associated with reduced hepatic expression of the GH receptor and the GH-dependent
ALS necessary to form the ternary binding complex which contains most circulating IGF-1
in postnatal life (Rhoads et al., 2000a,b).
Postnatal changes in superficial indices of carbohydrate and protein metabolism were little
affected by birth weight in small and normal lambs that were artificially reared with ad libitum
access to milk replacer. The very high concentrations of plasma GH in small, newborn lambs
decreased markedly within 2 days of birth but remained significantly higher than concentra-
tions in normal lambs for about 2 weeks. During the same period, plasma IGF-1 increased
steadily in both groups but remained significantly lower in the small lambs (Greenwood et al.,
2002). These observations suggest that the apparent immaturity of the GH/IGF axis in growth-
retarded newborn lambs persists for several weeks after birth. Interestingly, only during this
early postnatal phase did the absolute growth rates of low-birth-weight lambs (248 g/d) lag
significantly behind those of normal birth weight lambs (353 g/d) (Greenwood et al., 1998).
Thereafter, during rapid growth from about 2 weeks of age to slaughter at 20 kg (attained at
6.5–8 weeks of age), plasma IGF-1 concentrations were persistently higher but GH concen-
trations were not different in low-versus normal-birth-weight lambs, perhaps related to the
higher relative energy intakes and plasma insulin concentrations (see below) of the small lambs.
This study did not examine the consequences of low birth weight after weaning. However,
plasma GH concentrations tended to be higher during adolescence (~132 days of age) and
adulthood (~378 days of age) in low-birth-weight male lambs from carunclectomized ewes
compared to lambs of normal birth weight and were negatively correlated with indices of
birth size (Gatford et al., 2002).
Plasma insulin concentrations increased rapidly during the early postnatal period in small
lambs feeding ad libitum, consistent with their very high levels of energy intake. Then, from
about 2 weeks of age until slaughter at 20 kg, plasma insulin concentrations were persistently
higher in low-compared with normal-birth-weight lambs. We speculate that this relative
hyperinsulinemia may be due to the predisposition of growth-retarded neonates to develop
insulin resistance (Hales et al., 1996).
Plasma leptin concentrations were somewhat higher in rapidly fattening, low-birth-weight
lambs during the first week post partum, but not thereafter (Ehrhardt et al., 2003), despite the
fact that at any subsequent body weight up to 20 kg LW, these lambs were significantly fatter
than their normal-birth-weight counterparts (Greenwood et al., 1998).
Regulation of metabolism and growth during prenatal life 23

Additional aspects of whole-body and tissue growth and development in lambs suffering
IUGR are discussed in another, recent review (Greenwood and Bell, 2003).

5.2. Fetal programming of postnatal pathophysiology

The human epidemiological evidence for fetal programming has implicated IUGR as an
important risk factor for mature onset of diseases including hypertension and type II diabetes
(Barker, 1998). Although the methodology and interpretation of aspects of this work recently
have been challenged (Huxley et al., 2002), these epidemiological associations have been
replicated experimentally in rodents (Langley-Evans, 2001) and various models of IUGR in
sheep (McMillen et al., 2001; Greenwood and Bell, 2003). For example, low-birth-weight
offspring born to protein-deprived rats (Langley and Jackson, 1994) and placentally insuffi-
cient ewes (McMillen et al., 2001) display hypertension during postweaning growth and
adulthood. The rat model also has been used to demonstrate a relation between prenatal nutri-
tion and the later development of insulin resistance (Langley et al., 1994), and similar
evidence is emerging from sheep studies (Greenwood and Bell, 2003).
Mechanisms linking prenatal nutrition, organ and tissue development, and the programming of
later pathophysiology are unclear. However, excessive fetal exposure to glucocorticoids is a con-
sistent feature of most animal studies involving prenatal nutrient deprivation, especially during
late gestation. Also, treatment of pregnant rats and sheep with glucocorticoids during late preg-
nancy can replicate some of the programming effects of fetal undernutrition on later development
of hypertension and insulin resistance (Langley-Evans, 2001; Greenwood and Bell, 2003).
Growing evidence from studies on sheep and other species indicates that fetal programming
can involve long-term sequelae to changes in the early prenatal environment that do not neces-
sarily cause changes in fetal gross morphology. For example, modest undernutrition of ewes
during the first half of pregnancy had no effect on growth of lambs during fetal or postnatal life
but caused relative hypertension and increased activity of the hypothalamic–pituitary–adrenal
(HPA) axis in lambs aged 12–13 weeks (Hawkins et al., 2000). Consistent with these responses,
maternal undernutrition between early and mid-gestation caused increased expression of the
glucocorticoid receptor in adrenals, kidney, liver, lungs, and perirenal adipose tissue of the fetus
at term (~145 days) (Whorwood et al., 2001). At the same time, there were marked changes in
the enzymatic capacity of several fetal tissues to deactivate cortisol, which may have led to
excessive fetal exposure to this hormone during late gestation. Some of these tissue-specific
fetal responses were evident as early as day 77 of gestation.
A central role for corticosteroids in the mediation of fetal programming was further impli-
cated by the remarkable finding that exposure of ewes to high doses of dexamethasone for
only 2 days in early pregnancy resulted in hypertensive offspring at 3–4 months of age (Dodic
et al., 1998). This hypertension amplified with age to beyond 3 years and was associated with
increased cardiac output (Dodic et al., 1999) but no change in responsiveness of the HPA axis
(Dodic et al., 2002). Glucose metabolic responses to insulin were unaltered, but the ability of
insulin to suppress net fatty acid release from adipose tissue (plasma nonesterified fatty acid
concentration) was moderately enhanced (Gatford et al., 2000).

6. FUTURE PERSPECTIVES
The development almost four decades ago of novel techniques for studying fetal and placental
physiology and metabolism in utero has led to considerable understanding of the regulation
24 A. W. Bell et al.

of the metabolic and developmental processes that culminate in the birth of a healthy, well-
grown neonate. Nevertheless, unexplained dysfunctions of conceptus growth remain,
associated with unacceptable incidence of perinatal morbidity and mortality in many human
and domestic animal populations. There also is a new awareness of the possible longer-term
effects of nutritional and other environmental insults during fetal life, some of which may be
quite subtle and without influence on gross morphology. Unraveling the mechanisms under-
lying such effects will be the major challenge of prenatal biology for the foreseeable future
and should lead to a greater understanding of both human mature-onset pathologies and
unexplained variation in the productivity and disease resistance of domestic animals.

REFERENCES
Ahima, R.S., Flier, J.S., 2000. Leptin. Annu. Rev. Physiol. 62, 413–437.
Aldoretta, P.W., Carver, T.D., Hay, W.W. Jr., 1998. Maturation of glucose-stimulated insulin secretion in
fetal sheep. Biol. Neonate 73, 375–396.
Aldoretta, P.W., Hay, W.W. Jr., 1999. Effect of glucose supply on ovine uteroplacental glucose metabo-
lism. Amer. J. Physiol. 277, R947–R958.
Alexander, D.P., Assan, R., Britton, H.G., Fenton, E., Redstone, D., 1976. Glucagon release in the sheep
fetus I. Effect of hypo- and hyperglycemia and arginine. Biol. Neonate 30, 1–10.
Alexander, D.P, Britton, H.G., Cohen, N.M., Nixon, D.A., Parker, R.A., 1968. Insulin concentrations in
the foetal plasma and foetal fluids of the sheep. J. Endocrinol. 40, 389–390.
Alexander, G., 1964. Studies on the placenta of the sheep (Ovis aries L.): effect of reduction in the
number of caruncles. J. Reprod. Fertil. 7, 307–322.
Alexander, G., Williams, D., 1971. Heat stress and development of the conceptus in domestic sheep.
J. Agr. Sci. 76, 53–72.
Anderson, A.H., Fennessey, P.V., Meschia, G., Wilkening, R.B., Battaglia, F.C., 1997. Placental transport
of threonine and its utilization in the normal and growth-restricted fetus. Amer. J. Physiol. 272,
E892–E900.
Anderson, M.S., Flowers-Ziegler, J., Das, U.G., Hay, W.W. Jr., Devaskar, S.U., 2001a. Glucose trans-
porter protein responses to selective hyperglycemia or hyperinsulinemia in fetal sheep. Amer.
J. Physiol. 281, R1545–R1552.
Anderson, M.S., He, J., Flowers-Ziegler, J., Devaskar, S.U., Hay, W.W. Jr., 2001b. Effects of selective
hyperglycemia and hyperinsulinemia on glucose transporters in fetal ovine skeletal muscle. Amer.
J. Physiol. 281, R1256–R1263.
Anthony, R.V., Pratt, S.L., Liang, R., Holland, M.D., 1995. Placental-fetal hormonal interactions: impact
on fetal growth. J. Anim. Sci. 73, 1861–1871.
Barbera, A., Wilkening, R.B., Teng, C., Battaglia, F.C., Meschia, G., 1997. Metabolic alterations in the
fetal hepatic and umbilical circulations during glucocorticoid-induced parturition in sheep. Pediat.
Res. 41, 242–248.
Barker, D.J.P., 1998. Mothers, Babies and Health in Later Life, 2nd Edition. Churchill Livingstone,
Edinburgh.
Barnes, R.J., Comline, R.S., Silver, M., 1978. Effect of cortisol on liver glycogen concentrations in
hypophysectomized, adrenalectomized and normal foetal lambs during late or prolonged gestation.
J. Physiol. 275, 567–579.
Bassett, J.M., Hanson, C., 1998. Catecholamines inhibit growth in fetal sheep in the absence of hypox-
emia. Amer. J. Physiol. 274, R1536–R1545.
Bassett, J.M., Hanson, C., 2000. Prevention of hypoinsulinemia modifies catecholamine effects in fetal
sheep. Amer. J. Physiol. 278, R1171–R1181.
Regulation of metabolism and growth during prenatal life 25

Bassett, J.M., Thorburn, G.D., Wallace, A.L.C., 1970. The plasma growth hormone concentration of the
foetal lamb. J. Endocrinol. 48, 251–263.
Bassett, N.S., Oliver, M.H., Breier, B.H., Gluckman, P.D., 1990. The effect of maternal starvation on plasma
insulin-like growth factor I concentrations in the late gestation ovine fetus. Pediat. Res. 27, 401–404.
Battaglia, F.C., Meschia, G., 1988. Fetal nutrition. Annu. Rev. Nutr. 8, 43–61.
Battaglia, F.C., Regnault, T.R.H., 2001. Placental transport and metabolism of amino acids. Placenta 22,
145–161.
Bauer, M.K., Harding, J.E., Breier, B.H., Gluckman, P.D., 2000. Exogenous GH infusion to late-gestational
fetal sheep does not alter fetal growth and metabolism. J. Endocrinol. 166, 591–597.
Bell, A.W., 1993. Pregnancy and fetal metabolism. In: Forbes, J.M., France, J. (Eds.), Quantitative
Aspects of Ruminant Digestion and Metabolism. CAB International, Oxford, pp. 406–431.
Bell, A.W., Battaglia, F.C., Meschia, G., 1987a. Relation between metabolic rate and body size in the
ovine fetus. J. Nutr. 117, 1181–1186.
Bell, A.W., Ehrhardt, R.A., 2002. Regulation of placental nutrient transport and implications for fetal
growth. Nutr. Res. Rev. 15, 211–230.
Bell, A.W., Hay, W.W. Jr., Ehrhardt, R.A., 1999. Placental transport of nutrients and its implications for
fetal growth. J. Reprod. Fertil., Suppl. 54, 401–410.
Bell, A.W., Kennaugh, J.M., Battaglia, F.C., Makowski, E.L., Meschia, G., 1986. Metabolic and circula-
tory studies of fetal lamb at midgestation. Amer. J. Physiol. 250, E538–E544.
Bell, A.W., Kennaugh, J.M., Battaglia, F.C., Meschia, G., 1989. Uptake of amino acids and ammonia at
mid-gestation by the fetal lamb. Quart. J. Exp. Physiol. 74, 635–643.
Bell, A.W., Slepetis, R., Ehrhardt, R.A., 1995. Growth and accretion of energy and protein in the gravid
uterus during late pregnancy in Holstein cows. J. Dairy Sci. 78, 1954–1961.
Bell, A.W., Wilkening, R.B., Meschia, G., 1987b. Some aspects of placental function in chronically heat-
stressed ewes. J. Dev. Physiol. 9, 17–29.
Boisclair, Y.R., Rhoads, R.P., Ueki, I., Wang, J., Ooi, G.T., 2001. The acid-labile subunit (ALS) of the
150 kDa IGF-binding protein complex: an important but forgotten component of the circulating IGF
system. J. Endocrinol. 170, 63–70.
Campbell, F.M., Gordon, M.J., Dutta-Roy, A.K., 1994. Plasma membrane fatty acid-binding protein
(FABPpm) from the sheep placenta. Biochim. Biophys. Acta 1214, 187–192.
Campbell, F.M., Taffesse, S., Gordon, M.J., Dutta-Roy, A.K., 1995. Plasma membrane fatty acid-binding
protein from human placenta: identification and characterisation. Biochem. Biophys. Res. Comm.
209, 1011–1017.
Carter, B.S., Moores, R.R., Jr., Teng, C., Meschia, G., Battaglia, F.C., 1995. Main routes of plasma lac-
tate carbon disposal in the midgestation fetal lamb. Biol. Neonate 67, 295–300.
Carver, T.D., Quick, A.A., Teng, C., Pike, A.W., Fennessey, P.V., Hay, W.W. Jr., 1997. Leucine metabo-
lism in chronically hypoglycemic hypoinsulinemic growth-restricted fetal sheep. Amer. J. Physiol.
272, E107–E117.
Char, V.C., Creasy, R.K., 1976. Acetate as a metabolic substrate in the fetal lamb. Amer. J. Physiol. 230,
357–361.
Chung, M., Teng, C., Timmerman, M., Meschia, G., Battaglia, F.C., 1998. Production and utilization of
amino acids by ovine placenta in vivo. Amer. J. Physiol. 274, E13–E22.
Clegg, R.A., 1981. Placental lipoprotein lipase activity in the rabbit, rat and sheep. Comp. Biochem.
Physiol. 69B, 585–591.
Cohen, W.R., Piasecki, G.J., Jackson, B.T., 1982. Plasma catecholamines during hypoxemia in fetal lamb.
Amer. J. Physiol. 243, R520–R525.
Comline, R.S., Silver, M., 1976. Some aspects of foetal and utero-placental metabolism in cows with
indwelling umbilical and uterine vascular catheters. J. Physiol. 260, 571–586.
26 A. W. Bell et al.

Constancia, M., Hemberger, M., Hughes, J., Dean, W., Ferguson-Smith, A., Fundeles, R., Stewart, F.,
Kelsey, G., Fowden, A., Sibley, C., Reik, W., 2002. Placental-specific IGF-II is a major modulator of
placental and fetal growth. Nature 417, 945–948.
Creasy, R.K., Barrett, C.T., de Swiet, M., Kahanpää, K.V., Rudolph, A.M., 1972. Experimental intrauterine
growth retardation in the sheep. Amer. J. Obstet. Gynecol. 112, 566–573.
Currie, M.J., Bassett, N.S., Gluckman, P.D., 1997. Ovine glucose transporter-1 and -3: cDNA partial
sequences and developmental gene expression in the placenta. Placenta 18, 393–401.
D’Agostino, J., Field, J.B., Frazier, M.L., 1985. Ontogeny of immunoreactive insulin in the fetal bovine
pancreas. Endocrinology 116, 1108–1116.
Das, U.G., He, J., Ehrhardt, R.A., Hay, W.W. Jr., Devaskar, S.U., 2000. Time-dependent physiological reg-
ulation of ovine placental GLUT-3 glucose transporter protein. Amer. J. Physiol. 279, R2252–R2261.
Das, U.G., Sadiq, H.F., Soares, M.J., Hay, W.W. Jr., Devaskar, S.U., 1998. Time-dependent physiological
regulation of rodent and ovine placental glucose transporter (GLUT-1) protein. Amer. J. Physiol. 274,
R339–R347.
DeChiara, T.M., Robertson, E.J., Efstratiadis, A., 1991. Parental imprinting of the mouse insulin-like
growth factor II gene. Cell 64, 849–859.
Delhanty, P.J., Han, V.K., 1993. The expression of insulin-like growth factor (IGF)-binding protein-2 and
IGF-II genes in the tissues of the developing ovine fetus. Endocrinology 132, 41–52.
Devaskar, S.U., Anthony, R., Hay, W. Jr., 2002. Ontogeny and insulin regulation of fetal ovine white
adipose tissue leptin expression. Amer. J. Physiol. 282, R431–R438.
de Zegher, F., Styne, D.M., Daaboul, J., Bettendorf, M., Kaplan, S.L., Grumbach, M.M., 1989. Hormone
ontogeny in the ovine fetus and neonatal lamb. XX. Effect of age, breeding season, and twinning on
the growth hormone (GH) response to GH-releasing factor: evidence for a homeostatic role of fetal
GH. Endocrinology 124, 124–128.
Dickson, M.C., Saunders, J.C., Gilmour, R.S., 1991. The ovine insulin-like growth factor-I gene:
characterization, expression and identification of a putative promoter. J. Mol. Endocrinol. 6, 17–31.
DiGiacomo, J.E., Hay, W.W. Jr., 1990a. Placental-fetal glucose exchange and placental glucose
consumption in pregnant sheep. Amer. J. Physiol. 258, E360–E367.
DiGiacomo, J.E., Hay, W.W. Jr., 1990b. Fetal glucose metabolism and oxygen consumption during
sustained hypoglycemia. Metabolism 39, 193–202.
Dodic, M., May, C.N., Wintour, E.M., Coghlan, J.P., 1998. An early prenatal exposure to excess gluco-
corticoid leads to hypertensive offspring in sheep. Clin. Sci. 94, 149–155.
Dodic, M., Peers, A., Coghlan, J.P., May, C.N., Lumbers, E., Yu, Z.-Y., Wintour, E.M., 1999. Altered car-
diovascular haemodynamics and baroreceptor-heart rate reflex in adult sheep after prenatal exposure
to dexamethasone. Clin. Sci. 97, 103–109.
Dodic, M., Peers, A., Moritz, K., Hantzis, V., Wintour, E.M., 2002. No evidence for HPA reset in adult
sheep with high blood pressure due to short prenatal exposure to dexamethasone. Amer. J. Physiol.
282, R343–R350.
Ehrhardt, R.A., Bell, A.W., 1995. Growth and metabolism of the ovine placenta during mid-gestation.
Placenta 16, 727–741.
Ehrhardt, R.A., Bell, A.W., 1997. Developmental increases in glucose transporter concentration in the
sheep placenta. Amer. J. Physiol. 273, R1132–R1141.
Ehrhardt, R.A., Bell, A.W., Boisclair, Y.R., 1999. Analysis of leptin and leptin receptor mRNA expres-
sion in the sheep placenta. J. Anim. Sci., 77, Suppl. 1, 159.
Ehrhardt, R.A., Bell, A.W., Boisclair, Y.R., 2002. Spatial and developmental regulation of leptin in fetal
sheep. Amer. J. Physiol. 282, R1628–R1635.
Ehrhardt, R.A., Boston, R.C., Slepetis, R.M., Bell, A.W., 1996. Moderate maternal undernutrition
elevates placental glucose transport capacity in sheep. J. Anim. Sci. 74, Suppl. 1, 152.
Regulation of metabolism and growth during prenatal life 27

Ehrhardt, R.A., Greenwood, P.L., Bell, A.W., Boisclair, Y.R. 2003. Plasma leptin is regulated predomi-
nantly by nutrition in preruminant lambs. J. Nutr. 133, 4196–4201.
Ehrhardt, R.A., Slepetis, R.M., Bell, A.W., 1998. Moderate maternal undernutrition alters glucose
transporter levels in maternal insulin-responsive tissues and in the placenta. J. Anim. Sci. 76,
Suppl. 1, 130.
Etherton, T.D., Bauman, D.E., 1998. Biology of somatotropin in growth and lactation of domestic
animals. Physiol. Rev. 78, 745–761.
Faichney, G.J., White, G.A., 1987. Effects of maternal nutritional status on fetal and placental growth and
on fetal urea synthesis in sheep. Aust. J. Biol. Sci. 40, 365–377.
Ferrell, C.L., 1991. Maternal and fetal influences on uterine and conceptus development in the cow: II.
Blood flow and nutrient flux. J. Anim. Sci. 69, 1954–1965.
Ferrell, C.L., Garrett, W.N., Hinman, N., 1976. Growth, development and composition of the udder and
gravid uterus of beef heifers during pregnancy. J. Anim. Sci. 42, 1477–1489.
Florini, J.R., Ewton, D.Z., Coolican, S.A., 1996. Growth hormone and insulin-like growth factor system
in myogenesis. Endocr. Rev. 17, 481–517.
Forhead, A.J., Li, J., Saunders, J.C., Dauncey, M.J., Gilmour, R.S., Fowden, A.L., 2000. Control of ovine
hepatic growth hormone receptor and insulin-like growth factor I by thyroid hormones in utero. Amer.
J. Physiol. 278, E1166–E1174.
Forhead, A.J., Thomas, L., Crabtree, J., Hoggard, N., Gardner, D.S., Giussani, D.A., Fowden, A.L., 2002.
Plasma leptin concentration in fetal sheep during late gestation: ontogeny and effect of glucocorti-
coids. Endocrinology 143, 1166–1173.
Fowden, A.L., 1997. Comparative aspects of fetal carbohydrate metabolism. Equine Vet. J. 24, 19–25.
Fowden, A.L., Apatu, R.S.K., Silver, M., 1995. The glucogenic capacity of the fetal pig: developmental
regulation by cortisol. Exp. Physiol. 80, 457–467.
Fowden, A.L., Forhead, A.J., Silver, M., Macdonald, A.A., 1997. Glucose, lactate and oxygen metabo-
lism in the fetal pig during late gestation. Exp. Physiol. 82, 171–182.
Fowden, A.L., Forhead, White, K.L., Taylor, P.M., 2000b. Equine uteroplacental metabolism at mid- and
late gestation. Exp. Physiol. 85, 539–545.
Fowden, A.L., Hay, W.W. Jr. 1988. The effects of pancreatectomy on the rates of glucose utilization,
oxidation and production in the sheep fetus. Quart. J. Exp. Physiol. 73, 973–984.
Fowden, A.L., Li, J., Forhead, A.L., 1998. Glucocorticoids and the preparation for life after birth: are
there long-term consequences of the life insurance? Proc. Nutr. Soc. 57, 113–122.
Fowden, A.L., Mapstone, J., Forhead, A.J., 2001. Regulation of glucogenesis by thyroid hormones in
fetal sheep during late gestation. J. Endocrinol. 170, 461–469.
Fowden, A.L., Silver, M., 1995. The effect of thyroid hormones on oxygen and glucose metabolism in
the sheep fetus during late gestation. J. Physiol. 482, 203–213.
Fowden, A.L., Taylor, P.M., White, K.L., Forhead, A.J., 2000a. Ontogenic and nutritionally induced
changes in fetal metabolism in the horse. J. Physiol. 528, 209–219.
Freemark, M., Handwerger, S., 1984. Ovine placental lactogen stimulates glycogen synthesis in fetal rat
hepatocytes. Amer. J. Physiol. 246, E21–E24.
Freemark, M., Handwerger, S., 1986. The glycogenic effects of placental lactogen and growth hormone
in ovine fetal liver are mediated through binding to specific fetal ovine placental lactogen receptors.
Endocrinology 118, 613–618.
Gatford, K.L., Clarke, I.J., De Blasio, M.J., McMillen, I.C., Robinson, J.S., Owens, J.A., 2002. Perinatal
growth and plasma GH profiles in adolescent and adult sheep. J. Endocrinol. 173, 151–159.
Gatford, K.L., Wintour, E.M., De Blasio, M.J., Owens, J.A., Dodic, M., 2000. Differential timing for pro-
gramming of glucose homeostasis, sensitivity to insulin and blood pressure by in utero exposure to
dexamethasone in sheep. Clin. Sci. 98, 553–560.
28 A. W. Bell et al.

Gertler, A., Djiane, J., 2002. Mechanism of ruminant placental lactogen action: molecular and in vivo
studies. Mol. Genet. Metab. 75, 189–201.
Gluckman, P.D., Butler, J.H., Comline, R., Fowden, A., 1987. The effects of pancreatectomy on the plasma
concentrations of insulin-like growth factors 1 and 2 in the sheep fetus. J. Dev. Physiol. 9, 79–88.
Gluckman, P.D., Mueller, P.L., Kaplan, S.L., Rudolph, A.M., Grumbach, M.M., 1979. Hormone
ontogeny in the ovine fetus. I. Circulating growth hormone in mid and late gestation. Endocrinology
104, 162–168.
Goalstone, M.L., Leitner, J.W., Wall, K., Dolgonos, L., Rothen, K.I., Accili, D., Draznin, B., 1998. Effect
of insulin on farnesyltransferase: specificity of insulin action and potentiation of nuclear effects of
insulin-like growth factor-1, epidermal growth factor, and platelet-derived growth factor. J. Biol.
Chem. 273, 23892–23896.
Grajwer, L.A., Sperling, M.A., Sack, J., Fisher, D.A., 1977. Possible mechanisms and significance of the
neonatal surge in glucagon secretion: studies in newborn lambs. Pediat. Res. 11, 833–836.
Greenwood, P.L, Bell, A.W., 2003. Consequences of intrauterine growth retardation for postnatal growth,
metabolism and pathophysiology. Reproduction, Suppl. 61, 195–206.
Greenwood, P.L., Hunt A.S., Hermanson, J.W., Bell, A.W., 1998. Effects of birth weight and postnatal
nutrition on neonatal sheep: I. Body growth and composition, and some aspects of energetic
efficiency. J. Anim. Sci. 76, 2354–2367.
Greenwood, P.L., Hunt, A.S., Slepetis, R.M., Finnerty, K.D., Alston, C., Beermann, D.H., Bell, A.W.,
2002. Effects of birth weight and postnatal nutrition on neonatal sheep: III. Regulation of energy
metabolism. J. Anim. Sci. 80, 2850–2861.
Greenwood, P.L., Slepetis, R.M., Bell, A.W., 2000. Influences on fetal and placental weights during mid to
late gestation in prolific ewes well nourished throughout pregnancy. Reprod. Fertil. Dev. 12, 149–156.
Greenwood, P.L., Slepetis, R.M., Hermanson, J.W., Bell, A.W., 1999. Intrauterine growth retardation is
associated with reduced cell cycle activity, but not myofibre number, in ovine fetal muscle. Reprod.
Fertil. Dev. 11, 281–291.
Hales, C.N., Desai, M., Ozanne, S.E., Crowther, N.J., 1996. Fishing in the stream of diabetes: from meas-
uring insulin to the control of fetal organogenesis. Biochem. Soc. Trans. 24, 341–350.
Han, V.K., Lund, P.K., Lee, D.C., D’Ercole, A.J., 1988. Expression of somatomedin/insulin-like growth
factor messenger ribonucleic acids in the human fetus: identification, characterization, and tissue
distribution. J. Clin. Endocrinol. Metab. 66, 422–429.
Harding, J.E., Jones, C.T., Robinson, J.S., 1985. Studies on experimental growth retardation in sheep: the
effects of a small placenta in restricting transport to and growth of the fetus. J. Dev. Physiol. 7, 427–442.
Harding, J.E., Liu, L., Evans, P.C., Gluckman, P.D., 1994. Insulin-like growth factor I alters feto-placental
protein and carbohydrate metabolism in fetal sheep. Endocrinology 134, 1509–1514.
Harwell, C.M., Padbury, J.F., Anand, R.S., Martinez, A.M., Ipp, E., Thio, L., Burnell, E.E., 1990. Fetal
catecholamine responses to maternal hypoglycemia. Amer. J. Physiol. 259, R1126–R1130.
Hawkins, P., Steyn, C., McGarrigle, H.H.G., Calder, N.A., Saito, T., Stratford, L.L., Noakes, D.E.,
Hanson, M.A., 2000. Cardiovascular and hypothalamic-pituitary-adrenal axis development in late
gestation fetal sheep and young lambs following modest maternal nutrient restriction in early gestation.
Reprod. Fertil. Dev. 12, 443–456.
Hay, W.W. Jr., 1995. Regulation of placental metabolism by glucose supply. Reprod. Fertil. Dev. 7, 365–375.
Hay, W.W., 1998. Fetal requirements, placental transfer of nitrogenous compounds. In: Polin, R.A., Fox,
W.W (Eds.), Fetal, Neonatal Physiology, 2nd Edition. Saunders, Philadelphia, pp. 619–634.
Hay, W.W. Jr., Meznarich, H.K., 1986. The effect of hyperinsulinaemia on glucose utilization and oxidation
and on oxygen consumption in the fetal lamb. Quart. J. Exp. Physiol. 71, 689–698.
Hay, W.W. Jr., Meznarich, H.K., 1988. Effect of maternal glucose concentration on uteroplacental glucose
consumption and transfer in pregnant sheep. Proc. Soc. Exp. Biol. Med. 190, 63–69.
Regulation of metabolism and growth during prenatal life 29

Hay, W.W. Jr., Meznarich, H.K., DiGiacomo, J.E., Hirst, K., Zerbe, G., 1988. Effects of insulin and glucose
concentrations on glucose utilization in fetal sheep. Pediat. Res. 23, 381–387.
Hay, W.W. Jr., Molina, R.D., DiGiacomo, J.E., Meschia, G., 1990. Model of placental glucose consump-
tion and glucose transfer. Amer. J. Physiol. 258, R569–R577.
Hay, W.W. Jr., Myers, S.A., Sparks, J.W., Wilkening, R.B., Meschia, G., Battaglia, F.C., 1983. Glucose
and lactate oxidation rates in the fetal lamb. Proc. Soc. Exp. Biol. Med. 173, 553–563.
Hay, W.W. Jr., Sparks, J.W., Wilkening, R.B., Battaglia, F.C., Meschia, G., 1984. Fetal glucose
uptake and utilization as functions of maternal glucose concentration. Amer. J. Physiol. 246,
E237–EE242.
Hodgson, J.C., Mellor, D.J., Field, A.C., 1982. Foetal and maternal rates of urea production and disposal
in well-nourished and undernourished sheep. Brit. J. Nutr. 48, 49–58.
Holzman, I.R., Lemons, J.A., Meschia, G., Battaglia, F.C., 1977. Ammonia production by the pregnant
uterus. Proc. Soc. Exp. Biol. Med. 156, 27–30.
Hooper, S.B., Bocking, A.D., White, S.E., Fraher, L.J., McDonald, T.J., Han, V.K., 1994. Catecholamines
stimulate the synthesis and release of insulin-like growth factor binding protein-1 (IGFBP-1) by fetal
sheep liver in vivo. Endocrinology 134, 1104–1112.
Hopkins, P.S., Thorburn, G.D., 1972. The effects of fetal thyroidectomy on the development of the ovine
fetus. J. Endocrinol. 54, 55–66.
Horn, J., Stern, M.D.R., Young, M., Noakes, D.E., 1983. Influence of insulin and substrate concentration
on protein synthetic rate in fetal tissues. Res. Vet. Sci. 35, 35–41.
Huxley, R., Neil, A., Collins, R., 2002. Unravelling the fetal origins hypothesis: is there really an inverse
association between birthweight and subsequent blood pressure? Lancet 360, 659–665.
Illsley, N.P., 2000. Glucose transporters in the human placenta. Placenta 21, 14–22.
Jodarski, G.D., Shanahan, M.F., Rankin, J.H.G., 1985. Fetal insulin and placental 3-O-methyl glucose
clearance in near-term sheep. J. Dev. Physiol. 7, 251–258.
Jones, C.T., Ritchie, J.W., Walker, D., 1983. The effects of hypoxia on glucose turnover in the fetal sheep.
J. Dev. Physiol. 5, 223–235.
Jones, J.I., Clemmons, D.R., 1995. Insulin-like growth factors and their binding proteins: biological
actions. Endocr. Rev. 16, 3–34.
Jozwik, M., Teng, C., Battaglia, F.C., Meschia, G., 1999. Fetal supply of amino acids and amino
nitrogen after maternal infusion of amino acids in pregnant sheep. Amer. J. Obstet. Gynecol. 180,
447–453.
Jozwik, M., Teng, C., Timmerman, M., Chung, M., Meschia, G., Battaglia, F.C., 1998. Uptake and trans-
port by the ovine placenta of neutral nonmetabolizable amino acids with different transport system
affinities. Placenta 19, 531–538.
Kennaugh, J.M., Bell, A.W., Teng, C., Meschia, G., Battaglia, F.C., 1987. Ontogenetic changes in the
rates of protein synthesis and leucine oxidation during fetal life. Pediat. Res. 22, 688–692.
Klempt, M., Bingham, B., Breier, B.H., Baumbach, W.R., Gluckman, P.D., 1993. Tissue distribution
and ontogeny of growth hormone receptor messenger ribonucleic acid and ligand binding to hepatic
tissue in the midgestation sheep fetus. Endocrinology 132, 1071–1077.
Krishnamurti, C.R., Schaefer, A.L., 1984. Effect of acute maternal starvation on tyrosine metabolism and
protein synthesis in fetal sheep. Growth 48, 391–403.
Langley, S.C., Browne, R.F., Jackson, A.A., 1994. Altered glucose tolerance in rats exposed to maternal
low protein diets in utero. Comp. Biochem. Physiol. 109A, 223–229.
Langley, S.C., Jackson, A.A., 1994. Increased systolic blood pressure in adult rats induced by fetal
exposure to maternal low protein diet. Clin. Sci. 86, 217–222.
Langley-Evans, S.C., 2001. Fetal programming of cardiovascular function through exposure to maternal
undernutrition. Proc. Nutr. Soc. 60, 505–513.
30 A. W. Bell et al.

Lee, C.Y., Bazer, F.W., Etherton, T.D., Simmen, F.H., 1991. Ontogeny of insulin-like growth factors
(IGF-I and IGF-II) and IGF-binding proteins in porcine serum during fetal and postnatal develop-
ment. Endocrinology 128, 2336–2344.
Lee, C.Y., Chung, C.S., Simmen, F.A., 1993. Ontogeny of the porcine insulin-like growth factor system.
Mol. Cell. Endocrinol. 93, 71–80.
Leibovich, H., Gertler, A., Bazer, F.W., Gootwine, E., 2000. Active immunization of ewes against ovine
placental lactogen increases birth weight of lambs and milk production with no adverse effect on
conception rate. Anim. Reprod. Sci. 64, 33–47.
Lemons, J.A., Schreiner, R.L., 1983. Amino acid metabolismin the ovine fetus. Amer. J. Physiol. 244,
E459–E466.
Leury, B.J., Bird, A.R., Chandler, K.D., Bell, A.W., 1990a. Effects of maternal undernutrition and
exercise on glucose kinetics in fetal sheep. Brit. J. Nutr. 64, 463–472.
Leury, B.J., Bird, A.R., Chandler, K.D., Bell, A.W., 1990b. Glucose partitioning in the pregnant ewe:
effects of undernutrition and exercise. Brit. J. Nutr. 64, 449–462.
Li, J., Forhead, A.J., Dauncey, M.J., Gilmour, R.S., Fowden, A.L., 2002. Control of growth hormone
receptor and insulin-like growth factor-I expression by cortisol in ovine fetal skeletal muscle.
J. Physiol. 541, 581–589.
Li, J., Owens, J.A., Owens, P.C., Saunders, J.C., Fowden, A.L., Gilmour, R.S., 1996. The ontogeny of
hepatic growth hormone receptor and insulin-like growth factor I gene expression in the sheep fetus
during late gestation: developmental regulation by cortisol. Endocrinology 137, 1650–1657.
Li, J., Saunders, J.C., Gilmour, R.S., Silver, M., Fowden, A.L., 1993. Insulin-like growth factor-II mes-
senger ribonucleic acid expression in fetal tissues of the sheep during late gestation: effects of
cortisol. Endocrinology 132, 2083–2089.
Liechty, E.A., Boyle, D.W., Moorehead, H., Lee, W.-H., Bowsher, R.R., Denne, S.C., 1996. Effects of cir-
culating IGF-I on glucose and amino acid kinetics in the ovine fetus. Amer. J. Physiol. 271, E177–E185.
Liechty, E.A., Boyle, D.W., Moorehead, H., Liu, Y.M., Denne, S.C., 1993. Increased fetal glucose
concentration decreases ovine fetal leucine oxidation independent of insulin. Amer. J. Physiol. 265,
E617–E623.
Liechty, E.A., Denne, S., Lemons, J.A., Klein, C.L., 1991a. Effects of glucose infusion on leucine
transamination and oxidation in the ovine fetus. Pediat. Res. 30, 423–429.
Liechty, E.A., Kelley, J., Lemons, J.A., 1991b. Effect of fasting on uteroplacental amino acid metabolism
in the pregnant sheep. Biol. Neonate 60, 207–214.
Liu, J.-L., Baker, J., Perkins, A.S., Roberson, E.J., Efstratiadis, A., 1993. Mice carrying null mutations of
the genes encoding insulin-like growth factor-I and type 1 receptor. Cell 75, 59–72.
Lopez, M.F., Dikkes, P., Zurakowski, D., Villa-Komaroff, L., Majzoub, J.A., 1999. Regulation of hepatic
glycogen in the insulin-like growth factor II-deficient mouse. Endocrinology 140, 1442–1448.
Lorijn, R.H.W., Nelson, J.C., Longo, L.D., 1980. Induced fetal hyperthyroidism: cardiac output and
oxygen consumption. Amer. J. Physiol. 235, H302–H307.
Loy, G.L., Quick, A.N. Jr., Hay, W.W. Jr., Meschia, G., Battaglia, F.W., Fennessey, P.V., 1990.
Fetoplacental deamination and decarboxylation of leucine. Amer. J. Physiol. 259, E492–E497.
McMillen, I.C., Adams, M.B., Ross, J.T., Coulter, C.L., Simonetta, G., Owens, J.A., Robinson, J.S., Edwards,
L.J., 2001. Fetal growth restriction: adaptations and consequences. Reproduction 122, 195–204.
McNeill, D.M., Slepetis, R., Ehrhardt, R.A., Smith, D.M., Bell, A.W., 1997. Protein requirements of
sheep in late pregnancy: partitioning of nitrogen between gravid uterus and maternal tissues. J. Anim.
Sci. 75, 809–816.
Mellor, D.J., Matheson, I.C., Small, J., 1977. Some changes in the composition of maternal and fetal
plasma from chronically catheterised sheep during short periods of reduced feed intake in late
pregnancy. Res. Vet. Sci. 23, 119–121.
Regulation of metabolism and growth during prenatal life 31

Meschia, G., Battaglia, F.C., Hay, W.W. Jr., Sparks, J.W., 1980. Utilization of substrates by the ovine
placenta in vivo. Fed. Proc. 39, 245–249.
Mesiano, S., Young, I.R., Hey, A.W., Browne, C.A., Thorburn, G.D., 1989. Hypophysectomy of the fetal
lamb leads to a fall in the plasma concentration of insulin-like growth factor I (IGF-I), but not
IGF-II. Endocrinology 124, 1485–1491.
Meznarich, H.K., Hay, W.W. Jr., Sparks, J.W., Meschia, G., Battaglia, F.C., 1987. Fructose disposal and
oxidation rates in the ovine fetus. Quart. J. Exp. Physiol. 72, 617–625.
Milley, J.R., 1994. Effects of insulin on ovine fetal leucine kinetics and protein metabolism. J. Clin.
Invest. 93, 1616–1624.
Milley, J.R., 1996. Fetal substrate uptake during increased ovine fetal cortisol concentration. Amer.
J. Physiol. 271, E186–E191.
Molina, R.D., Meschia, G., Battaglia, F.C., Hay, W.W. Jr., 1991. Gestational maturation of placental
glucose transfer capacity in sheep. Amer. J. Physiol. 261, R697–R704.
Moores, R.R. Jr., Vaughn, P.R., Battaglia, F.C., Fennessey, P.V., Wilkening, R.B., Meschia, G., 1994.
Glutamate metabolism in fetus and placenta of late-gestation sheep. Amer. J. Physiol. 267, R89–R96.
Mounzih, K, Qiu, J., Ewart-Toland, A., Chehab, F.F., 1998. Leptin is not necessary for gestation and par-
turition but regulates maternal nutrition via a leptin resistance state. Endocrinology 139, 5259–5262.
Mühlhäusler, B.S., Roberts, C.T., McFarlane, J.R., Kauter, K.G., McMillen, I.C., 2002. Fetal leptin is a
signal of fat mass independent of maternal nutrition in ewes fed at or above maintenance energy
requirements. Biol. Reprod. 67, 493–499.
Noble, R.C., Shand, J.H., Christie, W.W., 1985. Synthesis of C20 and C22 polyunsaturated fatty acids by
the placenta of the sheep. Biol. Neonate 47, 333–338.
Noden, D.M., deLahunta, A., 1985. The Embryology of Domestic Animals: Developmental Mechanisms,
Malformations. Williams & Wilkins, Baltimore.
Osborn, B.H., Fowlkes, J., Han, V.K.M., Freemark, M., 1992. Nutritional regulation of insulin-like
growth factor-binding protein gene expression in the ovine fetus and pregnant ewe. Endocrinology
131, 1743–1750.
Osgerby, J.C., Wathes, D.C., Howard, D., Gadd, T.S., 2002. The effect of maternal undernutrition on
ovine fetal growth. J. Endocrinol. 173, 131–141.
Owens, J. A., Falconer, J., Robinson, J. S., 1986. Effect of restriction of placental growth on umbilical
and uterine blood flows. Amer. J. Physiol. 250, R427–R434.
Owens, J.A., Falconer, J., Robinson, J.S., 1987a. Effect of restriction of placental growth on fetal and
utero-placental metabolism. J. Dev. Physiol. 9, 225–238.
Owens, J.A., Falconer, J., Robinson, J.S., 1987b. Restriction of placental size in sheep enhances
efficiency of placental transfer of antipyrine, 3-O-methyl-D-glucose but not urea. J. Dev. Physiol. 9,
457–464.
Owens, J.A., Kind, K.L., Carbone, F., Robinson, J.S., Owens, P.C., 1994. Circulating insulin-like growth
factors-I and -II and substrates in fetal sheep following restriction of placental growth. J. Endocrinol.
140, 5–13.
Paolini, C.L., Meschia, G., Fennessey, P.V., Pike, A.W., Teng, C., Battaglia, F.C., Wilkening, R.B., 2001.
An in vivo study of ovine placental transport of essential amino acids. Amer. J. Physiol. 280, E31–E39.
Parkes, M.J., Bassett, J.M., 1985. Antagonism by growth hormone of insulin action in fetal sheep.
J. Endocrinol. 105, 379–382.
Philipps, A.F., Dubin, J.W., Matty, P.J., Raye, J.R., 1983. Influence of exogenous glucagon on fetal glucose
metabolism and ketone production. Pediat. Res. 17, 51–56.
Phillips, I.D., Simonetta, G., Owens, J.A., Robinson, J.S., Clarke, I.J., McMillen, I.C., 1996. Placental
restriction alters the functional development of the pituitary-adrenal axis in the sheep fetus during late
gestation. Pediat. Res. 40, 861–866.
32 A. W. Bell et al.

Polk, D.H., Wu, S.Y., Wright, C., Reviczky, A.L., Fisher, D.A., 1988. Ontogeny of thyroid hormone effect
on tissue 5′-monoiodinase activity in fetal sheep. Amer. J. Physiol. 254, E337–E341.
Rankin, J.H.G., Jodarski, G., Shanahan, M.R., 1986. Maternal insulin and placental 3-O-methyl glucose
transport. J. Dev. Physiol. 8, 247–253.
Rehfeldt, C., Kuhn, G., Nürnberg, G., Kanitz, E., Schneiders, F., Beyer, M., Nürnberg, K., Ender, K.,
2001. Effects of exogenous somatotropin during early gestation on maternal performance, fetal
growth, and composition traits in sheep. J. Anim. Sci. 79, 1789–1799.
Reynolds, L.P., Ferrell, C.L., Robertson, D.A., Ford, S.P., 1986. Metabolism of the gravid uterus, foetus
and utero-placenta at several stages of gestation in cows. J. Agr. Sci. 106, 437–444.
Rhind, S.M., Robinson, J.J., McDonald, I., 1980. Relationships among uterine and placental factors in
prolific ewes and their relevance to variations in foetal weight. Anim. Prod. 30, 115–124.
Rhoads, R.P., Greenwood, P.L., Bell, A.W., Boisclair, Y.R., 2000a. Organization and regulation of the
gene encoding the sheep acid-labile subunit of the 150-kilodalton insulin-like growth factor-binding
protein complex. Endocrinology 141, 1425–1433.
Rhoads, R.P., Greenwood, P.L., Bell, A.W., Boisclair, Y.R., 2000b. Nutritional regulation of the genes
encoding the acid-labile subunit and other components of the circulating insulin-like growth factor
system in the sheep. J. Anim. Sci. 78, 2681–2689.
Robinson, J.S., Chidzanja, S., Kind, K., Lok, F., Owens, P., Owens, J., 1995. Placental control of fetal
growth. Reprod. Fertil. Dev. 7, 333–344.
Robinson, J.S., Hart, I.C., Kingston, E.J., Jones, C.T., Thorburn, G.D., 1980. Studies on the growth of the
fetal sheep: the effects of reduction of placental size on hormone concentration in fetal plasma. J. Dev.
Physiol. 2, 239–248.
Ross, J.C., Fennessey, P.V., Wilkening, R.B., Battaglia, F.C., Meschia, G., 1996. Placental transport and
fetal utilization of leucine in a model of fetal growth retardation. Amer. J. Physiol. 270, E491–E503.
Rudolph, C.D., Roman, C., Rudolph, A.M., 1989. Effect of cortisol on hepatic gluconeogenesis in the
fetal sheep. J. Dev. Physiol. 11, 219–223.
Schoknecht, P.A., Currie, W.B., Bell, A.W., 1992. Kinetics of placental lactogen in mid- and late-gestation
ovine fetuses. J. Endocrinol. 188, 95–100.
Schoknecht, P.A., McGuire, M.A., Cohick, W.S., Currie, W.B., Bell, A.W., 1996. Effect of chronic
infusion of placental lactogen on ovine fetal growth in late gestation. Domest. Anim. Endocrinol.
13, 519–528.
Schreiner, R.L., Nolen, P.A., Bonderman, P.W., Moorehead, H.C., Gresham, E.L., Lemons, J.A.,
Escobedo, M.B., 1980. Fetal and maternal hormonal response to starvation in the ewe. Pediat. Res. 14,
103–108.
Shen, W.-H., Yang, X., Boyle, D.W., Lee, W.H., Liechty, E.A., 2001. Effects of intravenous insulin-like
growth factor-I and insulin administration on insulin-like growth factor binding proteins in the ovine
fetus. J. Endocrinol. 171, 143–151.
Silver, M., Fowden, A.L., Taylor, P.M., Knox, J., Hill, C.M., 1994. Blood amino acids in the pregnant
mare and fetus: the effects of maternal fasting and intrafetal insulin. Exp. Physiol. 79, 423–433.
Simmons, M.A., Battaglia, F.C., Meschia, G., 1979. Placental transfer of glucose. J. Dev. Physiol. 1, 227–243.
Slotkin, T.A., Seidler, F.J., 1988. Adrenomedullary catecholamine release in the fetus and newborn:
secretory mechanisms and their role in stress and survival. J. Dev. Physiol. 10, 1–16.
Smeaton, T.C., Owens, J.A., Kind, K.L., Robinson, J.S., 1989. The placenta releases branched-chain amino
acids into the umbilical and uterine circulations in the pregnant sheep. J. Dev. Physiol. 12, 95–99.
Sparks, J.W., Hay, W.W. Jr., Bonds, D., Meschia, G., Battaglia, F.C., 1982. Simultaneous measurements
of lactate turnover rate and umbilical lactate uptake in the fetal lamb. J. Clin. Invest. 70, 179–192.
Stegeman, J.H.G., 1974. Placental development in the sheep and its relation to fetal development. Bjd. Dierk.
44, 1–72.
Regulation of metabolism and growth during prenatal life 33

Stephens, E., Thureen, P.J., Goalstone, M.L., Serson, M., Leitner, J.W., Hay, W.W. Jr., Draznin, B., 2001.
Fetal hyperinsulinemia increases farnesylation of p21 Ras in fetal tissues. Amer. J. Physiol. 281,
E217–E223.
Sterle, J.A., Cantley, T.C., Lamberson, W.R., Lucey, M.C., Gerrard, D.E., Matteri, R.L., Day, B.N., 1995.
Effects of recombinant porcine somatotropin on placental size, fetal growth, and IGF-I and IGF-II
concentrations in pigs. J. Anim. Sci. 73, 2980–2985.
Stevens, D., Alexander, G., 1986. Lipid deposition after hypophysectomy and growth hormone treatment
in the sheep fetus. J. Dev. Physiol. 8, 139–145.
Stevens, D., Alexander, G., Bell, A.W., 1990. Effect of prolonged glucose infusion into fetal sheep on
body growth, fat deposition and gestation length. J. Dev. Physiol. 13, 277–281.
Teasdale, F., 1976. Numerical density of nuclei in the sheep placenta. Anat. Rec. 185, 187–196.
Thomas, L., Wallace, J.M., Aitken, R.P., Mercer, J.G., Trayhurn, P., 2001. Circulating leptin during ovine
pregnancy in relation to maternal nutrition, body composition and pregnancy outcome. J. Endocrinol.
169, 465–476.
Thureen, P.J., Merson, S., Hay, W.W. Jr., 2000. Regulation of uterine and umbilical amino acid uptakes
by maternal amino acid concentrations. Amer. J. Physiol. 279, R849–R859.
Thureen, P.J., Padbury, J.F., Hay, W.W. Jr., 2001. The effect of maternal hypoaminoacidaemia on
placental uptake and transport of amino acids in pregnant sheep. Placenta 22, 162–170.
Thureen, P., Trembler, K.A., Meschia, G., Makowski, E.L., Wilkening, R.B., 1992. Placental glucose
transport in heat-induced fetal growth retardation. Amer. J. Physiol. 263, R578–R585.
Timmerman, M., Chung, M., Wilkening, R.B., Fennessey, P.V., Battaglia, F.C., Meschia, G., 1998.
Relationship of fetal alanine uptake and placental alanine metabolism to maternal plasma alanine
concentration. Amer. J. Physiol. 275, E942–E950.
Timmerman, M., Teng, C., Wilkening, R.B., Fennessey, P., Battaglia, F.C., Meschia, G., 2000. Effect of
dexamethasone on fetal hepatic glutamine-glutamate exchange. Amer. J. Physiol. 278, E839–E845.
Townsend, S.F., Rudolph, C.D., Rudolph, A.M., 1991. Cortisol induces perinatal hepatic gluconeogene-
sis in the lamb. J. Dev. Physiol. 16, 71–79.
Van Veen, L.C.P., Teng, C., Hay, W.W., Jr., Meschia, G., Battaglia, F.C., 1987. Leucine disposal and
oxidation rates in the fetal lamb. Metabolism 36, 48–53.
Van Vliet, G., Styne, D.M., Kaplan, S.L., Grumbach, M.M., 1983. Hormone ontogeny in the ovine fetus.
XVI. Plasma immunoreactive somatomedin C/insulin-like growth factor I in the fetal and neonatal
lamb and in the pregnant ewe. Endocrinology 113, 1716–1720.
Vatnick, I., Bell, A.W., 1992. Ontogeny of fetal hepatic and placental growth and metabolism in sheep.
Amer. J. Physiol. 263, R619–R623.
Vaughn, P.R., Lobo, C., Battaglia, F.C., Fennessey, P.V., Wilkening, R.B., Meschia, G., 1995. Glutamine-
glutamate exchange between placenta and fetal liver. Amer. J. Physiol. 268, E705–E711.
Vernon, R.G., Robertson, J.P., Clegg, R.A., Flint, D.J., 1981. Aspects of adipose-tissue metabolism in
foetal lambs. Biochem. J. 196, 819–824.
Wallace, J.M., Bourke D.A., Aitken, R.P., Leitch, N., Hay, W.W. Jr., 2002. Blood flows and nutrient
uptakes in growth-restricted pregnancies induced by overnourishing adolescent sheep. Amer.
J. Physiol. 282, R1027–R1036.
Wallace, J.M., Bourke, D.A., Aitken, R.P., Palmer, R.M., Da Silva, P., Cruickshank, M.A., 2000.
Relationship between nutritionally-mediated placental growth restriction and fetal growth, body com-
position and endocrine status during late gestation in adolescent sheep. Placenta 21, 100–108.
Ward, J.W., Wooding, F.B.P., Fowden, A.L., 2002. The effects of cortisol on the binucleate cell popula-
tion in the ovine placenta during late gestation. Placenta 23, 451–458.
Whorwood, C.B., Firth, K.M., Budge, H., Symonds, M.E., 2001. Maternal undernutrition during early to
midgestation programs tissue-specific alterations in the expression of the glucocorticoid receptor,
34 A. W. Bell et al.

11β-hydroxysteroid dehydrogenase isoforms, and type 1 angiotensin II receptor in neonatal sheep.


Endocrinology 142, 2854–2864.
Widdas, W.F., 1952. Inability of diffusion to account for placental glucose transfer in the sheep and con-
sideration of the kinetics of a possible carrier transfer. J. Physiol. 118, 23–39.
Wilkening, R.B., Molina, R.D., Battaglia, F.C., Meschia, G., 1987. Effect of insulin on glucose/oxygen
and lactate/oxygen quotients across the hindlimb of fetal lambs. Biol. Neonate 51, 18–23.
Wooding, F.B.P., Morgan, G., Fowden, A.L., Allen, W.R., 2000. Separate sites and mechanisms for
placental transport of calcium, iron and glucose in the equine placenta. Placenta 21, 635–645.
Young, M., McFadyen, I.R., 1973. Placental transfer and fetal uptake of amino acids in the pregnant ewe.
J. Perinatal Med. 1, 174–182.
Yuen, B.S.J., McMillen, I.C., Symonds, M.E., Owens, P.C., 1999. Abundance of leptin mRNA in fetal
adipose tissue is related to fetal body weight. J. Endocrinol. 163, R11–R14.
Yuen, B.S.J., Owens, P.C., McFarlane, J.R., Symonds, M.E., Edwards, L.J., Kauter, K.G., McMillen, I.C.,
2002. Circulating leptin concentrations are positively related to leptin messenger RNA expression in
adipose tissue of fetal sheep in the pregnant ewe fed at or below maintenance energy requirements
during late gestation. Biol. Reprod. 67, 911–916.
Yuen, B.S.J., Owens, P.C., Mühlhäusler, B.S., Roberts, C.T., Symonds, M.E., Keisler, D.H., McFarlane,
J.R., Kauter, K.G., Evens, Y., McMillen, I.C., 2003. Leptin alters the structural and functional char-
acteristics of adipose tissue before birth. FASEB J. Express article 10.1096/fj.02–0756fje. Published
online, April 22, 2003.
Zhou, J., Bondy, C.A., 1993. Placental glucose transporter gene expression and metabolism in the rat.
J. Clin. Invest. 91, 845–852.
PART II
Protein metabolism
2 Regulation of skeletal muscle protein
metabolism in growing animals

T. A. Davis and M. L. Fiorotto

United States Department of Agriculture/Agricultural Research Service,


Children’s Nutrition Research Center, Department of Pediatrics,
Baylor College of Medicine, Houston, TX 77030, USA

The skeletal musculature is not only of great significance to the physiological function and
long-term well-being of the growing animal, but, by virtue of its large mass, has tremendous
impact on the overall rate of protein metabolism in the whole body. Protein deposition is very
rapid during early life and this is largely driven by the high fractional rate of protein synthesis
in skeletal muscle. A number of factors regulate the growth, development, and metabolic
activity of the skeletal musculature, and these include the intrinsic or genetic factors that
influence muscle differentiation as well as the extrinsic factors such as nutrients, hormones,
and activity that influence muscle hypertrophy.

1. INTRODUCTION
In the mature adult, the skeletal musculature constitutes the largest single protein pool in the
body, and comprises approximately 60% of the body’s metabolically active mass. Thus,
despite its relatively low basal rate of metabolism, skeletal muscle mass is such that changes
to its composition and/or its size have implications for the overall metabolism of the body.
Until relatively recently, the primary interests in skeletal muscle metabolism were related to its
functional role in dictating locomotor performance, specifically speed, strength, and endurance,
and its influence on the quantity and quality of meat products that constitute a primary source
of protein and micronutrients in the human diet. More recently, however, a renewed interest
in the contribution of the muscle metabolism to the overall health of the human individual has
emerged. This interest, together with the developments in our understanding of the regulation
of gene expression and cellular signalling, have spurred substantial amounts of research to
advance our understanding of how muscle mass, metabolism, and function respond to nutrients,
hormones, growth factors, activity, and other anabolic agents.
The fully differentiated skeletal muscle is made up of multinucleated myofibres. The post-
natal growth rate of muscle mass is a function of the total number of fibres, and the growth

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
37 © 2005 Elsevier Limited. All rights reserved.
38 T. A. Davis and M. L. Fiorotto

rate of each fibre. Thus, understanding the regulation of myofibre formation during prenatal life,
and the rate of muscle protein accretion in postnatal life, are critical for evaluating the animal’s
maximal capacity for muscle growth. The mature myofibre is composed primarily of the three
basic systems required for muscle contraction: the myofibrillar proteins composed primarily of
the contractile elements and the associated structural proteins; an extensive membrane system
that regulates the release and uptake of ions in response to the neural inputs; and the mitochon-
drial and cytoplasmic system of enzymes involved in the generation of the ATP required to drive
functional processes. These components vary in their relative abundance, as well as in their level
of activity, thereby giving rise to substantial diversity in fibre function and size.
The combination of functional and metabolic properties of fibres is used as the basis for
the standard classification of muscle fibres in the adult. Muscle fibres can be divided into
two main categories on the basis of their twitch characteristics, that is, slow-twitch (S) or
fast-twitch (F), which correspond with the Type I and Type II nomenclature, respectively. The
contractile property of the myofibrils is largely determined by the ATPase activity of the
myosin heavy chain (MHC) isoform expressed within each fibre (Schiaffino and Reggiani,
1996). Fibres are also identified in a variety of ways by their metabolic properties: slow-twitch
fibres, and a subcategory of fast-twitch fibres, generate ATP predominantly by oxidative
metabolism (O). All fast-twitch fibres also can generate limited amounts of ATP by glycolysis
and, hence, they are categorized as glycolytic (FG or Type IIB) or fast-twitch, oxidative-
glycolytic (FOG or Type IIA). The capacity for oxidative metabolism is supported by a rela-
tively high abundance of mitochondria, enzymes for fatty acid oxidation, and oxygen delivery
and uptake. The latter is effected by a high capillary density, and the presence of large amounts
of myoglobin in the sarcoplasm, both of which impart a red tint to the muscle, so that SO and
FOG fibres also are referred to as red muscle. Because these metabolic properties render to
the fibres a greater degree of fatigue-resistance, they constitute primarily those muscles that
undergo prolonged periods of sustained slow isometric contraction, such as postural muscles
(e.g. soleus and rhomboides), or muscles that are required for continuous episodes of isotonic
contraction (e.g. diaphragm and jaw muscles). In contrast, FG fibres have a paucity of these
components and thus, muscles where these fibres predominate (e.g. longissimus dorsi) are
much lighter in colour and are referred to as white muscles.
The origin of myofibre heterogeneity is complex and not entirely understood. There is
evidence that the various fibre types are derived from distinct myoblast lineages. However,
within an organism, the relative abundance of different fibre types varies between muscles
according to their physiological function. There is also variability between individuals of the
same breed, between different breeds of the same species, and between different species
(Rehfeldt et al., 1999). Fibre-type proportions are of relevance with respect to livestock, as they
are a key determinant of meat quality (Koohmaraie, 2003). Fibre diversity has its origins in the
embryo, is amplified during the process of differentiation during fetal and early postnatal life,
and thereafter is fine-tuned in response to the functional demands placed on the muscle. By and
large, however, fibre-type composition is a genetically determined, inherited trait. Muscle fibre
hypertrophy, on the other hand, occurs after the fibres have differentiated. Hypertrophy is
largely a postnatal event and metabolically is dominated by the accretion of muscle-specific
proteins. Unlike the early processes of determination, commitment, and differentiation that are
normally orchestrated by signalling molecules and intracellular pathways inherent to the devel-
oping organism and are primarily under genetic control, muscle hypertrophy is highly
responsive to external cues, such as nutrient availability, muscle use, and various hormones.
The aim of this review is to consider the factors that influence muscle protein metabolism
in the growing organism. We shall address the “heritable/congenital” component, i.e. the
Regulation of skeletal muscle protein metabolism 39

developmental aspects of muscle growth that principally determine fibre number and fibre-
type diversity, and the role of external influences that modulate muscle hypertrophy. Owing
to space limitations, this review cannot cover all aspects of muscle metabolism in equal depth.
Rather, we have selected those areas in which substantial progress has been made recently,
and have focused largely on early life, when the most marked changes occur.

2. MUSCLE DIFFERENTIATION
2.1. From myoblast to myofibre

Much of our understanding of early myogenesis is derived from studies on the chick.
However, the overall pattern is similar in mammals as indicated by the extensive study of the
mouse in which the advent of transgenic technology has permitted the function of individual
genes to be identified. Myogenesis begins in early embryonic life and, with the exception of
craniofacial muscles, the skeletal musculature develops from mesodermal progenitor cells in
the somites (Perry and Rudnicki, 2000; Buckingham, 2001). The somites develop in pairs
from aggregates of epithelial cells on either side of the neural tube and notochord and mature
in a rostro-caudal direction under the regulation of positive signals and negative regulators, in
the form of diffusible molecules, produced by the tissues adjacent to the somites (Buckingham,
2001; Buckingham et al., 2003; Francis-West et al., 2003) (fig. 1). Cells of the dorsal surface of
the somite are compartmentalized into the dermomyotome and are specified to form myogenic
and dermal progenitors, whereas signals to cells on the ventral aspect of the somite specify
the formation of the sclerotome which gives rise to the ribs and axial skeleton (Buckingham,
2001). Myogenic precursor cells originate from the dorso-medial (epaxial) and ventro-lateral
(hypaxial) edges of the dermomyotome. The epaxial precursors delaminate and translocate
ventrolaterally to form the myotome, under the dermomyotome, and eventually they expand
and differentiate to form the deep back muscles. The hypaxial myogenic precursors of the
dermomyotome migrate ventrally to form the ventral body wall muscles, tongue, and diaphragm,
or delaminate and migrate into the limb buds to give rise to the limb musculature.
Specification of myogenic cells occurs upon the activation of myogenic regulatory factors
(MRF), specifically those that encode the basic helix–loop–helix transcription factors Myf5,
MyoD, MRF4, and myogenin (Molkentin and Olson, 1996; Perry and Rudwicki, 2000).

Fig. 1. Schematic representation of somite structure and key molecules responsible for myogenic specifica-
tion, determination, delamination, and migration.
40 T. A. Davis and M. L. Fiorotto

Myf5 and MyoD are the earliest MRFs to be expressed in the dermomyotome, and their pat-
terns of expression appear to be spatially and temporally regulated. Wnt-1 produced in the
dorsal neural tubes induces Myf5 expression in the epaxial region, whilst Wnts produced by
the dorsal ectoderm induce MyoD expression by the dermomyotome. Sonic hedgehog (Shh),
produced from the notochord and floor plate of the neural tube, also appears to play a role in
the early activation of Myf5 in the determination of the epaxial dermomyotome (Buckingham
et al., 2003). Cells committed to the muscle lineage express the homeobox gene, Pax-3, and its
expression is most likely activated by Wnts and Shh. During early myogenesis, however, the
differentiation of cells that express Pax-3 and are committed to the muscle lineage is blocked
by the expression of inhibitory factors. These include the growth factor, bone morphogenic
protein-4 (BMP-4), and fibroblast growth factors, both of which are produced by the lateral
plate mesoderm. Inhibition of differentiation is critical as it enables the cells to continue
proliferation, delaminate, and migrate. In the regions of the limbs, migrating cells do not
express MRFs until they reach the limbs where, after some delay during which they undergo
several rounds of replication, they give rise to the appendicular muscles (Buckingham, 2001;
Buckingham et al., 2003; Francis-West et al., 2003).
Delamination and migration of muscle precursor cells are crucial steps in myogenesis
(Birchmeier and Brohmann, 2000; Francis-West et al., 2003) and require the activation of the
c-met receptor by hepatocyte growth factor (HGF, also known as scatter factor) (Scaal et al.,
1999; Birchmeier and Brohmann, 2000). Transcription of the c-met gene is activated by Pax-3,
and in its absence no limb muscles form, whereas when a constitutively active form is
expressed, there is an overproduction of hypaxial muscles (Epstein et al., 1996). Lbx1 is another
transcription factor required for migration of the somitic cells, with particular importance for
those cells that give rise to the dorsal muscle masses of the limbs (Brohmann et al., 2000).
Once commitment of cells to the myogenic lineage has occurred, the cells are prevented
from differentiating further by a variety of gene products produced by the cells themselves, as
well as the cells’ matrix, and a variety of mitogens that promote proliferation (Perry and
Rudnicki, 2000; Fuchtbauer, 2002). The maintenance of the committed, but not fully differ-
entiated, state is a critical determinant of myofibre number and size because it permits
myoblast proliferation to continue and, hence, to expand the population of cells that can give
rise to myofibres (Fuchtbauer, 2002). Indeed, all instances of muscle hypertrophy that are
associated with increased myofibre number have their origin in embryonic and fetal life, and
by birth, fibre number is fixed (Rehfeldt et al., 1999). The degree of myoblast proliferation
in vivo is the product of the balance between activities of stimulatory and inhibitory factors
(Fuchtbauer, 2002). The former include Msx1, basic fibroblast growth factor (bFGF), HGF,
and the insulin-like growth factors (IGF)-I and -II. The transforming growth factor (TGF-β)
super-family of peptides exert a variety of effects which appear to be species-dependent, but
by and large, they inhibit terminal differentiation with either little effect or even inhibition
of proliferation (Fuchtbauer, 2002). Members of the TGF-β family that have been shown to
regulate myogenesis include TGF-β itself, activin, BMPs, and growth differentiation factor 8,
also known as myostatin (McPherron et al., 1997; Fuchtbauer, 2002). Their precise role and
the incurred responses vary between muscle beds and according to the net balance between
positive and negative regulators.
The in vivo significance of these factors in regulating myoblast hyperplasia and fibre
formation has been demonstrated in studies with an extensive variety of transgenic animals.
Notable among these has been the development of transgenic mice in which the myostatin
gene is inactivated (McPherron et al., 1997). Myostatin is expressed beginning early in
embryonic life and inhibits cell cycle progression, thereby limiting myoblast proliferation
Regulation of skeletal muscle protein metabolism 41

(Lee and McPherron, 1999). Myostatin also inhibits differentiation by down-regulating


MyoD and myogenin expression and activity (Langley et al., 2002). Hence, in its absence,
there is greater myoblast proliferation and this ultimately results in the formation of a larger
number of muscle fibres. The characterization of myostatin and its mechanism of action
pointed to the identification of the genetic basis for the muscle hypertrophy in some breeds
of cattle, commonly known as “double-muscling” (Grobet et al., 1997). In vivo, myostatin is
regulated by follistatin, and follistatin overproduction increases muscle mass, whereas
impaired production reduces muscle mass.
The IGFs are unique in their actions in that they can stimulate both proliferation and dif-
ferentiation by activating the type 1 IGF receptor; these effects, however, are temporally
separated. The switch from proliferative to myogenic effects is associated with changes in the
intracellular signalling pathways from primary activation of the mitogen-activated protein
kinase (MAPK) pathway during the proliferative phase, to signalling through the phos-
phatidylinositol 3-kinase (PI 3-kinase) pathway for differentiation (Coolican et al., 1997). In
cell culture, the proliferative phase is associated with IGF-I-stimulated expression of cell
cycle markers and cell proliferation and reduced expression of myogenic markers, whereas
during differentiation IGF-I promotes expression of myogenin and muscle-specific gene
expression (Engert et al., 1996). The exact trigger responsible for switching the response from
proliferation to differentiation is unclear, especially in vivo. Studies in the chick embryo have
demonstrated that increased local expression of IGF-I in the limb at an early stage of devel-
opment before cells have differentiated increases myoblast number with the formation of
larger muscles containing an increased number of myofibres (Mitchell et al., 2002), in much
the same way as, although independently of, decreased myostatin expression. Similarly,
administration of growth hormone (GH) to pregnant sows in early gestation (10–24 days)
indirectly stimulates fetal IGF secretion and results in greater myofibre number at birth
(Rehfeldt et al., 1993). This response contrasts with the response to over-expression of IGF-I
in the differentiated muscle, where muscle hypertrophy is not associated with increased fibre
number (Musaro et al., 2001; Fiorotto et al., 2003).
The stimulation of muscle differentiation by the MRFs entails not only up-regulation of
their expression within presumptive muscle cells, but also the co-ordinated orchestration of
a series of events that enables them to be transcriptionally active. These include: the down-
regulation of Id proteins which mitigate the binding of MRFs to their E-box consensus
sequence (Benezra et al., 1990); the association with the myogenic enhancer factor 2 (MEF2)
family of transcription factors that bind to both their own DNA site and form protein–protein
interactions with the MRFs (Molkentin and Olson, 1996); and the dissociation of histone
deacetylases from the transcription factors and subsequent recruitment of histone acetylases
to E-boxes associated with muscle-specific genes. The resulting acetylation of histones
produces the chromatin relaxation necessary to permit transcription of the underlying gene
(McKinsey et al., 2001, 2002). This differentiation step is associated concomitantly with
inhibition of cell cycling, and the alignment and subsequent fusion of the adhered myoblasts
to form myotubes.
Fibres form in two waves: the first wave occurs during early embryogenesis and results
in the formation of primary myotubes that shape and position the orientation of individual
muscles (Ontell and Dunn, 1978; Ontell, 1982). These primary fibres are originally in clusters,
but progressively become separated by the basal lamina. Subsequently, a secondary population
of myoblasts located under the basal lamina of the primary fibres begins to proliferate using
the plasma membrane of the primary fibres as scaffolding, but without fusing to them. These
then fuse among themselves to form secondary myotubes, and gradually separate from the
42 T. A. Davis and M. L. Fiorotto

primaries by forming their own basal lamina. There is a third population of myoblasts, the
satellite cells, that remain quiescent, sandwiched between the plasma membrane and basal
lamina. The expression of the homeobox protein, Pax-7, distinguishes satellite cells from
other lineages of myoblast, and appears to be essential for the formation of satellite cells
(Seale et al., 2000). Satellite cells proliferate to enlarge the myonuclear population present in
the myofibres to which they are attached, and contribute only to myofibre hypertrophy, not
new fibres (except to repair muscle damage).
To this point in the process of skeletal muscle development, the primary effect of pertur-
bations in the growth process are likely to manifest themselves as variations in the number of
myofibres that form. The timing of the perturbation will determine if primary or secondary
fibres are affected. It is unclear if impairment of satellite cell replication during the terminal
phase of muscle formation can permanently impact the size of a muscle’s reserve of satellite
cells and, hence, postnatal muscle growth potential.

2.2. Compositional development

Once myoblasts have fused, the expression of muscle-specific proteins dominates muscle
growth. The cells undergo a complex set of transformations to create the highly structured
myofibrils which have the capacity to perform contractile work. This process of maturation
is critical for the developing offspring as it is essential for postnatal survival: skeletal muscles
are a prerequisite for independent breathing and suckling. Indeed, the offspring die at birth in
all transgenic animals in which normal skeletal muscle development is impaired by targeted
disruption of key regulatory molecules (e.g. Venuti et al., 1995). However, there is a wide
variation in the level of muscle maturity at birth among species, and between muscle groups
within the same individual. From these observations, it is evident that birth occurs at different
stages of muscle development across species, and that within an individual, maturation
proceeds in a rostral to caudal direction, and from proximal to distal in the appendages.
Hence, altricial species, like rabbits, rodents, and most carnivores, are born with functional
head and thoracic muscles, whereas their lower abdominal and limb muscles, especially those
of the hindlimbs, are still immature. In contrast, precocial species, such as ungulates, have
relatively longer gestations and the newborn muscle is at a fairly advanced stage of maturity.
Indeed, locomotor function is attained very soon after birth.
The maturation of myotubes into fully functional myofibres involves the co-ordinated
development of the metabolic machinery of the cells, the ion transport membrane system, and
the contractile elements. The complex sarcoplasmic reticulum and T-tubular system respon-
sible for coupling excitation and contraction followed by muscle relaxation develop in a
co-ordinated fashion and attain their mature configuration at approximately 2 weeks of age in
the rat. At this point, the membrane system constitutes approximately 40% of the non-
myofibrillar compartment in the muscle fibre (Schiaffino and Margreth, 1969). Contractile
proteins, not present in the myotube at fusion, comprise 55−65% of total muscle protein by
2 weeks of age in the rat (Yates and Greaser, 1983; Fiorotto et al., 2000a). The accretion of
myofibrillar proteins, therefore, is a major determinant of whole-body protein accretion
during this developmental period.
The diversity in fibre-type characteristics results from the combination of the inherent
properties of the myoblast lineage from which the fibres were derived and extrinsic signals
from the organism. Thus, in the development of a myofibre from a myotube, in addition to
the first-order general pattern of compositional development that occurs (described above),
Regulation of skeletal muscle protein metabolism 43

significant changes also occur in the isoform composition of muscle proteins and the
supporting metabolic machinery. These result in a second order of compositional changes that
produce the development of the adult phenotype of individual fibre types (Schiaffino and
Reggiani, 1996). A significant proportion of the proteins that constitute the thick, intermediate,
and thin filaments initially are expressed as immature isoforms and subsequently are replaced
by the adult isoforms during maturation (table 1). In the thick filaments, all primary fibres
initially express a combination of embryonic and slow MHC isoforms, or neonatal and
slow MHC isoforms. The associated myosin light chains (MLC) also differ according to the
MHC isoform with which they are associated and, therefore, their expression changes during
maturation. Secondary fibres, in contrast, initially express the embryonic and neonatal MHC
isoforms, which are replaced usually by any of the adult fast MHC isoforms according to the
functional characteristics of the individual muscle. This process of terminal maturation is
amenable to regulation by hormones, activity pattern, and neural inputs, and contrasts with the
earlier differentiation of myoblast lineage which proceeds independently of extrinsic factors.
A fundamental conundrum regarding the differentiation among fibre type relates to the
mechanism that co-ordinates the appropriate expression of the plethora of proteins responsible
for the metabolic and contractile characteristics of muscle. Recent studies have focused on the

Table 1
Contractile protein isoforms expressed in the developing and mature muscles of the rat

Developing muscles Adult muscles

Embryonic Neonatal Fast Slow

Myosin heavy chains


embryonic neonatal 2B β-slow
β-slow (embryonic) 2X
2A
Myosin light chains (MLC)
MLC-1embryonic MLC-1fast MLC-1fast MLC-1slow/ventricular
MLC-1slow-α (MLC-3fast) MLC-3fast (MLC-1slow-α)
MLC-1fast
MLC-2fast MLC-2fast MLC-2fast MLC-2slow
Actin
α-cardiac α-skeletal α-skeletal α-skeletal
α-skeletal (α-cardiac)
Troponins (Tn)
TnC-fast TnC-fast TnC-fast TnC-slow/cardiac
TnC-slow/cardiac
TnI-slow TnI-fast TnI-fast TnI-slow
TnT-cardiac TnT-fast, fetal TnT-fast, adult TnT-slow
isoforms isoforms
TnT-slow
Tropomyosins (TM)
TM-β TM-β TM-αfast TM-αslow
TM-αfast TM-αfast (TM-β) TM-αfast
TM-αslow TM-β
Minor isoforms are indicated in parentheses. Isoform profile indicated for neonatal developing muscles is that of
the majority of hindlimb muscle fibres, which are secondary generation fibres destined to become fast-type fibres.
Adapted from Schiaffino and Reggiano (1996).
44 T. A. Davis and M. L. Fiorotto

role of calcineurin (CaN), a calcium and calmodulin-dependent serine/threonine phosphatase,


in the regulation of expression of those genes responsible for the slow muscle phenotype
(Schiaffino and Serrano, 2002; Spangenburg and Booth, 2003). CaN is activated when there
are high intracellular steady-state levels of Ca2+, typical of slow fibres that are subjected to
chronic, low-frequency nerve stimulation. Once activated, CaN dephosphorylates the tran-
scription factors, nuclear factor of activated T-cells (NFAT) and MEF-2, so that they can then
be translocated into the nucleus where, in conjunction with MRFs, they effect changes in gene
expression (McKinsey et al., 2001, 2002). These factors bind to their respective DNA consen-
sus sequences which form a characteristic motif, the slow upstream element (SURE), present
in the promoter of a variety of slow muscle genes such as slow troponin I and myoglobin
(Calvo et al., 2001). Despite the ability of CaN to sense and transduce changes in cell calcium
levels into changes in gene expression, it is by no means a global regulator of the SO fibre
phenotype as was initially proposed. For example, it has been observed that expression of
MHC-IIa, the predominant MHC isoform in FOG muscles, is also highly responsive to CaN
activation (Allen and Leinwand, 2002).
Considerable progress has been made in identification of the mechanisms that co-ordinate
the contractile and metabolic characteristics of muscle. Again, the sustained elevation of
intracellular calcium appears to be a central factor in signalling not only the fast-to-slow shift
in muscle gene expression (Allen and Leinwand, 2002), but also an increase in mitochondrial
biogenesis (Ojuka et al., 2003). In addition to CaN, calcium activates calcium/calmodulin-
dependent protein kinases (CaMK) which catalyse a series of reactions that result in the
transcription of a coactivator of nuclear receptors, peroxisome proliferator-activated receptor
coactivator-1α (PGC-1α) (Handschin et al., 2003). PGC-1α plays a pivotal role in glucose
metabolism, mitochondrial biogenesis, and adaptive thermogenesis by activating various tran-
scription factors. Specifically, in muscle PGC-1α has been shown to stimulate mitochondrial
DNA replication, mitochondrial abundance, cytochrome c and cytochrome oxidase levels,
GLUT4 expression, and uncoupling protein expression. PGC-1α also enhances its own tran-
scription (Handschin et al., 2003). Consequently, once activated, an autoregulatory loop is set
up which sustains PGC-1α expression and its downstream effects, and thereby maintains
stable expression of the oxidative phenotype.
In addition to calcium, PGC-1α expression is regulated by thyroid hormone and AMP-
activated protein kinase, an enzyme that is activated by chronic reductions in the cellular
ATP/AMP ratio, for example, with energy deprivation (Irrcher et al., 2003; Ojuka et al., 2003;
Spangenburg and Booth, 2003). These mechanisms that co-ordinate the metabolic properties of
a muscle with its contractile properties and pattern of use, however, are pertinent primarily to the
development of slow-twitch and/or oxidative properties, presumably in muscles where these
characteristics are not present. This suggests that fast-twitch, glycolytic properties are the default
phenotype of skeletal muscle and there is, indeed, some evidence to support this suggestion.
During terminal maturation, the loss of polyinnervation and acquisition of single innervation
from a nerve with a low-frequency firing pattern is necessary for the development and mainte-
nance of slow-twitch characteristics. Moreover, if the soleus is denervated at birth in the rat, slow
myosin isoenzymes are gradually replaced by fast myosins (Gambke et al., 1983). Thus, the
replacement of the immature isoforms of myosin specifically by adult slow myosin occurs only
with the appropriate neural input. In the mature muscle, cross-innervation of a mature fast-twitch
muscle with the nerve from a slow-twitch muscle gradually transforms the entire contractile
phenotype and metabolic properties to those of a slow muscle (Barany and Close, 1971).
Thyroid hormone also plays a critical role in the maturation of skeletal muscle. Moreover,
because thyroid hormone is sensitive to changes in energy balance, it may serve as a signal
Regulation of skeletal muscle protein metabolism 45

to the muscle to produce adaptive changes in muscle metabolism. The importance of thyroid
hormone has been studied extensively with respect to the regulation of MHC expression
where it is required for down-regulation of the neonatal isoform of MHC in muscles destined
to become either fast or slow (Gambke et al., 1983; Adams et al., 1999). Furthermore, in the
absence of thyroid hormone, the accumulation of slow MHC is accelerated, whereas that
of IIA MHC is down-regulated. Thus, variations in energy balance that produce changes in
thyroid hormone might be anticipated to alter the postnatal pattern of muscle maturation. Our
studies in suckling rats suggest that the changes in thyroid hormone have to be relatively
severe in order to effect changes at the level of gene expression.
However, changes in muscle use and protein turnover also occur as a consequence of alter-
ations in food intake and growth rate. These changes serve to mitigate the effect of altered
gene expression and, consequently, the maturation of muscle phenotype is preserved (Fiorotto
and Davis, 1997; Fiorotto et al., 2000a). The suckling pig responds to mild hypothyroidism
during the suckling period by increasing slow MHC expression, although the effect is some-
what mitigated by increases in nuclear thyroid hormone receptor (Harrison et al., 1996).
Relatively severe energy restriction in post-weaned pigs has similar effects, increasing the
abundance of slow MHC at both the protein and mRNA level, and increasing the oxidative
properties of the muscles, but with substantial muscle-to-muscle variability (White et al.,
2000). Although the increase in slow MHC expression is compatible with the known effects
of hypothyroidism on MHC expression, the enhanced oxidative properties are the opposite of
those anticipated on the basis of PGC-1α regulation by thyroid hormone. However, they are
compatible with a change in AMP kinase activity, which increases with a chronic deficit in
energy intake, and thereby promotes mitochondrial biogenesis and fatty acid oxidation.
Overall, these responses to a chronic deficit in energy intake represent beneficial adaptations
by the muscle to enhance its metabolic efficiency: a slow muscle requires less energy than a
fast-twitch muscle to generate the same amount of tension, and it is able to derive more of its
energy by fatty acid oxidation and oxidative phosphorylation (Henriksson, 1990).
The regulation of the fast-twitch, glycolytic phenotype of muscles is much less clearly
understood than that of slow-twitch muscle. Some genes expressed in fast fibres contain
a characteristic binding motif, the fast intronic regulatory element (FIRE), analogous to the
SURE motif in slow fibre genes (Nakayama et al., 1996). Myoblast lineage established during
fetal life appears to be a primary determinant of whether a fibre matures into a fast fibre.
As noted previously, gene mutations that promote secondary myoblast proliferation result
primarily in fast fibre hypertrophy. During terminal maturation, thyroid hormone is required
for the down-regulation of neonatal MHC and, if present at high levels, thyroid hormone
tends to drive the expression of fast MHC in muscle fibres that normally would be slow
(Nakayama et al., 1996). The role of innervation also appears to be less critical in the devel-
opment of fast-twitch fibres than for slow fibres. In both rodents and chickens, denervation
delays the elimination of immature MHC isoforms, but does not prevent the development of
the fast phenotype. Inactivity also tends to promote the fast phenotype, although this is attrib-
utable in part to the preferential atrophy of slow fibres. In post-weaned pigs (Katsumata et al.,
2000), but possibly not neonatal pigs (Louveau and Le Dividich, 2002), mild undernutrition
up-regulates the expression of the growth hormone receptor (GHR) on FOG and FG fibres
which normally express the lowest level of GHR. Together with the reduction in thyroid
hormone expression, the resulting changes in hormone responsiveness may be responsible for
the metabolic shift that occurs in muscle during undernutrition and which enables it to derive
more of its energy from lipid oxidation. A broader implication of these findings, however,
relates to the anti-insulin effects of GH in the undernourished animal which serve to divert
46 T. A. Davis and M. L. Fiorotto

nutrients from muscle towards the visceral organs. This is demonstrated by the differential
response in rates of tissue growth in protein-malnourished piglets where body protein parti-
tioned into gastrointestinal tissue is preserved, while that of skeletal muscle is reduced (Ebner
et al., 1994).

2.3. Role of protein synthesis and degradation in the regulation


of compositional development

As the above discussion of the developmental changes in muscle composition indicates, pro-
tein synthetic rates in the immature muscle must sustain not only the de novo accretion of
myofibrillar proteins and membrane structures, but also their continuous and co-ordinated
replacement as the tissue develops its mature compositional and functional characteristics.
The faster accumulation of myofibrillar proteins compared to sarcoplasmic proteins is
explained almost entirely by their higher fractional synthesis rate compared to other protein
components (Fiorotto et al., 2000a). As compositional maturity is attained, the synthesis rate
of myofibrillar proteins decreases to a greater extent than sarcoplasmic proteins, and in the
mature muscle, sarcoplasmic protein synthesis rates are higher than for myofibrillar proteins.
However, once the mature composition is attained, the rates of degradation also differ in
parallel and results in the maintenance of constant composition. In altricial mammals such as
rodents and rabbits, the differential regulation of protein synthesis in the different protein
pools occurs in the immediate postnatal period. In precocial animals, the full complement of
myofibrillar proteins in fibres is mostly completed by birth, although they still undergo some
limited, second-order compositional maturation postnatally. Nonetheless, the intrauterine
pattern of development and mechanisms of regulation at comparable stages of maturity are
likely to be similar across species.
In the mature muscle, there are fibre-type differences in the rate of protein turnover that
reflect their compositional differences; slow fibres have higher rates of protein turnover than
fast-twitch fibres (Bark et al., 1998; Dardevet et al., 2002), and this diversity emerges only
upon maturation (Davis et al., 1989). These phenotypic differences are attributable to the
differences in the synthesis rate of muscle proteins in combination with the variation in their
relative abundances among muscles. In skeletal muscles from mature pigs, the average
synthesis rate of mitochondrial proteins is higher than for sarcoplasmic proteins and this, in turn,
tends to be slightly higher than for the myofibrillar proteins (Balagopal et al., 1997; Fiorotto
et al., 2000a). Although the ratio of myofibrillar to sarcoplasmic proteins tends to be greater
in slower muscles (Hemel-Grooten et al., 1995), the difference in synthesis rates is substan-
tially lower than that of mitochondrial proteins, which are more abundant in the slower,
oxidative muscles. The greater overall protein synthetic activity of the slow muscles is sup-
ported by a higher ribosomal abundance and entails minimal changes in protein synthetic
efficiency. In addition to the inherent variation in synthesis rates, the myofibre protein com-
ponents can also differ in their responses to extrinsic stimuli. For example, in adult porcine
muscle, stimulation of protein synthesis by insulin appears to be limited to the mitochondrial
proteins (Boirie et al., 2001); the developmental decline in muscle protein synthesis rates is
dominated by myofibrillar proteins (Fiorotto et al., 2000a). Clearly, these differences among
muscle protein components have to be factored into our understanding of the overall regula-
tion of skeletal muscle protein metabolism.
In the newly differentiated muscle, the high ribosomal abundance is the principal factor
that enables high rates of protein synthesis to be attained, and its reduction with maturation
is one mechanism that may underlie the general reduction in fractional synthesis rates
Regulation of skeletal muscle protein metabolism 47

observed for all muscle proteins. However, this cannot explain the differences in composition
of proteins synthesized, and some regulation must occur at the level of gene transcription.
During the transition from myoblast to myotube, genes encoding non-muscle proteins are
repressed, while those specific for muscle proteins are induced in a co-ordinated manner.
There is then a commensurate change in the composition of proteins expressed (Devlin
and Emerson, 1978, 1979; Shani et al., 1981). In vivo, it has been demonstrated that the
stoichiometry of the total mRNAs encoding all isoforms within a protein family is maintained
accurately, and that production of individual myofibrillar proteins in appropriate stoichio-
metric amounts, therefore, is regulated at the message level (Wade et al., 1990). However,
such changes would not explain the differential responses of sarcoplasmic and myofibrillar
proteins even if the decrease in ribosomal abundance were accompanied by a reduction in the
proportion of myofibrillar mRNAs. Such a change in mRNA composition would increase the
translational efficiency of sarcoplasmic proteins, but decrease the translational efficiency of
myofibrillar proteins. Translational efficiency, however, increases for all proteins in the
immature muscle. Clearly, therefore, although differences in mRNA abundances are involved,
as we have demonstrated in newborn pigs (Fiorotto et al., 2000b), this also cannot be entirely
responsible for the compositional differences in protein synthesis, suggesting that there must
also be regulation of mRNA at the translational and post-translational levels.

3. POSTNATAL MUSCLE GROWTH


3.1. Satellite cells and hyperplasia

Postnatal growth of skeletal muscle is driven by hypertrophy of the existing fibres. This
requires both an increase in myonuclear content, and the accretion of muscle proteins.
Myonuclei are postmitotic and, thus, satellite cells are entirely responsible for the postnatal
increase in muscle fibre DNA. This is clearly demonstrated in mice in which the expression
of Pax-7 is abolished and consequently no satellite cells form (Seale et al., 2000). These
muscles contain both primary and secondary fibres but they fail to hypertrophy postnatally.
Indeed, in a variety of circumstances normally associated with accelerated postnatal muscle
growth, the inhibition of satellite cell replication will prevent the growth response (Rosenblatt
and Parry, 1992).
Recent evidence suggests that subpopulations of satellite cells may exist that can be dis-
tinguished by their proliferative potential (Perry and Rudnicki, 2000; Seale and Rudnicki,
2000). There is a reserve population of quiescent, non-differentiated cells that retains its
mitogenic potential and has the capacity for self-renewal. Under the appropriate stimulation,
these satellite cells become activated, migrate as necessary, and undergo a limited number of
replications before they terminally differentiate. These cells can no longer divide and undergo
fusion into the myofibre. In the rat, satellite cells comprise approximately 32% of muscle
nuclei at birth and decrease to 10% at 4 weeks of age and less than 5% at sexual maturity
when the cells are largely mitotically quiescent (Allbrook et al., 1971). A similar pattern is
seen in the pig, in which satellite cells constitute approximately 20% of total muscle nuclei
at birth, and 4% at 64 weeks of age (Campion et al., 1981; Mesires and Doumit, 2002). These
values vary according to the metabolic properties of the muscle. The significance of myonu-
clear number in the context of muscle growth relates to the observation that in skeletal
muscle, “myonuclear domain” size, i.e. the quantity of cytoplasm regulated by a single
myonucleus (and reflected by the protein:DNA ratio), is tightly regulated (Allen et al., 1999).
This implies that the amount of protein that can be deposited without further addition of
48 T. A. Davis and M. L. Fiorotto

myonuclei is limited. Nuclear domain size under “steady state” conditions appears to vary
according to the metabolic activity of the fibre. It is smaller for oxidative than glycolytic
fibres, and for any given fibre type, values increase with age. Although satellite cells become
quiescent as growth rate plateaus, their proliferation can be reactivated in response to muscle
injury, denervation, or increased muscle stretch and it is essential for muscle repair and hyper-
trophy (Bischoff, 1994).
The close similarity between the developmental changes in satellite cell replication and
protein synthesis strongly suggests that these processes may be linked. This is further sup-
ported by the differential response of immature and mature skeletal muscles to suboptimal
nutrient intakes. Alterations in food intake (greater or less than average) in the neonatal animal,
provided they are not severe, alter DNA and protein accretion proportionally as indicated by
the maintenance of relatively normal, age-appropriate protein:DNA ratios despite a wide range
of growth rates (Fiorotto and Davis, 1997). In transgenic animals, sustained over-expression
of IGF-I in the skeletal muscle promotes satellite cell replication and transiently increases the
accretion of total muscle DNA; this increase precedes the enhanced accumulation of muscle
protein and results in protein:DNA ratios that temporarily are lower than normal (Fiorotto
et al., 2003). These ultimately increase to age-appropriate values, but never surpass those in
wild-type control animals.
A potential link between satellite cell replication and the capacity for protein synthesis is
through the regulation of ribosomal production, ribosomal abundance being the primary
determinant of a cell’s maximal capacity for protein synthesis. Regulation of ribosome
biogenesis is achieved in the majority of cells by altering the rate of rRNA synthesis by rDNA
transcription (Zahradka et al., 1991), the regulation of which is coupled to cell cycling via the
retinoblastoma gene product, pRb (Hannan et al., 2000). We have demonstrated that the
enhanced replicative capacity of satellite cells from muscles that overexpress IGF-I is associ-
ated with increased phosphorylation of pRb upon mitogen stimulation (Chakravarthy et al.,
2001). Thus, when rates of satellite cell division are high, pRB is phosphorylated, and in this
form it enables a key transcription factor for rDNA transcription, UBF, to transactivate rDNA
genes to promote rRNA synthesis. Thus, accretion of ribosomes is necessarily correlated to
the rate of cell division.

3.2. Role of protein synthesis in the regulation of muscle growth

A rapid increase in the absolute rate of growth occurs during early postnatal life and a majority
of this growth is comprised of skeletal muscle protein (Young, 1970). The more rapid accre-
tion of muscle protein than other tissue proteins results in an increase in the proportion of the
body protein pool that is represented by muscle protein from ~30% in the newborn to ~50%
in the adult (fig. 2). However, the fractional rate of growth, i.e. the amount of weight gained
in relation to the existing mass, is extremely high at birth and decreases with development,
with the most rapid change in the fractional rate of growth occurring during the neonatal
period. This developmental decline in the fractional rate of growth is largely explained by a
developmental decline in the fractional rate of protein deposition in skeletal muscle (Shields
et al., 1983; Mitchell et al., 2001).
Changes in the rate of protein deposition are driven by changes in the rates of protein syn-
thesis or protein degradation such that a decline in the fractional rate of protein deposition can
be due to a decline in the fractional rate of protein synthesis, an increase in the fractional rate
of protein degradation, or both. The developmental decline in protein deposition in skeletal
muscle is due to a developmental decline in the fractional rate of muscle protein synthesis
Regulation of skeletal muscle protein metabolism 49

Fig. 2. Relative changes in the proportion of whole-body protein mass attributable to skeletal muscle in the
rat between birth and weaning (Fiorotto et al., unpublished observations).

(Kelly et al., 1984; Denne and Kalhan, 1987; Davis et al., 1989) (fig. 3). In fact, the fractional
rate of muscle protein synthesis in the pig and rat is about 3-fold higher in the newborn than at
weaning, and the rate of decline is attenuated as development proceeds (Kelly et al., 1984; Davis
et al., 1989, 1996; Baillie and Garlick, 1992; Fiorotto et al., 2000a). This developmental decline
in skeletal muscle protein synthesis is more profound in muscles containing predominately
FG fibres than in those containing primarily SO fibres (Davis et al., 1989). By contrast, fractional
protein degradation rates in skeletal muscle decline modestly with development.
The rate of protein synthesis is determined by the abundance of ribosomes, the efficiency
of the translational process, and potentially, the concentration of translatable mRNA (Kimball
and Jefferson, 1988). Because the majority of RNA in tissues is rRNA, ribosomal abundance
can be estimated from the RNA to protein ratio, or can be measured more precisely from the
amount of 18S rRNA expressed per unit protein. The efficiency of the translation process can

Fig. 3. Relationship between the postnatal decline in the rate of muscle protein accretion and the fractional
synthesis rate of skeletal muscle proteins in the hindlimbs of rats. (Data compiled from Davis et al., 1989;
Fiorotto et al., 2000a.)
50 T. A. Davis and M. L. Fiorotto

be calculated from the amount of protein synthesized per unit RNA and reflects how well the
protein synthetic machinery is functioning. Chronic changes in protein synthesis are thought
to be a result of a change in ribosome number. Thus, the high rate of protein synthesis in
immature muscle and its overall decline with development are driven largely by an elevated
number of ribosomes at birth and a developmental decline in ribosome concentration as the
musculature matures (Kelly et al., 1984; Davis et al., 2001). Rapid changes in the rate of pro-
tein synthesis, including those due to food ingestion, are generally regulated by changes in the
efficiency of translation process secondary to modulation of the rate of translation initiation
(Harmon et al., 1984; Kimball and Jefferson, 1988; Kimball et al., 1994), the rate-limiting
step in protein synthesis.
One of the best characterized steps involved in the regulation of translation initiation,
depicted in fig. 4, is the binding of initiator methionyl-tRNA (met-tRNA) to the 40S ribosomal
subunit to form the 43S preinitiation complex via mediation of eukaryotic initiation factor
(eIF) 2 (Pain, 1996; Kimball et al., 1997; Webb and Proud, 1997). The eIF2-mediated met-tRNA
binding to the 40S subunit is further regulated by the activity of eIF2B, which exchanges
GDP for GTP on eIF2 (Kimball et al., 1996). A second well-characterized step in translation
initiation, shown in fig. 4, is the binding of mRNA to the 43S preinitiation complex via medi-
ation of the assembly of the eIF4F complex of proteins (Lin et al., 1994; Rhoads et al., 1994;
Sonenberg, 1994). The three proteins comprising the eIF4F complex are eIF4A, an RNA
helicase, eIF4E, the protein that binds to the m7GTP cap structure at the 5′-end of the mRNA,
and eIF4G, a scaffolding protein that binds to the 40S ribosomal subunit. Thus, mRNA binds
to the 40S ribosomal subunit through the association of eIF4E with eIF4G. The availability
of eIF4E for binding to eIF4G is regulated by its association with 4E-BP1, a repressor pro-
tein that competes with eIF4G for binding to eIF4E (Pause et al., 1994). Upon stimulation by
an anabolic agent, such as insulin, 4E-BP1 becomes phosphorylated, resulting in reduced
affinity of eIF4E for 4E-BP1, release of eIF4E, and enhanced binding of eIF4E to eIF4G to
form the active eIF4E:eIF4G complex (Gingras et al., 1999).
Activation of translation initiation is mediated through a signal transduction pathway
involving a protein kinase referred to as the mammalian target of rapamycin (mTOR) which,

Fig. 4. Regulation of translation initiation. Abbreviations: eIF, eukaryotic initiation factor; 4EBP1, eIF4E
binding protein; Met-tRNA, initiator methionyl tRNA; S6K, 70 kDa ribosomal protein S6 kinase 1; 43S, 43S
ribosomal subunit; 48S, 48S ribosomal subunit.
Regulation of skeletal muscle protein metabolism 51

in addition to phosphorylating 4E-BP1, also phosphorylates and activates the 70 kDa ribosomal
protein S6 kinase, S6K1 (Jefferies et al., 1994; von Manteuffel et al., 1997). These phospho-
rylation events lead to an increase in the rate at which most proteins are synthesized and, in
addition, the preferential increase in translation of mRNAs encoding elements of the transla-
tional apparatus, including ribosomal proteins and elongation factors.
Recent studies in growing pigs suggest that the overall developmental decline in the
response of skeletal muscle protein synthesis to feeding involves regulation by eIF2B (Davis
et al., 2000). Availability of eIF4E for 48S ribosomal complex formation follows a similar
pattern. This response is primarily modulated by the developmental change in the feeding-
induced activation of the factors involved in the binding of mRNA to the 43S preinitiation
complex.

3.3. Role of protein degradation in the regulation of muscle growth

Less is known about the mechanisms that regulate protein degradation than those that regu-
late protein synthesis. It is known that there are multiple pathways in mammalian tissues for
the degradation of proteins and that these pathways are highly controlled and selectively
degrade specific protein substrates. These pathways include the lysosomal–autophagic
system, the calpain–calpastatin system, and the ubiquitin–proteasome system (Goll et al.,
1989; Attaix et al., 1999). The lysosomal–autophagic systems involve primarily cathepsins.
Most evidence suggests that this pathway of degradation is unselective and may be of special
importance under conditions in which cellular proteolysis is maximally activated. The
calpain–calpastatin system is the major calcium-activated pathway of protein degradation.
At least two main calpain isoforms, μ− calpain and m-calpain, have been identified and the
system is subject to inhibition by the protein, calpastatin. The proteases play an important role
in muscle myofibrillar protein turnover by catalysing initial disruption of the structure via
proteolysis at the Z-disc. Released myofilaments can then be degraded into amino acids by
the proteasome and/or lysosomal enzymes (Goll et al., 1992). The ubiquitin–proteasome
pathway is widely distributed among tissues and has a relatively broad protein specificity.
It consists of a recognition system involving the protein ubiquitin, which is responsible for
targeting the protein substrates towards degradation by forming a polyubiquitin complex, and
a multifunctional protease, referred to as the proteasome, which degrades the proteins. The
role of these proteolytic pathways in the regulation of muscle growth and development remains
to be explored.

4. REGULATORS OF PROTEIN SYNTHESIS


4.1. Feeding

Dietary protein is utilized very efficiently for the deposition of whole-body protein during
early postnatal life (Pellett and Kaba, 1972; McCracken et al., 1980; Fiorotto et al., 1991;
Davis et al., 1993a). The accumulated evidence suggests that young animals utilize their
dietary amino acids more efficiently for growth because they are capable of a greater increase
in muscle protein synthesis in response to feeding than older animals (Davis et al., 1991,
1996). Feeding stimulates protein synthesis in the whole body of the newborn human (Denne
et al., 1991); in skeletal muscle of the suckling lamb (Oddy et al., 1987; Wester et al., 2000);
and in skeletal muscle of the post-weaned, but still growing, rat (Garlick et al., 1983).
However, the stimulation of muscle protein synthesis by feeding is blunted or absent in adult
mammals (Melville et al., 1989; Baillie and Garlick, 1992; Tessari et al., 1996).
52 T. A. Davis and M. L. Fiorotto

Fig. 5. Stimulation of protein synthesis by feeding decreases with development in neonatal pigs.

In the suckling pig (Davis et al., 1996, 1997; Burrin et al., 1997a) and rat (Davis et al.,
1991, 1993b), protein synthesis in skeletal muscle is maximally stimulated after eating.
Figure 5 shows that the postprandial rise in protein synthesis in skeletal muscle of the neona-
tal pig declines sharply during the first 4 weeks of life. Although feeding stimulates
protein synthesis in all tissues of the neonatal animal, the magnitude and the developmental
decline in the response to feeding are most pronounced in skeletal muscle (Burrin et al., 1991,
1995, 1997a; Davis et al., 1991, 1993b, 1996). This enhanced ability of skeletal muscle protein
synthesis to respond to the provision of nutrients in young growing animals should not be
surprising, because the rate of protein deposition during the postprandial period must be
higher than the rate of protein loss during the postabsorptive period to permit growth of skeletal
muscle.
Recent studies have examined the developmental changes in the expression and activation
of factors that regulate the feeding-induced stimulation of protein synthesis in skeletal muscle
of young, growing pigs. The results show that eIF2B activity, which regulates the binding of
met-tRNA to the 40S ribosomal subunit, is unaffected by feeding but decreases with devel-
opment. The stimulation of muscle protein synthesis by feeding, and the developmental
decline in this response, involve regulation by the eIF4F complex (Davis et al., 2000; Kimball
et al., 2002). In skeletal muscle of the neonatal pig, feeding increases the phosphorylation
of 4E-BP1, resulting in dissociation of the inactive 4E-BP1 . eIF4E complex, and increased
association of the active eIF4E . eIF4G complex. This response leads to a global increase in
the rate of muscle protein synthesis. These feeding-induced changes in the activity of factors
that regulate eIF4F formation decrease with development in parallel with the developmental
change in the feeding-induced stimulation of muscle protein synthesis. A response to feeding
has been observed in “teenage” rats (Yoshizawa et al., 1997). However, the magnitude of the
response is smaller than that in neonatal pigs, thus further supporting a developmental decline
in the feeding-induced formation of the eIF4F complex.
The developmental changes in the feeding-induced eIF4F activation occur in parallel with
increased phosphorylation of S6K1, which is involved in the translation of mRNAs encoding
specific proteins that regulate translation initiation (Davis et al., 2000; Kimball et al., 2002).
An increased phosphorylation of both 4E-BP1 and S6K1 suggests involvement of the mTOR
signalling pathway in this process. Furthermore, rapamycin, a specific inhibitor of mTOR,
Regulation of skeletal muscle protein metabolism 53

strongly attenuates the feeding-induced assembly of both eIF4F and S6K1 activation
(Kimball et al., 2000). Thus, the enhanced activation of the eIF4F complex following food
consumption likely plays an important role in the postprandial stimulation of muscle protein
synthesis in growing animals and the efficient use of dietary amino acids for muscle protein
deposition in the neonate.

4.2. Insulin

Studies performed in incubated muscles and in perfused hindlimbs of growing animals clearly
demonstrate that insulin stimulates protein synthesis (Jefferson et al., 1977; Davis et al., 1987;
Kimball et al., 1994). The infusion of physiological concentrations of insulin in fasted,
weaned rats stimulates muscle protein synthesis in vivo to rates similar to those found in the
fed state (Garlick et al., 1983). This response to feeding can be blocked by co-administration
of anti-insulin serum (Preedy and Garlick, 1986). Furthermore, insulin has been shown to
stimulate whole-body amino acid utilization and protein synthesis in the fetal sheep (Liechty
et al., 1992; Thureen et al., 2000), protein synthesis in hindlimb of the young lamb (Wester
et al., 2000), and skeletal muscle protein synthesis in the weaned rat (Garlick et al., 1983). In
marked contrast to studies conducted in growing animals, most studies in adult animals
(Baillie and Garlick, 1992; McNulty et al., 1993) and humans (Gelfand et al., 1987; Heslin
et al., 1992; Louard et al., 1992) show little, if any, response of muscle protein synthesis to
physiological increases in insulin. This suggests that the response of muscle protein synthe-
sis to insulin is developmentally regulated.
Insulin plays a key role in the increased response of skeletal muscle protein synthesis to feed-
ing, and thus the increased rate of protein deposition, during the early postnatal period. In fasted
and fed neonatal pigs, there is a positive curvilinear relationship between the postprandial
increase in fractional muscle protein synthesis rates and circulating insulin concentrations
(Davis et al., 1997). Studies using a hyperinsulinemic–euglycemic–euaminoacidemic clamp
technique show that when amino acids and glucose are maintained at fasting levels, insulin
infusion increases amino acid disposal, and that the insulin sensitivity and responsiveness of
amino acid disposal decrease with development (Wray-Cahen et al., 1997). This response
suggests that the developmental change in the insulin sensitivity of whole-body amino acid
disposal may underlie the developmental change in the efficiency of utilization of dietary
amino acids for protein deposition. Furthermore, raising insulin concentrations in the neonatal
pig to levels typical of the fed state increases the rate of skeletal muscle protein synthesis to
within the range normally present in the fed state, even when amino acids and glucose are
maintained at fasting levels (Wray-Cahen et al., 1998). This response to insulin, like the
response to feeding, is attenuated with development and is greater in muscles that are com-
posed primarily of FG fibres, and is not specific to myofibrillar proteins (Davis et al., 2001).
The insulin signalling cascade (fig. 6) leading to the stimulation of protein synthesis is
initiated by insulin binding to its receptor. This leads to autophosphorylation of the receptor,
the activation of insulin receptor tyrosine kinase, and the subsequent phosphorylation of
several cytosolic substrates including insulin receptor substrate (IRS)-1 and -2 (Sun et al.,
1991; White and Kahn, 1994). IRS-1 and -2 serve as “docking proteins”, transmitting insulin
signals to several proteins that contain Src-homology 2 (SH2) domains (Backer et al., 1992;
Sun et al., 1993) including phosphatidylinositol (PI) 3-kinase, which catalyses the phosphory-
lation of PI. The activation of PI 3-kinase triggers the activation of components of the insulin
signalling pathway leading to translation initiation, i.e. protein kinase B (Akt) and mTOR.
54 T. A. Davis and M. L. Fiorotto

Fig. 6. Insulin signalling pathway leading to translation initiation.

Studies focusing on the developmental changes in the insulin signalling pathway that leads
to translation initiation have shown that in the pig the abundance of insulin receptor protein
in muscle during the early suckling period is 2-fold higher than at weaning (Suryawan et al.,
2001). Although the abundance of IRS-1 and IRS-2 does not change with development, the
abundance of the downstream signalling proteins, protein kinase B and mTOR, decreases
with development (Kimball et al., 2002). This developmental decline in the abundance of
insulin receptor, protein kinase B, and mTOR in skeletal muscle likely contributes to the over-
all decline in the responsiveness of muscle protein synthesis to feeding that occurs over the
course of development.
Because insulin mediates the postprandial elevation in skeletal muscle protein synthesis
and this response decreases with development (Wray-Cahen et al., 1998; Davis et al., 2001),
it is not surprising that the feeding-induced activation of the insulin signalling pathway that
regulates protein synthesis decreases with development. Thus, the feeding-induced activation
of the insulin receptor, IRS-1, IRS-2, PI 3-kinase, and protein kinase B in skeletal muscle
decreases with development (Suryawan et al., 2001; Kimball et al., 2002), in parallel with
the developmental decline in the feeding-induced activation of translation initiation factors
and protein synthesis (Davis et al., 1996). This suggests that the developmental decline in
the postprandial stimulation of protein synthesis in skeletal muscle results from a reduction
in the capacity of the intracellular insulin signalling pathway to transduce to the translational
apparatus the stimulus provided by the feeding-induced rise in insulin and/or amino acid
concentrations.
A number of studies performed in cell culture, in the perfused hindlimb, and in intact growing
rats have demonstrated that the stimulation of protein synthesis by insulin involves increased
phosphorylation of the translational repressor protein, 4E-BP1, reduced interaction with
eIF4E, and increased assembly of the mRNA cap-binding complex, eIF4G:eIF4E (Kimball et al.,
1994, 1997). Furthermore, insulin increases phosphorylation of S6K1, thereby increasing the
translation of specific proteins involved in the regulation of translation. Recent in vivo studies
performed in neonatal pigs support these findings and further show that the insulin-induced
Regulation of skeletal muscle protein metabolism 55

changes in factors regulating translation initiation as well as the upstream components of the
insulin signalling pathway occur in a dose-response manner within the physiological range
(Suryawan et al., 2001; O’Connor et al., 2003). Recently, however, studies in adult rats
suggest that while insulin increases the phosphorylation of S6K1, insulin does not alter
4E-BP1 phosphorylation (Long et al., 2000). This lack of effect of insulin on 4E-BP1 phos-
phorylation and, by inference, eIF4F formation, is not surprising as physiological
hyperinsulinaemia has no effect on muscle protein synthesis in adults. Thus, insulin plays an
important role in the regulation of protein synthesis in muscle of growing animals, but its
importance during adulthood is less apparent.

4.3. Amino acids

Although amino acids are the precursors for the synthesis of proteins, they also play a key
role as nutritional signals in the regulation of muscle protein synthesis. Amino acids have the
capability to stimulate muscle protein synthesis throughout a substantial part of the life cycle,
in contrast to the developmental decline and loss of the capability of insulin to stimulate
muscle protein synthesis with age. In weaned but still growing rats (Preedy and Garlick,
1986), adult humans and rats (Bennet et al., 1990; McNulty et al., 1993; Vary et al., 1999),
and elderly people (Volpi et al., 1998), acute amino acid infusion, either alone or concurrent
with insulin infusion, stimulates protein synthesis in skeletal muscle. Recent studies suggest,
however, that the magnitude of the stimulation of muscle protein synthesis by amino acids
may decrease in the early postnatal period (Davis et al., 2002a).
When a balanced amino acid mixture is infused into fasted, growing pigs, muscle protein
synthesis increases and this response to amino acid infusion decreases with development,
in parallel with the developmental decline in the feeding-induced stimulation of skeletal
muscle protein synthesis (Davis et al., 2002a). In young pigs, the stimulation of skeletal
muscle protein synthesis by amino acids is greater in muscles that contain predominately
FG muscle fibres than in those that contain primarily SO fibres, and is similar for myofibril-
lar and sarcoplasmic proteins. The response to amino acid infusion occurs when insulin levels
either remain at the fasting level or are raised to the fed level by infusion (O’Connor et al.,
2003). Indeed, the magnitude of the increase in muscle protein synthesis with amino acid
stimulation is similar to that which occurs with insulin stimulation alone, implying that
insulin and amino acids may be interacting with the same signalling pathway within skeletal
muscle.
Studies performed in cell culture have shown that amino acid availability modulates pro-
tein synthesis by regulating both the met-tRNA and mRNA binding steps of translation
initiation (Fox et al., 1998; Hara et al., 1998; Kimball et al., 1998; Patti et al., 1998; Jefferson
and Kimball, 2001). In vivo studies in mature, food-deprived rats in which a large oral dose
of leucine was administered suggest that in muscle, leucine promotes the binding of eIF4G to
eIF4E, increases the phosphorylation of 4E-BP1, and represses the association of eIF4E with
4E-BP1 (Anthony et al., 2000). In neonatal pigs, raising amino acids from the fasting to the
fed levels in the presence of insulin produced a similar response (O’Connor et al., 2003).
In the absence of insulin, amino acids do not affect either the phosphorylation of S6K1 and
4E-BP1, or the association of eIF4E with 4E-BP1 and eIF4G, even though they stimulate
muscle protein synthesis. This suggests that amino acids stimulate muscle protein synthesis
in growing animals by modulating the availability of eIF4E for 48S ribosomal complex for-
mation, and by processes that do not require enhanced assembly of the mRNA cap-binding
complex.
56 T. A. Davis and M. L. Fiorotto

4.4. Insulin-like growth factors

Many (Douglas et al., 1991; Fryburg et al., 1995; Bark et al., 1998; Vary et al., 2000; Davis
et al., 2002b), although not all (Oddy and Owens, 1996; Boyle et al., 1998), studies have
demonstrated an anabolic effect of IGF-I on protein synthesis in skeletal muscle. However, in
some studies reductions in circulating concentrations of amino acids, insulin, and/or glucose
during the administration of IGF-I may have limited the ability of IGF-I to stimulate protein
synthesis. Thus, when amino acids, glucose, and insulin are maintained at fasting levels, infu-
sion of IGF-I to the level seen in the fed state stimulates muscle protein synthesis in growing
swine (Davis et al., 2002b).
IGF-I, however, is unlikely to play a role in the feeding-induced stimulation of muscle
protein synthesis. First, in contrast to insulin, the rise in circulating IGF-I after feeding
is not immediate (Buonomo and Baile, 1991; Goldstein et al., 1991; Davis et al., 1993b, 1996;
Svanberg et al., 1996). Second, the postprandial changes in muscle protein synthesis in
young animals are positively correlated with changes in circulating insulin, but not IGF-I,
concentrations (Davis et al., 1997, 1998). Third, with development, circulating IGF-I levels
increase, whereas skeletal muscle protein synthesis rates decrease (Davis et al., 1996).
Although circulating IGF-I is unlikely to be a physiologically significant regulator of
the feeding-induced stimulation of skeletal muscle protein synthesis, this does not negate
the potential role of IGF-I as a long-term regulator of growth, as has been suggested by
others (Buonomo and Baile, 1991; Donovan et al., 1991; VandeHaar et al., 1991), or the
potential usefulness of IGF-I as an anabolic agent to enhance protein deposition as discussed
previously.
IGF-I likely stimulates protein synthesis in skeletal muscle by acting on the same signalling
pathway as insulin that leads to translation initiation (Dardevet et al., 1996; Vary et al., 2000).
The receptors for both IGF-I and insulin share considerable homology of structure and func-
tion (Ullrich et al., 1986; Cheatham and Kahn, 1995; LeRoith et al., 1995) and both hormones
act on some of the same intracellular signalling pathways (Dardevet et al., 1996; Suryawan
et al., 2001). Furthermore, both insulin and IGF-I stimulate protein synthesis by increasing
the formation of the active eIF4E . eIF4G complex that regulates the binding of mRNA to the
ribosome (Kimball et al., 1997; Vary et al., 2000).

4.5. Growth hormone

Growth hormone treatment increases protein deposition, improves nitrogen retention, and
enhances the efficiency with which dietary protein is utilized for growth (Campbell et al.,
1990; Caperna et al., 1991; Vann et al., 2000a). Furthermore, GH treatment profoundly
decreases the synthesis and excretion of urea, and the oxidation of amino acids. Whole-body
protein balance is improved in response to GH treatment due to the minimization of protein
loss during fasting, and maximization of protein gain during meal absorption (Vann et al.,
2000b). GH treatment in GH-deficient (Bier, 1991; Russell-Jones et al., 1998) and normal,
mature animals and adult humans (Eisemann et al., 1989; Pell et al., 1990; Fryburg et al.,
1991; Bell et al., 1998) increases protein deposition by stimulating whole-body and skeletal
muscle protein synthesis. Chronic GH treatment in cattle and swine increases amino acid
uptake by the hindquarter (Boisclair et al., 1994; Bush et al., 2003a) and protein synthesis in
muscle (Eisemann et al., 1989; Seve et al., 1993) with no change in protein degradation across
the hindlimb (Bush et al., 2003a).
Regulation of skeletal muscle protein metabolism 57

In young, growing swine, GH treatment increases skeletal muscle protein synthesis in the
postprandial state, but not in the fasting condition. This increase is due to modulation of trans-
lational efficiency by GH and not by ribosome number (Bush et al., 2003b). The GH-induced
increase in translation initiation is attributable to modulation of the factors associated with the
binding of both mRNA and met-tRNA to the ribosomal complex, that is, the phosphorylation
of 4E-BP1, association of eIF4E with eIF4G, and eIF2B activity. Because GH increases
circulating IGF-I and insulin concentrations and this increase is greater in the fed than in the
fasting state, the GH-induced increase in protein synthesis may involve mediation by IGF-I
and/or insulin, or may be due to a direct effect of GH. In fact, GH indirectly activates some
of the same signalling components as insulin and IGF-I, i.e. IRS-1 and -2, PI 3-kinase, protein
kinase B, and S6K1 (Anderson, 1993; Yenush and White, 1997). In addition, the increased
substrate availability, i.e. amino acids, provided in the fed condition may be permissive for
the GH-induced increase in muscle protein synthesis.

4.6. Colostrum

Colostrum provides a rich source of nutrients for the newborn mammal that supports the rapid
growth and accretion of body protein during the first few days of postnatal life (Burrin et al.,
1997b). In addition to nutrients, colostrum also contains maternal immunoglobulins that for
many species are essential for passive immunity, and a variety of bioactive components that
include insulin, IGF-I, IGF-II, and epidermal growth factor. Although the benefits of the con-
sumption of nutrients and immune factors are readily apparent, the functional significance of
the numerous hormones and growth factors present in colostrum is unclear. Studies that have
compared the growth of newborns have demonstrated an enhanced anabolic response in asso-
ciation with the feeding of colostrum, especially of the visceral organs (Widdowson and
Crabb, 1976; Widdowson et al., 1976). Given their mitogenic and anabolic properties, this
response was often attributed to the presence of trophic factors in the colostrum. However, it
must also be considered that the consumption of colostrum entails the ingestion of a larger
quantity of nutrients than that typically provided by mature milk or, indeed, many formulas.
Studies designed to distinguish between the trophic effects of macronutrient intake and
those due to factors in colostrum (Burrin et al., 1995; Fiorotto et al., 2000b) showed that in
newborn pigs, feeding stimulates protein synthesis in all tissues, but the stimulation of pro-
tein synthesis in skeletal muscle is greater when colostrum, as opposed to a nutrient-matched
formula or mature sow’s milk, is fed. This suggests that the enhanced stimulation of skeletal
muscle protein synthesis in newborn pigs fed colostrum, as opposed to other feeds, is not due
solely to the provision of macronutrients. Furthermore, the stimulation of protein synthesis
by colostrum feeding was restricted specifically to the myofibrillar proteins, unlike the
general stimulation of protein synthesis by feeding which incurred a proportional stimulation
of the synthesis of both sarcoplasmic and myofibrillar proteins (Fiorotto et al., 2000b).
Feeding also resulted in a general increase in muscle mRNA concentration, but in the
colostrum-fed piglets the enhanced synthesis rate of myofibrillar proteins was associated with
a disproportionate increase in the abundance of myofibrillar mRNA, as exemplified by total
MHC mRNAs. Additionally, colostrum augmented the effect of feeding on protein
synthesis by promoting a greater accretion of ribosomes.
Thus, feeding colostrum has both quantitative consequences for the anabolic process in the
skeletal musculature of the newborn animal and qualitative consequences, with potential
implications for the development of muscle function. Improvement of skeletal muscle
58 T. A. Davis and M. L. Fiorotto

function is advantageous insofar as it is critical for the development of the newborn’s ability
to survive independently from its mother. The effects observed are likely attributable to non-
nutritive factors present in colostrum, although these have not yet been identified. However,
a number of potential factors, including insulin, IGF-I, thyroid hormone, and growth hor-
mone, have been excluded. Identification of the mechanisms underlying this phenomenon
will be critical for advancing our understanding of the biological role of early mammary
secretions in the regulation of neonatal growth and in establishing how diet contributes to the
regulation of skeletal muscle growth in early postnatal life.

5. FUTURE PERSPECTIVES
There are numerous issues concerning the regulation of skeletal muscle growth and meta-
bolism that need to be explored further. From the point of view of agriculture, the relative
significance of these is determined by the economic benefits to be gained. The ultimate aim
is to enhance feed efficiency. This, however, needs to be accomplished without compromis-
ing meat quality, especially tenderness and fat content. However, it is becoming increasingly
evident that consumers are becoming more resistant to the use by the livestock industry of
anabolic agents, growth promoters, and antibiotics, and frequently are prepared to pay a pre-
mium for products in which they have not been used. Although one may question the validity
of these concerns, it must be acknowledged that they are widespread and, therefore, should
not be ignored. In this regard, the application of genomics and proteomics to select for breed-
ing stock with desirable traits, and improved husbandry practices to reduce mortality and
morbidity in the birth to weaning period, are likely to be the most productive approaches.
Given the large increase in fish consumption, research on the growth and composition of
muscle of different fish varieties deserves substantially more attention.
Enhancement of muscle growth can be accomplished either by increasing the number of
myofibres, or by promoting myofibre hypertrophy. As should be evident, the former is a
prenatal event and is dictated by maternal and genetic factors. Thus, continued research on
the regulation of cell cycle progression and withdrawal of individual myoblast lineages, as
well as the factors that control terminal differentiation, are likely to yield relevant information,
especially when this can be merged with genomic trait analysis. Because mechanistic studies
are difficult to conduct in vivo, in normal animals, much of the basic research on these mech-
anisms must be performed in cell and tissue culture. However, the widespread use of
genetically engineered mice has been most productive and helpful in this regard because,
although far removed from livestock animals, transgenic mice provide an important and appo-
site tool with which to assess the relevance of specific cellular events in the context of the
whole animal. “The Myostatin Knockout Story” presents an excellent example of the useful-
ness of this approach.
For some mammalian species, the ability to increase muscle fibre number has limitations,
however, because the enhanced fetal growth may increase maternal morbidity and/or com-
promise maternal lactational capacity. In principle, therefore, enhanced postnatal muscle
hypertrophy would be preferable. As we have presented, postnatal hypertrophy is dictated by
two key factors, satellite cell number and protein accretion. Satellite cell number is the bal-
ance between the continued division of these cells acquired during the third phase of myoblast
determination, and their loss, either by terminal differentiation and fusion into the myofibre,
or by apoptosis. Although substantial progress has been made in understanding the factors
that regulate satellite cell division, there are still many unanswered questions. Satellite cells
have their origins in the latter part of fetal life and, therefore, are likely to be influenced
Regulation of skeletal muscle protein metabolism 59

by maternal variables; however, relatively little is known about the influence of maternal
physiology and metabolism on the satellite cells of the progeny. The factors, especially envi-
ronmental factors, that dictate terminal differentiation and apoptosis of satellite cells, and the
extent to which these processes can be manipulated through husbandry practices and diet, are
much less clearly understood, and warrant a closer examination as they are likely to have
long-term consequences.
More recently it is has been demonstrated that under certain in vitro conditions, satellite
cells can change lineage and form adipocytes. Clearly this has significant consequences not
only for overall muscle growth potential, but also for the composition of meat. The extent to
which this occurs in vivo, and the conditions that would favour such a change, merit further
attention.
The rate of protein accretion is the balance between protein synthesis and degradation. We
have demonstrated that in young animals the rate of protein synthesis is the principal regula-
tory factor. This unique feature is attributable to the ability of the immature muscle to
markedly increase translation when food is available. The latter is critical because it enables
amino acids to be diverted towards protein synthesis rather than to be oxidized. Thus, dietary
protein can be used with greater efficiency, provided the composition of amino acids and
energy intake are optimal. Thus, from the nutritional standpoint, the ability to meter the
amino acid composition of dietary protein to meet the needs for growth versus maintenance
during development can enable maximal exploitation of this high synthetic capacity of the
immature muscle. The characteristics of the immature muscle that enable this synthetic
response are primarily its high ribosomal content and an enhanced sensitivity and respon-
siveness of muscle protein synthesis to insulin. Clearly, therefore, further understanding of
those factors that are responsible for these unique features of the immature muscle, and their
down-regulation with maturation, would be warranted. Moreover, it is equally important that
the impact of environmental variables such as infection, temperature, activity (duration, type,
and intensity), and dietary nutrients other than protein and energy (e.g. micronutrients,
modified lipids, and various non-nutritive factors present in foodstuffs) on protein synthesis
during this anabolic phase of growth be investigated.
Our emphasis on protein synthesis rather than degradation does not negate the importance
of the latter in the regulation of protein accretion. Indeed, the regulation of protein degrada-
tion potentially represents a much more energetically efficient approach for improving the
efficiency of muscle protein deposition (Goll et al., 1989) beyond the early postnatal period.
However, much less is understood about the in vivo regulation of protein degradation, espe-
cially the factors that regulate myofibrillar breakdown, and the variability in these mechanisms
between muscles and among different species. In addition to the consequences for protein
deposition, protein degradation has consequences for meat quality because those enzymes
that are responsible for the degradation of muscle myofibrillar proteins are also important
determinants of post-mortem meat tenderisation. The evidence would suggest that in domestic
animals muscle hypertrophy resulting from suppression of protein degradation in vivo can
compromise meat tenderness. Examples of this negative consequence of suppressing protein
degradation is the callipyge lamb in which the degree of hypertrophy of certain muscles is
positively correlated to calpastatin expression, and increased toughness. The effects of some
β-adrenergic agonists in certain species is similar to the effect of the callipyge gene. Thus, a
clear understanding of the interplay between the structural characteristics of a muscle, the rel-
ative contribution of protein synthesis versus degradation to its overall growth, the variation
among species, and how these aspects of muscle structure and metabolism are influenced by
environmental factors and husbandry practices, are subjects that merit further study.
60 T. A. Davis and M. L. Fiorotto

Recently there has been much effort refocused on skeletal muscle metabolism in humans
with the recognition that there is an inevitable depletion of skeletal muscle with ageing
(sarcopenia). The consequence of this loss is quite severe as it results in a loss of strength,
flexibility, and overall mobility, which thereby compromises the individual’s quality of life.
The resulting decrease in activity not only exacerbates the muscle loss, but also decreases
basal and activity-related energy expenditure, which therefore enhances the propensity for
excessive fat deposition and glucose intolerance. The causes of sarcopenia appear to be exten-
sive and include the loss in the replicative capacity of satellite cells, age-related increases in
factors that are antagonistic to muscle growth, such as myostatin and Id factors, and a loss in
the body’s capacity to produce anabolic agents such as growth hormone and testosterone. The
relative importance of these, however, is far from clear. Additionally, or possibly in conse-
quence to these changes, skeletal muscle loses its regenerative capacity with ageing. As the
average life expectancy of humans increases, understanding the causes of sarcopenia and the
development of therapies and modalities to mitigate its occurrence has enormous economic
implications. Importantly, from the metabolic perspective, a better appreciation of the nutri-
ent needs and dietary regimens that are required to sustain optimal muscle metabolism are
warranted.

REFERENCES
Adams, G.R., McCue, S.A., Zeng, M., Baldwin, K.M., 1999. Time course of myosin heavy chain transi-
tions in neonatal rats: importance of innervation and thyroid state. Amer. J. Physiol. 276, R954–R961.
Allbrook, D.B., Han, M.F., Hellmuth, A.E., 1971. Population of muscle satellite cells in relation to age
and mitotic activity. Pathology 3, 233−243.
Allen, D.L., Leinwand, L.A., 2002. Intracellular calcium and myosin isoform transitions: calcineurin and
calcium-calmodulin kinase pathways regulate preferential activation of the IIa myosin heavy chain
promoter. J. Biol. Chem. 277, 45323−45330.
Allen, D.L., Roy, R.R., Edgerton, V.R., 1999. Myonuclear domains in muscle adaptation and disease.
Muscle Nerve 22, 1350−1360.
Anderson, N.G., 1993. Simultaneous activation of p90rsk and p70s6k S6 kinases by growth hormone in
3T3-F442A preadipocytes. Biochem. Biophys. Res. Commun. 193, 284−290.
Anthony, J.C., Anthony, T.G., Kimball, S.R., Vary, T.C., Jefferson, L.S., 2000. Orally administered
leucine stimulates protein synthesis in skeletal muscle of postabsorptive rats in association with
increased eIF4F formation. J. Nutr. 130, 139−145.
Attaix, D., Combaret, L., Taillandier, D., 1999. Mechanisms and regulation in protein degradation.
In: Lobley, G.E., White, A., MacRae, J.C. (Eds.), Protein Metabolism and Nutrition. Wageningen
Press, Wageningen, The Netherlands, pp. 51−67.
Backer, J.M., Myers, M.G., Jr., Shoelson, S.E., Chin, D.J., Sun, X.J., Miralpeix, M., Hu, P., Margolis, B.,
Skolnik, E.Y., Schlessinger, J., 1992. Phosphatidylinositol 3′-kinase is activated by association with
IRS-1 during insulin stimulation. EMBO J. 11, 3469−3479.
Baillie, A.G., Garlick, P.J., 1992. Attenuated responses of muscle protein synthesis to fasting and insulin
in adult female rats. Amer. J. Physiol. 262, E1−E5.
Balagopal, P., Rooyackers, O.E., Adey, D.B., Ades, P.A., Nair, K.S., 1997. Effects of aging on in vivo
synthesis of skeletal muscle myosin heavy-chain and sarcoplasmic protein in humans. Amer. J.
Physiol. 273, E790−E800.
Barany, M., Close, R.I., 1971. The transformation of myosin in cross-innervated rat muscles. J. Physiol.
213, 455−474.
Bark, T.H., McNurlan, M.A., Lang, C.H., Garlick, P.J., 1998. Increased protein synthesis after acute IGF-I
or insulin infusion is localized to muscle in mice. Amer. J. Physiol. 275, E118−E123.
Bell, A.W., Bauman, D.E., Beermann, D.H., Harrell, R.J., 1998. Nutrition, development and efficacy of
growth modifiers in livestock species. J. Nutr. 128, 360S−363S.
Benezra, R., Davis, R.L., Lockshon, D., Turner, D.L., Weintraub, H., 1990. The protein Id: a negative
regulator of helix-loop-helix DNA binding proteins. Cell 61, 49−59.
Regulation of skeletal muscle protein metabolism 61

Bennet, W.M., Connacher, A.A., Scrimgeour, C.M., Rennie, M.J., 1990. The effect of amino acid infu-
sion on leg protein turnover assessed by L-[15N]phenylalanine and L-[1-13C]leucine exchange. Eur. J.
Clin. Invest. 20, 41−50.
Bier, D.M., 1991. Growth hormone and insulin-like growth factor I: nutritional pathophysiology and ther-
apeutic potential. Acta Paediatr. Scand., Suppl. 374, 119−128.
Birchmeier, C., Brohmann, H., 2000. Genes that control the development of migrating muscle precursor
cells. Curr. Opin. Cell Biol. 12, 725−730.
Bischoff, R., 1994. The satellite cell and muscle regeneration. In: Engel, A.G., Franzini-Armstrong, C. (Eds.),
Myology, McGraw-Hill, New York, Vol. 1, pp. 97−118.
Boirie, Y., Short, K.R., Ahlman, B., Charlton, M., Nair, K.S., 2001. Tissue-specific regulation of
mitochondrial and cytoplasmic protein synthesis rates by insulin. Diabetes 50, 2652−2658.
Boisclair, Y.R., Bauman, D.E., Bell, A.W., Dunshea, F.R., Harkins, M., 1994. Nutrient utilization
and protein turnover in the hindlimb of cattle treated with bovine somatotropin. J. Nutr. 124,
664−673.
Boyle, D.W., Denne, S.C., Moorehead, H., Lee, W.H., Bowsher, R.R., Liechty, E.A., 1998. Effect of
rhIGF-I infusion on whole fetal and fetal skeletal muscle protein metabolism in sheep. Amer.
J. Physiol. 275, E1082−E1091.
Brohmann, H., Jagla, K., Birchmeier, C., 2000. The role of Lbx1 in migration of muscle precursor cells.
Development 127, 437−445.
Buckingham, M., 2001. Skeletal muscle formation in vertebrates. Curr. Opin. Genet. Dev. 11, 440−448.
Buckingham, M., Bajard, L., Chang, T., Daubas, P., Hadchouel, J., Meilhac, S., Montarras, D.,
Rocancourt, D., Relaix, F., 2003. The formation of skeletal muscle: from somite to limb. J. Anat. 202,
59−68.
Buonomo, F.C., Baile, C.A., 1991. Influence of nutritional deprivation on insulin-like growth factor I,
somatotropin, and metabolic hormones in swine. J. Anim. Sci. 69, 755−760.
Burrin, D.G., Davis, T.A., Ebner, S., Schoknecht, P.A., Fiorotto, M.L., Reeds, P.J., 1997a. Colostrum
enhances the nutritional stimulation of vital organ protein synthesis in neonatal pigs. J. Nutr. 127,
1284−1289.
Burrin, D.G., Davis, T.A., Ebner, S., Schoknecht, P.A., Fiorotto, M.L., Reeds, P.J., McAvoy, S., 1995.
Nutrient-independent and nutrient-dependent factors stimulate protein synthesis in colostrum-fed
newborn pigs. Pediat. Res. 37, 593−599.
Burrin, D.G., Davis, T.A., Fiorotto, M.L., Reeds, P.J., 1991. Stage of development and fasting affect
protein synthetic activity in the gastrointestinal tissues of suckling rats. J. Nutr. 121, 1099−1108.
Burrin, D.G., Davis, T.A., Fiorotto, M.L., Reeds, P.J., 1997b. Role of milk-borne vs endogenous insulin-
like growth factor I in neonatal growth. J. Anim. Sci. 75, 2739−2743.
Bush, J.A., Burrin, D.G., Suryawan, A., O’Connor, P.M., Nguyen, H.V., Reeds, P.J., Steele, N.C.,
van Goudoever, J.B., Davis, T.A., 2003a. Somatotropin-induced protein anabolism in hindquarters
and portal-drained viscera of growing pigs. Amer. J. Physiol. Endocrinol. Metab. 284, E302−E312.
Bush, J.A., Kimball, S.R., O’Connor, P.M., Suryawan, A., Orellana, R.A., Nguyen, H.V., Jefferson, L.S.,
Davis, T.A., 2003b. Translational control of protein synthesis in muscle and liver of growth hormone-
treated pigs. Endocrinology 144, 1273−1283.
Calvo, S., Vullhorst, D., Venepally, P., Cheng, J., Karavanova, I., Buonanno, A., 2001. Molecular dissec-
tion of DNA sequences and factors involved in slow muscle-specific transcription. Mol. Cell Biol. 21,
8490−8503.
Campbell, R.G., Johnson, R.J., King, R.H., Taverner, M.R., Meisinger, D.J., 1990. Interaction of dietary
protein content and exogenous porcine growth hormone administration on protein and lipid accretion
rates in growing pigs. J. Anim. Sci. 68, 3217−3225.
Campion, D.R., Richardson, R.L., Reagan, J.O., Kraeling, R.R., 1981. Changes in the satellite cell pop-
ulation during postnatal growth of pig skeletal muscle. J. Anim Sci. 52, 1014−1018.
Caperna, T.J., Komarek, D.R., Gavelek, D., Steele, N.C., 1991. Influence of dietary protein and recom-
binant porcine somatotropin administration in young pigs: II. Accretion rates of protein, collagen, and
fat. J. Anim. Sci. 69, 4019−4029.
Chakravarthy, M.V., Fiorotto, M.L., Schwartz, R.J., Booth, F.W., 2001. Long-term insulin-like growth
factor-I expression in skeletal muscles attenuates the enhanced in vitro proliferation ability of the
resident satellite cells in transgenic mice. Mech. Ageing Dev. 122, 1303−1320.
Cheatham, B., Kahn, C.R., 1995. Insulin action and the insulin signaling network. Endocr. Rev. 16,
117−142.
62 T. A. Davis and M. L. Fiorotto

Coolican, S.A., Samuel, D.S., Ewton, D.Z., McWade, F.J., Florini, J.R., 1997. The mitogenic and myo-
genic actions of insulin-like growth factors utilize distinct signaling pathways. J. Biol. Chem. 272,
6653−6662.
Dardevet, D., Sornet, C., Bayle, G., Prugnaud, J., Pouyet, C., Grizard, J., 2002. Postprandial stimulation
of muscle protein synthesis in old rats can be restored by a leucine-supplemented meal. J. Nutr. 132,
95−100.
Dardevet, D., Sornet, C., Vary, T., Grizard, J., 1996. Phosphatidylinositol 3-kinase and p70 S6 kinase par-
ticipate in the regulation of protein turnover in skeletal muscle by insulin and insulin-like growth
factor I. Endocrinology 137, 4087−4094.
Davis, T.A., Burrin, D.G., Fiorotto, M.L., Nguyen, H.V., 1996. Protein synthesis in skeletal muscle
and jejunum is more responsive to feeding in 7- than 26-day-old pigs. Amer. J. Physiol. 270,
E802−E809.
Davis, T.A., Burrin, D.G., Fiorotto, M.L., Reeds, P.J., Jahoor, F., 1998. Roles of insulin and amino acids
in the regulation of protein synthesis in the neonate. J. Nutr. 128, 347S−350S.
Davis, T.A., Fiorotto, M.L., Beckett, P.R., Burrin, D.G., Reeds, P.J., Wray-Cahen, D., Nguyen, H.V.,
2001. Differential effects of insulin on peripheral and visceral tissue protein synthesis in neonatal
pigs. Amer. J. Physiol. Endocrinol. Metab. 280, E770−E779.
Davis, T.A., Fiorotto, M.L., Burrin, D.G., Pond, W.G., Nguyen, H.V., 1997. Intrauterine growth restric-
tion does not alter response of protein synthesis to feeding in newborn pigs. Amer. J. Physiol. 272,
E877−E884.
Davis, T.A., Fiorotto, M.L., Burrin, D.G., Reeds, P.J., Nguyen, H.V., Beckett, P.R., Vann, R.C.,
O’Connor, P.M., 2002a. Stimulation of protein synthesis by both insulin and amino acids is unique to
skeletal muscle in neonatal pigs. Amer. J. Physiol. Endocrinol. Metab. 282, E880−E890.
Davis, T.A., Fiorotto, M.L., Burrin, D.G., Vann, R.C., Reeds, P.J., Nguyen, H.V., Beckett, P.R., Bush, J.A.,
2002b. Acute IGF-I infusion stimulates protein synthesis in skeletal muscle and other tissues of
neonatal pigs. Amer. J. Physiol. Endocrinol. Metab. 283, E638−E647.
Davis, T.A., Fiorotto, M.L., Nguyen, H.V., Burrin, D.G., Reeds, P.J., 1991. Response of muscle protein
synthesis to fasting in suckling and weaned rats. Amer. J. Physiol. 261, R1373−R1380.
Davis, T.A., Fiorotto, M.L., Nguyen, H.V., Reeds, P.J., 1989. Protein turnover in skeletal muscle of suckling
rats. Amer. J. Physiol. 257, R1141−R1146.
Davis, T.A., Fiorotto, M.L., Nguyen, H.V., Reeds, P.J., 1993b. Enhanced response of muscle protein
synthesis and plasma insulin to food intake in suckled rats. Amer. J. Physiol. 265, R334−R340.
Davis, T.A., Fiorotto, M.L., Reeds, P.J., 1993a. Amino acid compositions of body and milk protein
change during the suckling period in rats. J. Nutr. 123, 947−956.
Davis, T.A., Klahr, S., Karl, I.E., 1987. Insulin-stimulated protein metabolism in chronic azotemia and
exercise. Amer. J. Physiol. 253, F164−F169.
Davis, T.A., Nguyen, H.V., Suryawan, A., Bush, J.A., Jefferson, L.S., Kimball, S.R., 2000.
Developmental changes in the feeding-induced stimulation of translation initiation in muscle of
neonatal pigs. Amer. J. Physiol. Endocrinol. Metab. 279, E1226−E1234.
Denne, S.C., Kalhan, S.C., 1987. Leucine metabolism in human newborns. Amer. J. Physiol. 253,
E608−E615.
Denne, S.C., Rossi, E.M., Kalhan, S.C., 1991. Leucine kinetics during feeding in normal newborns.
Pediat. Res. 30, 23−27.
Devlin, R.B., Emerson, C.P. Jr., 1978. Coordinate regulation of contractile protein synthesis during
myoblast differentiation. Cell 13, 599−611.
Devlin, R.B., Emerson, C.P. Jr., 1979. Coordinate accumulation of contractile protein mRNAs during
myoblast differentiation. Dev. Biol. 69, 202−216.
Donovan, S.M., Atilano, L.C., Hintz, R.L., Wilson, D.M., Rosenfeld, R.G., 1991. Differential regulation
of the insulin-like growth factors (IGF-I and -II) and IGF binding proteins during malnutrition in the
neonatal rat. Endocrinology 129, 149−157.
Douglas, R.G., Gluckman, P.D., Ball, K., Breier, B., Shaw, J.H., 1991. The effects of infusion of insulin-
like growth factor (IGF) I, IGF-II, and insulin on glucose and protein metabolism in fasted lambs.
J. Clin. Invest. 88, 614−622.
Ebner, S., Schoknecht, P., Reeds, P., Burrin, D., 1994. Growth and metabolism of gastrointestinal and
skeletal muscle tissues in protein-malnourished neonatal pigs. Amer. J. Physiol. 266, R1736−R1743.
Eisemann, J.H., Hammond, A.C., Rumsey, T.S., 1989. Tissue protein synthesis and nucleic acid concen-
trations in steers treated with somatotropin. Brit. J. Nutr. 62, 657−671.
Regulation of skeletal muscle protein metabolism 63

Engert, J.C., Berglund, E.B., Rosenthal, N., 1996. Proliferation precedes differentiation in IGF-I-
stimulated myogenesis. J. Cell Biol. 135, 431−440.
Epstein, J.A., Shapiro, D.N., Cheng, J., Lam, P.Y., Maas, R.L., 1996. Pax3 modulates expression of the
c-Met receptor during limb muscle development. Proc. Natl. Acad. Sci. USA 93, 4213−4218.
Fiorotto, M.L., Davis, T.A., 1997. Food intake alters muscle protein gain with little effect on Na+–K+-
ATPase and myosin isoforms in suckled rats. Amer. J. Physiol. 272, R1461−R1471.
Fiorotto, M.L., Burrin, D.G., Perez, M., Reeds, P.J., 1991. Intake and use of milk nutrients by rat pups
suckled in small, medium, or large litters. Amer. J. Physiol. 260, R1104−R1113.
Fiorotto, M.L., Davis, T.A., Reeds, P.J., 2000a. Regulation of myofibrillar protein turnover during matu-
ration in normal and undernourished rat pups. Amer. J. Physiol. 278, 845−854.
Fiorotto, M.L., Davis, T.A., Reeds, P.J., Burrin, D.G., 2000b. Nonnutritive factors in colostrum enhance
myofibrillar protein synthesis in the newborn pig. Pediat. Res. 48, 1−7.
Fiorotto, M.L., Schwartz, R.J., Delaughter, M.C., 2003. Persistent IGF-I overexpression in skeletal
muscle transiently enhances DNA accretion and growth. FASEB J. 17, 59−60.
Fox, H.L., Pham, P.T., Kimball, S.R., Jefferson, L.S., Lynch, C.J., 1998. Amino acid effects on transla-
tional repressor 4E-BP1 are mediated primarily by L-leucine in isolated adipocytes. Amer. J. Physiol.
275, C1232−C1238.
Francis-West, P.H., Antoni, L., Anakwe, K., 2003. Regulation of myogenic differentiation in the devel-
oping limb bud. J. Anat. 202, 69−81.
Fryburg, D.A., Gelfand, R.A., Barrett, E.J., 1991. Growth hormone acutely stimulates forearm muscle
protein synthesis in normal humans. Amer. J. Physiol. 260, E499−E504.
Fryburg, D.A., Jahn, L.A., Hill, S.A., Oliveras, D.M., Barrett, E.J., 1995. Insulin and insulin-like growth
factor-I enhance human skeletal muscle protein anabolism during hyperaminoacidemia by different
mechanisms. J. Clin. Invest. 96, 1722−1729.
Fuchtbauer, E.M., 2002. Inhibition of skeletal muscle development: less differentiation gives more
muscle. Results Probl. Cell Differ. 38, 143−161.
Gambke, B., Lyons, G.E., Haselgrove, J., Kelly, A.M., Rubinstein, N.A., 1983. Thyroidal and neural
control of myosin transitions during development of rat fast and slow muscles. FEBS Lett. 156,
335−339.
Garlick, P.J., Fern, M., Preedy, V.R., 1983. The effect of insulin infusion and food intake on muscle pro-
tein synthesis in postabsorptive rats. Biochem. J. 210, 669−676.
Gelfand, R.A., Barrett, E.J., 1987. Effect of physiologic hyperinsulinemia on skeletal muscle protein syn-
thesis and breakdown in man. J. Clin. Invest. 80, 1−6.
Gingras, A.C., Raught, B., Sonenberg, N., 1999. eIF4 initiation factors: effectors of mRNA recruitment
to ribosomes and regulators of translation. Annu. Rev. Biochem. 68, 913−963.
Goldstein, S., Harp, J.B., Phillips, L.S., 1991. Nutrition and somatomedin. XXII: Molecular regulation
of insulin-like growth factor-I during fasting and refeeding in rats. J. Mol. Endocrinol. 6, 33−43.
Goll, D.E., Kleese, W.C., Szapchenko, A., 1989. Skeletal muscle proteases and protein turnover. In:
Campion, D.R., Hausman, G.J., Martin, R.J. (Eds.), Animal Growth Regulation. Plenum Publishing
Company, New York, pp. 141−182.
Goll, D.E., Thompson, V.F., Taylor, R.G., Christiansen, J.A., 1992. Role of the calpain system in muscle
growth. Biochimie 74, 225−237.
Grobet, L., Martin, L.J., Poncelet, D., Pirottin, D., Brouwers, B., Riquet, J., Schoeberlein, A., Dunner, S.,
Menissier, F., Massabanda, J., Fries, R., Hanset, R., Georges, M., 1997. A deletion in the bovine myo-
statin gene causes the double-muscled phenotype in cattle. Nat. Genet. 17, 71−74.
Handschin, C., Rhee, J., Lin, J., Tarr, P.T., Spiegelman, B.M., 2003. An autoregulatory loop controls per-
oxisome proliferator-activated receptor gamma coactivator 1α expression in muscle. Proc. Natl. Acad.
Sci. USA 100, 7111−7116.
Hannan, K.M., Kennedy, B.K., Cavanaugh, A.H., Hannan, R.D., Hirschler-Laszkiewicz, I., Jefferson, L.S.,
Rothblum, L.I., 2000. RNA polymerase I transcription in confluent cells: Rb downregulates rDNA
transcription during confluence-induced cell cycle arrest. Oncogene 19, 3487−3497.
Hara, K., Yonezawa, K., Weng, Q.P., Kozlowski, M.T., Belham, C., Avruch, J., 1998. Amino acid suffi-
ciency and mTOR regulate p70 S6 kinase and eIF-4E BP1 through a common effector mechanism.
J. Biol. Chem. 273, 14484−14494.
Harmon, C.S., Proud, C.G., Pain, V.M., 1984. Effects of starvation, diabetes and acute insulin
treatment on the regulation of polypeptide-chain initiation in rat skeletal muscle. Biochem. J. 223,
687−696.
64 T. A. Davis and M. L. Fiorotto

Harrison, A.P., Tivey, D.R., Clausen, T., Duchamp, C., Dauncey, M.J., 1996. Role of thyroid hormones
in early postnatal development of skeletal muscle and its implications for undernutrition. Brit. J. Nutr.
76, 841−855.
Hemel-Grooten, H.N., Koohmaraie, M., Yen, J.T., Arbona, J.R., Rathmacher, J.A., Nissen, S.L.,
Fiorotto, M.L., Garssen, G.J., Verstegen, M.W., 1995. Comparison between 3-methylhistidine
production and proteinase activity as measures of skeletal muscle breakdown in protein-deficient
growing barrows. J. Anim. Sci. 73, 2272−2281.
Henriksson, J., 1990. The possible role of skeletal muscle in the adaptation to periods of energy
deficiency. Eur. J. Clin. Nutr. 44, Suppl. 1, 55−64.
Heslin, M.J., Newman, E., Wolf, R.F., Pisters, P.W., Brennan, M.F., 1992. Effect of hyperinsulinemia
on whole body and skeletal muscle leucine carbon kinetics in humans. Amer. J. Physiol. 262,
E911−E918.
Irrcher, I., Adhihetty, P.J., Sheehan, T., Joseph, A.M., Hood, D.A., 2003. PPAR-γ coactivator-1α expres-
sion during thyroid hormone- and contractile activity-induced mitochondrial adaptations. Amer.
J. Physiol. Cell Physiol. 284, C1669−C1677.
Jefferies, H.B., Reinhard, C., Kozma, S.C., Thomas, G., 1994. Rapamycin selectively represses transla-
tion of the “polypyrimidine tract” mRNA family. Proc. Natl. Acad. Sci. USA 91, 4441−4445.
Jefferson, L.S., Kimball, S.R., 2001. Amino acid regulation of gene expression. J. Nutr. 131,
2460S−2466S.
Jefferson, L.S., Li, J.B., Rannels, S.R., 1977. Regulation by insulin of amino acid release and protein
turnover in the perfused rat hemicorpus. J. Biol. Chem. 252, 1476−1483.
Katsumata, M., Cattaneo, D., White, P., Burton, K.A., Dauncey, M.J., 2000. Growth hormone receptor
gene expression in porcine skeletal and cardiac muscles is selectively regulated by postnatal under-
nutrition. J. Nutr. 130, 2482−2488.
Kelly, F.J., Lewis, S.E., Anderson, P., Goldspink, D.F., 1984. Pre- and postnatal growth and protein
turnover in four muscles of the rat. Muscle Nerve 7, 235−242.
Kimball, S.R., Jefferson, L.S., 1988. Cellular mechanisms involved in the action of insulin on protein
synthesis. Diabetes Metab. Rev. 4, 773−787.
Kimball, S.R., Farrell, P.A., Nguyen, H.V., Jefferson, L.S., Davis, T.A., 2002. Developmental decline in
components of signal transduction pathways regulating protein synthesis in pig muscle. Amer.
J. Physiol. Endocrinol. Metab. 282, E585−E592.
Kimball, S.R., Horetsky, R.L., Jefferson, L.S., 1998. Signal transduction pathways involved in the regu-
lation of protein synthesis by insulin in L6 myoblasts. Amer. J. Physiol. 274, C221−C228.
Kimball, S.R., Jefferson, L.S., Nguyen, H.V., Suryawan, A., Bush, J.A., Davis, T.A., 2000. Feeding stim-
ulates protein synthesis in muscle and liver of neonatal pigs through an mTOR-dependent process.
Amer. J. Physiol. Endocrinol. Metab. 279, E1080−E1087.
Kimball, S.R., Jurasinski, C.V., Lawrence, J.C. Jr., Jefferson, L.S., 1997. Insulin stimulates protein
synthesis in skeletal muscle by enhancing the association of eIF-4E and eIF-4G. Amer. J. Physiol.
272, C754−C759.
Kimball, S.R., Mellor, H., Flowers, K.M., Jefferson, L.S., 1996. Role of translation initiation factor
eIF-2B in the regulation of protein synthesis in mammalian cells. Prog. Nucleic Acid Res. Mol. Biol.
54, 165−196.
Kimball, S.R., Vary, T.C., Jefferson, L.S., 1994. Regulation of protein synthesis by insulin. Annu. Rev.
Physiol. 56, 321−348.
Koohmaraie, M., 2003. Meat tenderness and muscle growth: is there relationship? Meat Sci. 62,
345−352.
Langley, B., Thomas, M., Bishop, A., Sharma, M., Gilmour, S., Kambadur, R., 2002. Myostatin
inhibits myoblast differentiation by down-regulating MyoD expression. J. Biol. Chem. 277,
49831−49840.
Lee, S.J., McPherron, A.C., 1999. Myostatin and the control of skeletal muscle mass. Curr. Opin. Genet.
Dev. 9, 604−607.
LeRoith, D., Werner, H., Beitner-Johnson, D., Roberts, C.T. Jr., 1995. Molecular and cellular aspects of
the insulin-like growth factor I receptor. Endocr. Rev. 16, 143−163.
Liechty, E.A., Boyle, D.W., Moorehead, H., Liu, Y.M., Denne, S.C., 1992. Effect of hyperinsuline-
mia on ovine fetal leucine kinetics during prolonged maternal fasting. Amer. J. Physiol. 263,
E696−E702.
Regulation of skeletal muscle protein metabolism 65

Lin, T.A., Kong, X., Haystead, T.A., Pause, A., Belsham, G., Sonenberg, N., Lawrence, J.C. Jr., 1994.
PHAS-I as a link between mitogen-activated protein kinase and translation initiation. Science 266,
653−656.
Long, W., Saffer, L., Wei, L., Barrett, E.J., 2000. Amino acids regulate skeletal muscle PHAS-I and p70
S6-kinase phosphorylation independently of insulin. Amer. J. Physiol. Endocrinol. Metab. 279,
E301−E306.
Louard, R.J., Fryburg, D.A., Gelfand, R.A., Barrett, E.J., 1992. Insulin sensitivity of protein and glucose
metabolism in human forearm skeletal muscle. J. Clin. Invest. 90, 2348−2354.
Louveau, I., Le Dividich, J., 2002. GH and IGF-I binding in adipose tissue, liver, and skeletal muscle in
response to milk intake level in piglets. Gen. Comp. Endocrinol. 126, 310−317.
McCracken, K.J., Eddie, S.M., Stevenson, W.G., 1980. Energy and protein nutrition of early-weaned
pigs. 1. Effect of energy intake and energy:protein on growth, efficiency and nitrogen utilization of
pigs between 8–32 days. Brit. J. Nutr. 43, 289−304.
McKinsey, T.A., Zhang, C.L., Olson, E.N., 2001. Control of muscle development by dueling HATs and
HDACs. Curr. Opin. Genet. Dev. 11, 497−504.
McKinsey, T.A., Zhang, C.L., Olson, E.N., 2002. Signaling chromatin to make muscle. Curr. Opin. Cell
Biol. 14, 763−772.
McNulty, P.H., Young, L.H., Barrett, E.J., 1993. Response of rat heart and skeletal muscle protein in vivo
to insulin and amino acid infusion. Amer. J. Physiol. 264, E958−E965.
McPherron, A.C., Lawler, A.M., Lee, S.J., 1997. Regulation of skeletal muscle mass in mice by a new
TGF-β superfamily member. Nature 387, 83−90.
Melville, S., McNurlan, M.A., McHardy, K.C., Broom, J., Milne, E., Calder, A.G., Garlick, P.J.,
1989. The role of degradation in the acute control of protein balance in adult man: failure of
feeding to stimulate protein synthesis as assessed by L-[1-13C]leucine infusion. Metabolism 38,
248−255.
Mesires, N.T., Doumit, M.E., 2002. Satellite cell proliferation and differentiation during postnatal growth
of porcine skeletal muscle. Amer. J. Physiol. Cell Physiol. 282, C899−C906.
Mitchell, A.D., Scholz, A.M., Mersmann, H.J., 2001. Growth and body composition. In: Pond, W.G.,
Mersmann, H.J. (Eds.), Biology of the Domestic Pig. Comstock Publishing Association, Ithaca, NY,
pp. 225−308.
Mitchell, P.J., Johnson, S.E., Hannon, K., 2002. Insulin-like growth factor I stimulates myoblast
expansion and myofiber development in the limb. Dev. Dyn. 223, 12−23.
Molkentin, J.D., Olson, E.N., 1996. Combinatorial control of muscle development by basic
helix-loop-helix and MADS-box transcription factors. Proc. Natl. Acad. Sci. USA 93,
9366−9373.
Musaro, A., McCullagh, K., Paul, A., Houghton, L., Dobrowolny, G., Molinaro, M., Barton, E.R.,
Sweeney, H.L., Rosenthal, N., 2001. Localized Igf-1 transgene expression sustains hypertrophy and
regeneration in senescent skeletal muscle. Nat. Genet. 27, 195–200.
Nakayama, M., Stauffer, J., Cheng, J., Banerjee-Basu, S., Wawrousek, E., Buonanno, A., 1996. Common
core sequences are found in skeletal muscle slow- and fast-fiber-type-specific regulatory elements.
Mol. Cell Biol. 16, 2408−2417.
O’Connor, P.M., Bush, J.A., Suryawan, A., Nguyen, H.V., Davis, T.A., 2003. Insulin and amino acids
independently stimulate skeletal muscle protein synthesis in neonatal pigs. Amer. J. Physiol.
Endocrinol. Metab. 284, E110−E119.
Oddy, V.H., Owens, P.C., 1996. Insulin-like growth factor I inhibits degradation and improves retention
of protein in hindlimb muscle of lambs. Amer. J. Physiol. 271, E973−E982.
Oddy, V.H., Lindsay, D.B., Barker, P.J., Northrop, A.J., 1987. Effect of insulin on hind-limb and whole-
body leucine and protein metabolism in fed and fasted lambs. Brit. J. Nutr. 58, 437−452.
Ojuka, E.O., Jones, T.E., Han, D.H., Chen, M., Holloszy, J.O., 2003. Raising Ca2+ in L6 myotubes
mimics effects of exercise on mitochondrial biogenesis in muscle. FASEB J. 17, 675−681.
Ontell, M., 1982. The growth and metabolism of developing muscle. In: Jones, C.T. (Ed.),
Biochemical Development of the Fetus and Neonate. Elsevier Biomedical Press, Amsterdam,
pp. 213−247.
Ontell, M., Dunn, R.F., 1978. Neonatal muscle growth: a quantitative study. Amer. J. Anat. 152,
539−556.
Pain, V.M., 1996. Initiation of protein synthesis in eukaryotic cells. Eur. J. Biochem. 236, 747−771.
66 T. A. Davis and M. L. Fiorotto

Patti, M.E., Brambilla, E., Luzi, L., Landaker, E.J., Kahn, C.R., 1998. Bidirectional modulation of insulin
action by amino acids. J. Clin. Invest 101, 1519−1529.
Pause, A., Belsham, G.J., Gingras, A.C., Donze, O., Lin, T.A., Lawrence, J.C. Jr., Sonenberg, N.,
1994. Insulin-dependent stimulation of protein synthesis by phosphorylation of a regulator of 5′-cap
function. Nature 371, 762−767.
Pell, J.M., Elcock, C., Harding, R.L., Morrell, D.J., Simmonds, A.D., Wallis, M., 1990. Growth, body
composition, hormonal and metabolic status in lambs treated long-term with growth hormone. Brit.
J. Nutr. 63, 431−445.
Pellett, P.L., Kaba, H., 1972. Carcass amino acids of the rat under conditions of determination of net
protein utilization. J. Nutr. 102, 61−68.
Perry, R.L., Rudnicki, M.A., 2000. Molecular mechanisms regulating myogenic determination and
differentiation. Front Biosci. 5, D750−D767.
Preedy, V.R., Garlick, P.J., 1986. The response of muscle protein synthesis to nutrient intake in postab-
sorptive rats: the role of insulin and amino acids. Biosci. Rep. 6, 177−183.
Rehfeldt, C., Fiedler, I., Weikard, R., Kanitz, E., Ender, K., 1993. It is possible to increase skeletal muscle
fibre number in utero. Biosci. Rep. 13, 213−220.
Rehfeldt, C., Stickland, N.C., Fiedler, I., Wegner, J., 1999. Environmental and genetic factors as sources
of variation in skeletal muscle fibre number. Basic Appl. Myol. 9, 235−253.
Rhoads, R.E., Joshi, B., Minich, W.B., 1994. Participation of initiation factors in the recruitment of
mRNA to ribosomes. Biochimie 76, 831−838.
Rosenblatt, J.D., Parry, D.J., 1992. Gamma irradiation prevents compensatory hypertrophy of overloaded
mouse extensor digitorum longus muscle. J. Appl. Physiol. 73, 2538−2543.
Russell-Jones, D.L., Bowes, S.B., Rees, S.E., Jackson, N.C., Weissberger, A.J., Hovorka, R., Sonksen, P.H.,
Umpleby, A.M., 1998. Effect of growth hormone treatment on postprandial protein metabolism in
growth hormone-deficient adults. Amer. J. Physiol. 274, E1050−E1056.
Scaal, M., Bonafede, A., Dathe, V., Sachs, M., Cann, G., Christ, B., Brand-Saberi, B., 1999. SF/HGF is
a mediator between limb patterning and muscle development. Development 126, 4885−4893.
Schiaffino, S., Margreth, A., 1969. Coordinated development of the sarcoplasmic reticulum and T system
during postnatal differentiation of rat skeletal muscle. J. Cell Biol. 41, 855−875.
Schiaffino, S., Reggiani, C., 1996. Molecular diversity of myofibrillar proteins: gene regulation and
functional significance. Physiol. Rev. 76, 371−423.
Schiaffino, S., Serrano, A., 2002. Calcineurin signaling and neural control of skeletal muscle fiber type
and size. Trends Pharmacol. Sci. 23, 569−575.
Seale, P., Rudnicki, M.A., 2000. A new look at the origin, function, and “stem-cell” status of muscle
satellite cells. Dev. Biol. 218, 115−124.
Seale, P., Sabourin, L.A., Girgis-Gabardo, A., Mansouri, A., Gruss, P., Rudnicki, M.A., 2000. Pax7 is
required for the specification of myogenic satellite cells. Cell 102, 777−786.
Seve, B., Ballevre, O., Ganier, P., Noblet, J., Prugnaud, J., Obled, C., 1993. Recombinant porcine
somatotropin and dietary protein enhance protein synthesis in growing pigs. J. Nutr. 123,
529−540.
Shani, M., Zevin-Sonkin, D., Saxel, O., Carmon, Y., Katcoff, D., Nudel, U., Yaffe, D., 1981. The
correlation between the synthesis of skeletal muscle actin, myosin heavy chain, and myosin light
chain and the accumulation of corresponding mRNA sequences during myogenesis. Dev. Biol. 86,
483−492.
Shields, R.G. Jr., Mahan, D.C., Graham, P.L., 1983. Changes in swine body composition from birth to
145 kg. J. Anim. Sci. 57, 43−54.
Sonenberg, N., 1994. Regulation of translation and cell growth by eIF-4E. Biochimie 76, 839−846.
Spangenburg, E.E., Booth, F.W., 2003. Molecular regulation of individual skeletal muscle fibre types.
Acta Physiol. Scand. 178, 413−424.
Sun, X.J., Crimmins, D.L., Myers, M.G. Jr., Miralpeix, M., White, M.F., 1993. Pleiotropic insulin signals
are engaged by multisite phosphorylation of IRS-1. Mol. Cell Biol. 13, 7418−7428.
Sun, X.J., Rothenberg, P., Kahn, C.R., Backer, J.M., Araki, E., Wilden, P.A., Cahill, D.A., Goldstein, B.J.,
White, M.F., 1991. Structure of the insulin receptor substrate IRS-1 defines a unique signal transduc-
tion protein. Nature 352, 73−77.
Suryawan, A., Nguyen, H.V., Bush, J.A., Davis, T.A., 2001. Developmental changes in the feeding-
induced activation of the insulin-signaling pathway in neonatal pigs. Amer. J. Physiol. Endocrinol.
Metab. 281, E908−E915.
Regulation of skeletal muscle protein metabolism 67

Svanberg, E., Zachrisson, H., Ohlsson, C., Iresjo, B.M., Lundholm, K.G., 1996. Role of insulin and IGF-I
in activation of muscle protein synthesis after oral feeding. Amer. J. Physiol. 270, E614−E620.
Tessari, P., Zanetti, M., Barazzoni, R., Vettore, M., Michielan, F., 1996. Mechanisms of postprandial
protein accretion in human skeletal muscle: insight from leucine and phenylalanine forearm kinetics.
J. Clin. Invest. 98, 1361−1372.
Thureen, P.J., Scheer, B., Anderson, S.M., Tooze, J.A., Young, D.A., Hay, W.W. Jr., 2000. Effect of hyper-
insulinemia on amino acid utilization in the ovine fetus. Amer. J. Physiol. Endocrinol. Metab. 279,
E1294−E1304.
Ullrich, A., Gray, A., Tam, A.W., Yang-Feng, T., Tsubokawa, M., Collins, C., Henzel, W., Le Bon, T.,
Kathuria, S., Chen, E., 1986. Insulin-like growth factor I receptor primary structure: comparison
with insulin receptor suggests structural determinants that define functional specificity. EMBO J. 5,
2503−2512.
VandeHaar, M.J., Moats-Staats, B.M., Davenport, M.L., Walker, J.L., Ketelslegers, J.M., Sharma, B.K.,
Underwood, L.E., 1991. Reduced serum concentrations of insulin-like growth factor-I (IGF-I) in
protein-restricted growing rats are accompanied by reduced IGF-I mRNA levels in liver and skeletal
muscle. J. Endocrinol. 130, 305−312.
Vann, R.C., Nguyen, H.V., Reeds, P.J., Burrin, D.G., Fiorotto, M.L., Steele, N.C., Deaver, D.R., Davis, T.A.,
2000a. Somatotropin increases protein balance by lowering body protein degradation in fed, growing
pigs. Amer. J. Physiol. Endocrinol. Metab. 278, E477−E483.
Vann, R.C., Nguyen, H.V., Reeds, P.J., Steele, N.C., Deaver, D.R., Davis, T.A., 2000b. Somatotropin
increases protein balance independent of insulin’s effects on protein metabolism in growing pigs.
Amer. J. Physiol. Endocrinol. Metab. 279, E1−E10.
Vary, T.C., Jefferson, L.S., Kimball, S.R., 1999. Amino acid-induced stimulation of translation initiation
in rat skeletal muscle. Amer. J. Physiol. 277, E1077−E1086.
Vary, T.C., Jefferson, L.S., Kimball, S.R., 2000. Role of eIF4E in stimulation of protein synthesis by IGF-I
in perfused rat skeletal muscle. Amer. J. Physiol. Endocrinol. Metab. 278, E58−E64.
Venuti, J.M., Morris, J.H., Vivian, J.L., Olson, E.N., Klein, W.H., 1995. Myogenin is required for late but
not early aspects of myogenesis during mouse development. J. Cell Biol. 128, 563−576.
Volpi, E., Ferrando, A.A., Yeckel, C.W., Tipton, K.D., Wolfe, R.R., 1998. Exogenous amino acids
stimulate net muscle protein synthesis in the elderly. J. Clin. Invest. 101, 2000−2007.
von Manteuffel, S.R., Dennis, P.B., Pullen, N., Gingras, A.C., Sonenberg, N., Thomas, G., 1997. The
insulin-induced signalling pathway leading to S6 and initiation factor 4E binding protein 1 phospho-
rylation bifurcates at a rapamycin-sensitive point immediately upstream of P70S6K. Mol. Cell Biol.
17, 5426−5436.
Wade, R., Sutherland, C., Gahlmann, R., Kedes, L., Hardeman, E., Gunning, P., 1990. Regulation of
contractile protein gene family mRNA pool sizes during myogenesis. Dev. Biol. 142, 270−282.
Webb, B.L., Proud, C.G., 1997. Eukaryotic initiation factor 2B (eIF2B). Int. J. Biochem. Cell Biol. 29,
1127−1131.
Wester, T.J., Lobley, G.E., Birnie, L.M., Lomax, M.A., 2000. Insulin stimulates phenylalanine uptake
across the hind limb in fed lambs. J. Nutr. 130, 608−611.
White, M.F., Kahn, C.R., 1994. The insulin signaling system. J. Biol. Chem. 269, 1−4.
White, P., Cattaneo, D., Dauncey, M.J., 2000. Postnatal regulation of myosin heavy chain isoform expres-
sion and metabolic enzyme activity by nutrition. Brit. J. Nutr. 84, 185−194.
Widdowson, E.M., Crabb, D.E., 1976. Changes in the organs of pigs in response to feeding for the first
24h after birth. I. The internal organs and muscles. Biol. Neonate 28, 261−271.
Widdowson, E.M., Colombo, V.E., Artavanis, C.A., 1976. Changes in the organs of pigs in response to
feeding for the first 24h after birth. II. The digestive tract. Biol. Neonate 28, 272−281.
Wray-Cahen, D., Beckett, P.R., Nguyen, H.V., Davis, T.A., 1997. Insulin-stimulated amino acid utiliza-
tion during glucose and amino acid clamps decreases with development. Amer. J. Physiol. 273,
E305−E314.
Wray-Cahen, D., Nguyen, H.V., Burrin, D.G., Beckett, P.R., Fiorotto, M.L., Reeds, P.J., Wester, T.J.,
Davis, T.A., 1998. Response of skeletal muscle protein synthesis to insulin in suckling pigs decreases
with development. Amer. J. Physiol. 275, E602−E609.
Yates, L.D., Greaser, M.L., 1983. Quantitative determination of myosin and actin in rabbit skeletal
muscle. J. Mol. Biol. 168, 123−141.
Yenush, L., White, M.F., 1997. The IRS-signalling system during insulin and cytokine action. Bioessays
19, 491−500.
68 T. A. Davis and M. L. Fiorotto

Yoshizawa, F., Kimball, S.R., Jefferson, L.S., 1997. Modulation of translation initiation in rat skeletal
muscle and liver in response to food intake. Biochem. Biophys. Res. Commun. 240, 825−831.
Young, V.R., 1970. The role of skeletal and cardiac muscle in the regulation of protein metabolism. In:
Munro, H.M. (Ed.), Mammalian Protein Metabolism. Academic Press, New York, pp. 585−674.
Zahradka, P., Larson, D.E., Sells, B.H., 1991. Regulation of ribosome biogenesis in differentiated rat
myotubes. Mol. Cell Biochem. 104, 189−194.
3 Whole animal and tissue proteolysis
in growing animals

V. E. Baracos

Department of Oncology, University of Alberta, 11560 University Avenue,


Edmonton, Alberta, Canada T6G 1Z2

While it is convenient to conceptualize protein synthesis as being associated with growth and pro-
tein degradation as being associated only with atrophy or senescence, both processes proceed
continuously in all tissues at all stages of life. Since the highest rates of protein degradation occur
during the most rapid growth, protein deposition is inefficient. Our current understanding of pro-
teolytic processes comes in large part from studies of skeletal muscle, including methods and
approaches for its determination, contributing proteases and regulators, physiological controls,
and post-mortem proteolytic events contributing to meat quality. Because of the contribution of
protein catabolism to deposition of marketable muscle tissue, to its energetic cost, and to product
quality, there is interest in different strategies increasing or modulating the rate of animal growth,
especially the relative rates of protein synthesis and catabolism in skeletal muscle.

1. PROTEIN DEGRADATION: A KEY DETERMINANT OF GROWTH,


METABOLIC RATE, AND GROWTH EFFICIENCY
This review covers a variety of topics pertaining to protein degradation and regulation of growth
in animals. Numerous excellent reviews are cited, and I have not attempted to cover in detail
domains for which recent synthesis articles are available. Not all of the work relevant to the
topic of growth and protein degradation has been conducted in domestic animal species. One of
the reasons for this is that there are many difficulties associated with measures of protein degra-
dation and these approaches are generally more difficult to implement in domestic animals
because of their cost, invasiveness, and need for the use of stable or radioactive isotopes and
related analytical equipment not routinely available in Animal Science Departments, such as
isotope ratio mass spectrometry (Patterson et al., 1997). The reader is thus strongly encouraged
to be open-minded to the broader scope of the protein degradation literature, in all species.
A fundamental concept in metabolism is that all proteins are in a continual state of turnover.
While protein synthesis is associated with growth and protein degradation is a dominant fea-
ture of atrophy or senescence, both processes proceed continuously. In fact the highest rates

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
69 © 2005 Elsevier Limited. All rights reserved.
70 V. E. Baracos

of protein degradation occur during the most rapid growth and protein deposition. This means
that a large fraction of proteins synthesized in any period of growth are broken back down.
The amount of protein formed and broken down again over any period of time is dependent
on developmental stage, species, and organ. For example, protein breakdown rates in indi-
vidual tissues of animals and poultry undergoing high rates of growth vary from <10%/day
to >80%/day. This means that in a matter of 10 days, an animal at this stage will have broken
down and resynthesized a slowly turning over tissue once and a rapidly turning over tissue
eight times! The vast majority of animal protein formed, thus never goes to market.
The results of Lapierre et al. (1999) further illustrate the impact of protein degradation
across different growth rates. These authors evaluated whole-body protein metabolism in rela-
tion to intake (0.6, 1.0, and 1.6 × maintenance requirements) in growing beef steers. Protein
retention in the whole body increased with intake, as a result of a greater increase in protein
synthesis compared with protein degradation. Protein breakdown had a major impact, as 65%
of the protein synthesized was degraded when intake varied from 1.0 to 1.6 times maintenance.
Protein degradation is sometimes determined on a tissue- or organ-specific basis (i.e. Biolo
et al., 1994; Samuels and Baracos, 1996; Zhang et al., 1996a,b; Lapierre et al., 1999), although
relatively few tissue-specific determinations have been done in domestic species. Where this has
been looked at, there is considerable emphasis on the skeletal muscles, since this organ is the
main product of meat animal agriculture. Splanchnic tissues drained by the portal and hepatic
veins are organ sites amenable to determination of protein degradation by tracer techniques (i.e.
Lapierre et al., 1999). These organ systems are also a considerable focus, since because of their
high turnover rates they constitute a major fraction of whole-body catabolism. In spite of its rel-
atively small size, the liver in a growing monogastric can comprise 25% of whole-body protein
catabolism. In growing cattle, the total splanchnic tissues inclusive of the liver accounted for 44%
of whole-body turnover (25% from the portal-drained viscera and 19% from the liver) (Lapierre
et al., 1999). There is increased protein synthesis in gut epithelium of cattle in response to feed-
ing (Kelly et al., 1995), and this has implications for energy expenditure.
When the degradation of protein is considered, the production of animal protein seems
startlingly inefficient. This apparently wasteful metabolism, however, performs essential
functions. Protein turnover is metabolically costly but is thought to convey flexibility in
protein and amino acid metabolism. Protein breakdown is a particular feature of remodelling,
such as during involution or metamorphosis when entire structures or organs are removed
or replaced. Protein breakdown provides a means to be able to make a rapid change in the
metabolic mass of any protein or group of proteins. Proteins not needed, non-functional, or
damaged may be rapidly removed by activating their catabolism. Proteins may be rapidly
induced, by activation of their synthesis and simultaneous suppression of their catabolism.
Amino acids can be mobilized by degradation of existing proteins, and used for the synthesis
of other proteins or other purposes such as gluconeogenesis. This is an essential function,
which is capable of providing a continuous source of essential and non-essential amino acids,
in a fashion independent of dietary intake.
Continuous turnover of proteins incurs a considerable cost in ATP (Mitch and Goldberg, 1996).
This cost in ATP may be largely associated with the energetic costs of protein synthesis; however,
it is now recognized that at least some elements of the process of protein degradation also require
ATP (Mitch and Goldberg, 1996). Energy economy could be improved by reducing flow through
cyclical metabolic pathways that use ATP, such as protein turnover (Gill et al., 1989).
Because of the contribution of protein catabolism to deposition of marketable muscle tissue and
to its energetic cost, there is interest in different strategies increasing or modulating the rate of
animal growth, especially the relative rates of protein synthesis and catabolism in skeletal muscle.
Whole animal and tissue proteolysis 71

2. MECHANISMS OF DEGRADATION: PROTEOLYTIC SYSTEMS


Intracellular proteolytic systems are extensively characterized in a wide variety of cells, tissues,
and species, including domestic livestock. The intracellular proteolytic systems of muscle that
degrade the myofibrillar proteins are now well characterized, and are the main focus of this
section. Three intracellular proteolytic systems are known to have the potential to contribute
to myofibrillar protein degradation: the lysosomal system, the calcium-dependent proteolytic
system (Tan et al., 1988), and the ATP-ubiquitin–proteasome-dependent proteolytic system
(Attaix et al., 1998). There is a low level of lysosomal proteolytic activity in skeletal muscle
and its overall contribution to catabolism in this tissue is small, except in the case of tissue
injury (Farges et al., 2002).
Cytosolic Ca2+-activated proteases (calpains) and their inhibitors have been extensively
studied in muscle of domestic animals, because of their putative relationship with post-
mortem proteolysis and hence of meat quality (see below). The calpains constitute a large
family comprising ubiquitous, tissue-specific, and atypical calpains (reviewed by Sorimachi
et al., 1997; Kinbara et al., 1998). The calpains are cysteine proteases with a Ca2+ requirement
for activation, in either the millimolar or micromolar concentration range. There are also
atypical calpains, such as p94 (also called calpain 3), a mammalian calpain homologue pre-
dominantly expressed in skeletal muscle, which has been shown to be responsible for a form
of limb-girdle muscular dystrophy. Calpastatin is a specific inhibitor of the calpains and the
isolation of this protein from animal species such as cattle, cloning of its complementary
DNA, and nucleotide sequencing have been completed (Killefer and Koomaraie, 1994). The
contribution of this system to overall muscle protein breakdown in vivo is difficult to esti-
mate; however, in incubated muscles inhibition of this system decreases protein degradation
by less than 10% in most studies. Although some authors suggest that calpains are rate-limiting
for release of filaments from the myofibrillar superstructure, if this were rate-limiting for
myofibrillar proteins to be degraded, inhibition of this system would be expected to block all
myofibrillar proteolysis and this is clearly not so (Attaix et al., 1998, 2001).
Muscle protein catabolism appears primarily mediated by the ATP-dependent ubiquitin–
proteasome system, which is responsible for degrading the bulk of intracellular proteins
including myofibrillar proteins. This conclusion is based on the use of specific proteasome
inhibitors, which are able to block upwards of 60% of total myofibrillar protein catabolism
(Attaix et al., 1998, 2001, 2002). This has been well established in a wide range of animal
models (reviewed by Mitch and Goldberg, 1996; Attaix et al., 2002). The ubiquitination/
deubiquitination system is a complex machine responsible for the specific tagging and proof-
reading of substrates degraded by the proteasome. Polyubiquitination of substrates targets
them for degradation by the proteasome, a multiprotein complex conserved from archaebacteria
to humans. Ubiquitin is an evolutionarily highly conserved 76-amino-acid polypeptide that is
abundant in all eukaryotic cells. The initial step in the ubiquitin pathway is ATP-dependent
and involves the linkage of ubiquitin to a ubiquitin-activating enzyme, or E1, in a high-energy
thioester bond. Ubiquitin is then transferred in a second thioester linkage to a ubiquitin-
conjugating enzyme, which in turn catalyses the transfer of ubiquitin to the substrate protein
in a covalent bond. In some cases, substrate polyubiquitination requires another enzyme, the
ubiquitin ligase (Bodine et al., 2001; Gomes et al., 2001). The ubiquitin ligase can participate
in the hierarchic transfer of ubiquitin into the substrate, or can function as an adaptor to facil-
itate positioning and transfer of ubiquitin from the ubiquitin-conjugating enzyme directly
onto the substrate. The substrates tagged by ubiquitin are then recognized by the proteasome
and degraded into peptides. How this proteolytic pathway degrades muscle proteins, and
72 V. E. Baracos

more particularly contractile proteins, remains largely unknown. Information on the ubiquitin-
conjugating enzymes and ubiquitin ligases that operate in muscle is still scarce. Similarly,
neither the signals that target myofibrillar proteins for breakdown nor the precise substrates
of the pathway have been identified. Finally, the possible relationships between the ubiquitin
proteasome pathway and the lysosomal cathepsins and calpains are not well understood.
The matrix metalloproteinases (MMP) represent a family of enzymes responsible for con-
nective tissue catabolism. Extensive studies in a variety of tissues suggest that the regulation
of MMP activities is complex. MMP are secreted in a latent form as zymogens and activated
sequentially in a cascade initiated by other proteases including plasmin or membrane-type
MMP (MT-MMP). A third level of regulation involves local production of polypeptide tissue
inhibitors of metalloproteinases. The matrix metalloproteinase system involved in intramus-
cular connective tissue degradation has effectively just been described (Balcerzak et al.,
2001). The genetic and physiological modulation of this system is barely characterized.

3. DETERMINATION OF PROTEIN CATABOLISM: WHOLE BODY,


TISSUES, INDIVIDUAL PROTEINS
Measurement of protein catabolism is technically and conceptually difficult. It is not the
intent of this chapter to cover all of the methodological considerations; however, it is impor-
tant to understand thoroughly the inherent limitations of the method used, in the interpretation
of any given set of results. The clearest picture is based on multiple independent approaches
giving the same overall conclusion. A physiologically relevant alteration in rates of protein
catabolism may be small, and a major problem is the size of these changes relative to the error
term of the measurement.

3.1. Degradation by difference: protein synthesis + net protein accretion (loss)

It should be noted at the outset that if the experimental system can be shown to be in a steady
state (i.e. no net protein accretion or loss), rates of protein synthesis and degradation are by
definition identical. In this case protein synthesis is a useful surrogate for measures of protein
degradation. Under non-steady-state conditions, protein degradation may be estimated as the
difference between protein synthesis and net protein gain or loss (i.e. Samuels and Baracos,
1995; Wheeler et al., 2000). Protein gain or loss over time is determined by serial slaughter
of groups of animals on the experimental treatments and protein synthesis is determined in
each group immediately before animals are killed, often using the “flooding dose” approach.
Using this method, the degradative rates of many organs and tissues can be estimated. This
method is applicable to small, inexpensive animals such as chicks or lambs and has been
extensively used in laboratory rodents. Calculated degradation rates include the summed
errors inherent in the directly measured variables, and may be considerable.

3.1.1. Urinary excretion of 3-methylhistidine

Post-translationally modified amino acids released on protein catabolism are not re-incorporated
into proteins and provide an index of catabolism of the proteins of which they are character-
istic. Measurement of urinary 3-methylhistidine (3-MH) excretion is used to estimate
myofibrillar protein breakdown. Similarly, urinary OH− proline reflects the appearance of
this amino acid from the catabolism of connective tissue proteins, mainly collagen (Funaba
et al., 1996). This approach requires quantitative collection of urine and is based on the
Whole animal and tissue proteolysis 73

assumption that no metabolism of 3-MH occurs once it is released from actin and myosin.
This is true in most species, but in sheep and swine a proportion is retained in muscle as a
dipeptide, balenine. In neither of these species does urinary 3-MH yield any data on protein
breakdown. Rathmacher and Nissen (1998) proposed a compartmental model of 3-MH that is
applicable in domestic animals and does not involve the collection of urine. In this approach,
3-MH metabolism in cattle, swine, and sheep was defined from a single bolus infusion of a
stable isotope, 3-[methyl-2H3]-methylhistidine. Following the bolus dose of the stable isotope
tracer, serial blood and urine samples are collected. At least three exponentials were required
to describe the plasma decay curve adequately. A simple three-compartment model described
the plasma kinetics of 3-[methyl-2H3]-MH/3-MH for cattle with one urinary exit from the
plasma compartment. The de novo production of 3-MH as calculated by the compartmental
model in cattle was not different when compared to total urinary 3-MH production. A plasma-
urinary kinetic three-compartment model with two exits was used for sheep with a urinary
exit out of the plasma compartment and a balenine exit out of a tissue compartment. A plasma
three-compartment model was used in swine with an exit out of a tissue compartment. The
kinetic parameters reflect the differences in known physiology of 3-MH metabolism of the
respective species. Steady-state model calculations define masses and fluxes of 3-MH
between three compartments and, importantly, the de novo production of 3-MH.

3.2. Isotopic tracer approaches for whole-body and tissue catabolism in vivo

A primed, constant infusion of an isotopically labelled amino acid such as leucine or phe-
nylalanine may be used to estimate whole protein degradation (i.e. Lapierre et al., 1999; Vann
et al., 2000). Splanchnic tissues drained by the portal and hepatic veins (i.e. Lapierre et al.,
1999) and the hindlimb drained by the femoral vein (i.e. Savary et al., 2001) are organ sites
amenable to determination of protein degradation by techniques based on arterio-venous
differences combined with radioactive or stable isotope tracers.
Wolfe and co-workers have produced a steady stream of methodological advances in this
area (Biolo et al., 1994; Ferrando et al., 1995; Zhang et al., 1996a,b; Patterson et al., 1997).
Zhang et al. (1996a) developed an attractive method to measure the fractional breakdown rate
of muscle protein. This method involves infusing labelled amino acid to reach an isotopic
equilibrium and then observing its decay in the arterial blood and muscle intracellular pool.
The calculation of fractional breakdown rate is based on the rate at which tracer released from
breakdown dilutes the intracellular enrichment using a modified precursor-product equation.
The measured fractional breakdown rates were in agreement with the results from the arterio-
venous balance method. This provides a feasible approach for measurement of muscle protein
catabolism. This method can be combined with the tracer incorporation method to measure
both breakdown and synthesis in the same infusion study.
One limitation of tracer/arterio-venous balance approaches is that the degradative rates of the
individual organs and tissues within the organ system(s) within the studied vascular bed cannot be
descriminated. While limb protein metabolism is often interpreted as equivalent to skeletal muscle
protein metabolism, skin protein synthesis and degradation accounted for approximately 10−15%
of the total leg protein kinetics in different species (Baracos et al., 1991; Biolo et al., 1994).

3.3. Protease gene expression

Protein degradation is clearly regulated, at least to some extent, at the level of gene expres-
sion, and in a wide variety of physiological and pathological states expression of various
74 V. E. Baracos

elements of proteolytic systems varies with measured overall degradative rates. Regulation
of protease gene expression in muscle has been the subject of several elegant studies and
comprehensive review articles by Attaix and co-workers (1997, 1998, 2001, 2002; Larbaud
et al., 2001).
While assessment of gene expression is not a primary measure of degradation, this
approach has applications where it is presently not possible to make direct determinations. A
small amount of data on local protease gene expression is emerging in tissues of the gas-
trointestinal tract, which are suggestive of regulation of proteolysis at this level. For example,
Samuels et al. (1996) measured mRNA levels for components of the lysosomal (cathepsins B
and D), Ca2+-activated (m-calpain), and ubiquitin-dependent (ubiquitin, 14 kDa ubiquitin-
conjugating enzyme E2, and C8 and C9 proteasome subunits) proteolytic pathways, in the
small intestine of rats during food deprivation. mRNA levels for most of these components
increased during fasting, suggesting that a co-ordinated activation of multiple proteolytic
systems contributed to intestinal protein wasting. Adegoke et al. (1999) tested the effects of a
luminal infusion of an amino acid mixture on protease mRNA in jejeunal mucosa of piglets
after overnight food deprivation. Amino acids acutely suppressed mucosal levels of mRNA
encoding ubiquitin, 14 kDa ubiquitin-conjugating enzyme, and the C9 subunit of the prote-
asome by 20–30%, demonstrating the sensitivity of components of the ATP-ubiquitin
proteolytic pathway to acute regulation by nutrients.

3.3.1. In vitro techniques

Aside from poultry (i.e. Baracos et al., 1989), the in vitro incubation techniques developed by
A.L. Goldberg and used widely in muscles of laboratory rodents are not applicable in domes-
tic animal species. A major attribute of this system has been the ability to study the
differential regulation of the lysosomal, Ca2+-dependent and ubiquitin/proteasome-dependent
proteolytic pathways using inhibitors. (i.e. Larbaud et al., 2001).

4. PROTEIN DEGRADATION IN RELATION TO GENETIC MAKE-UP


Protein turnover rates are subject to genetic variation, and differences in protein turnover may
explain part of the inherent differences in efficiency and growth of different animal breeds
(Reeds et al., 1998). For example, Wheeler et al. (2000) evaluated the effect of the callipyge
phenotype in lambs on protein kinetics. These authors studied callipyge and normal lambs at
5, 8, and 11 weeks of age. The synthesis rates of proteins in various tissues were measured
using a primed, continuous infusion of [2H5]phenylalanine. Rates of protein degradation were
estimated by difference between protein synthesis and net protein accretion. Enhanced
muscle growth seems to be maintained in callipyge lambs by reduced protein degradation.
Consistent with this observation, Koohmaraie et al. (1995) reported that the activity of
calpastatin is about 80% higher in the callipyge phenotype.
Unpublished work from our group suggests that different breeds of cattle may exhibit char-
acteristic differences in protein degradation. When animals were fed identical amounts of
metabolizable energy and protein/kg BW·75, Brahman × Angus cross cattle showed a lower
level of urinary 3-MH excretion (1.82 mg/d/kg BW·75) than Charolais (3.06 mg/d/kg BW·75)
(SE = 0.25; P < 0.009). Lower protein degradation, together with a tendency towards lower
metabolic rate, could be responsible for increased protein deposition and consequently
a higher growth rate observed in the Brahman × Angus cattle.
Whole animal and tissue proteolysis 75

5. REGULATION OF DEGRADATION: ENDOCRINE


AND AUTOCRINE CONTROLS
Rates of protein degradation are precisely regulated and there are multiple sites of hormonal
and metabolic controls. In general, circulating hormones have an anabolic or catabolic effect
on a tissue, affecting rates of protein synthesis, degradation, or both. Tissues may be in a steady
state, enter into a catabolic state (wasting) or grow, in response to a concert of hormones and
factors in a given physiological or pathological state. Note that these factors are diverse and
may be hormones, growth factors, substrates, and metabolites. Tissue-specific factors, such
as contractile activity and stretch, also greatly influence muscle protein turnover. The list of
factors affecting the process of protein catabolism is most fully understood for skeletal
muscle (table 1). This symphony of signals act collectively to inform muscle protein catabo-
lism in three information subsets:
1. Contractile activity. Two components of muscular activity, active contraction and passive
stretch, are perceived as anabolic signals. Lack of activity and shortening of muscle
result in activation of the degradation of contractile proteins. This regulation allows for
the maintenance of a muscle mass appropriate to the level of work.
2. Nutritional status/glycemia. Since muscle protein comprises the principal gluconeogenic
precursor, the need for muscle protein mobilization is conveyed through the factors

Table 1
Regulatory factors in muscle protein synthesis and degradation

Synthesis Degradation Overall promotes

Factors related to level of contractile work


Contractile activity ↑ ↓ Protein deposition
Stretch ↑ → Protein deposition
Disuse (inactivity) ↓ ↑ Atrophy
Gonadal steroids ↑ ↓ Protein deposition
Factors related to nutritional status/glycemia
Insulin ↑ ↓ Protein deposition
Insulin-like growth factor I ↑ ↓ Protein deposition
Growth hormone ↑ → Protein deposition
Glucose → ↓ Protein deposition
Ketone bodies → ↓ Protein deposition
Glutamine ↑ ↓ Protein deposition
Branched-chain amino acids ↑ ↓ Protein deposition
Glucagon ↓ → Atrophy
Glucocorticoids → ↑ Atrophy
β-Adrenergic agonists ↑ ↓ Protein deposition
Thyroid hormones (normal) ↑ ↑ Protein deposition
Thyroid (excess) ↑ ↑↑ Atrophy
Factors related to the presence of injury/inflammation
Prostaglandin E2 → ↑ Atrophy
Prostaglandin F2α ↑ → Protein deposition
Interleukin-1β ↓ ↑ Atrophy
Interleukin-6 ↓ ↑ Atrophy
Tumor necrosis factor α ↓ ↑ Atrophy
Interferon γ ↓ ↑ Atrophy
76 V. E. Baracos

regulating glycemia and gluconeogenesis. Since muscle protein comprises a reserve


of amino acids and protein in case of food deprivation, muscle catabolism is sensitive
to indices of food intake, which indicate a state of plenty where growth may occur, or
a state of fasting or nutritional deprivation when net catabolism is required.
3. Stress, infection, and injury. Disease or injury impose additional metabolic demands
when food intake may also be low or zero. Factors arising in the context of the
stress response and immune and inflammatory responses, such as cytokines
and prostaglandins, communicate to muscle the need for extra protein catabolism
(Castaneda, 2002).
All of the above factors are simultaneously at play. When muscle is catabolized, several
metabolic changes (reduced food intake, impaired mobility, and perturbations in the production
or responsiveness of catabolic and anabolic hormones, cytokines, and/or proteolysis-inducing
factors) act in concert. In the context of this complex regulatory system, it is often difficult to
identify the primary and secondary factors, and a complete understanding of their interactions
remains to be developed.

6. PROTEIN DEGRADATION, NUTRITIONAL STATUS, AND GROWTH


6.1. Feeding and diet

The rapid loss of skeletal muscle protein during acute starvation occurs primarily through
increased rates of protein breakdown and activation of the ubiquitin–proteasome-dependent
proteolytic process (Wing et al., 1995). The levels of ubiquitin-conjugated proteins increased
50−250% after food deprivation in various muscles. Like rates of proteolysis, the amount of
ubiquitin–protein conjugates and the fraction of ubiquitin conjugated to proteins increased
progressively during food deprivation and returned to normal within 1 day of refeeding.
Larbaud et al. (1996) showed that euglycemic hyperinsulinemia and hyperaminoacidemia
decrease skeletal muscle ubiquitin mRNA in goats, suggesting insulin and amino acids as
possible mediators of this effect.
Restricted feeding during growth constitutes a stress to energy metabolism, and energy
utilization is curtailed in this circumstance by reduction in growth. In sheep and cattle, protein
turnover is positively related to plane of nutrition (Lobley et al., 1992; Reecy et al., 1996).
The turnover of actomyosin, the major myofibril constituent, is modulated in animals
tested on various planes of nutrition. Lobley et al. (2000) demonstrated that both fractional
degradation rate and fractional synthesis rate were lower in muscles of steers with a low
growth rate (1 kg/d) on restricted feeding in comparison with high growth rate (1.4 kg/d). This
and other studies (i.e. Boisclair et al., 1993) emphasized myofibrillar turnover and muscle;
however, the connective tissue protein catabolism is similarly affected. Lambs with a low
growth rate present less active matrix metalloproteinase-2, suggesting a decrease in collagen
catabolism (Sylvestre et al., 2002). Other studies suggested lower collagen turnover and reduced
deposition of neo-synthesized, immature, and non-cross-linked collagen related with a low
growth rate (Aberle et al., 1981; Crouse et al., 1985; Miller et al., 1987; McCormick, 1994).
The quality of dietary protein has an impact on protein degradation. Branched-chain amino
acids appear to have a specific regulatory effect on protein degradation and decrease the rate
of this process (Ferrando et al., 1995). Diets with protein of inferior quality may increase pro-
tein breakdown in skeletal muscle (Lohrke et al., 2001). These authors studied the activation
of skeletal muscle protein breakdown in pigs fed isoenergetic and isonitrogenous diets based
on soy protein isolate compared with casein.
Whole animal and tissue proteolysis 77

6.2. Role of the somatotrophic axis

The somatotrophic axis plays a key role in the co-ordination of protein metabolism during
postnatal growth (reviewed by Breier, 1999). Acute growth hormone treatment improves the
partitioning of nutrients by increasing protein synthesis and decreasing protein degradation.
Short-term infusion of IGF-1 also reduces whole-body protein breakdown and increases
protein synthesis, and Boyle et al. (1998) have also shown this to be true in fetal sheep during
late gestation. More recently, Vann et al. (2000) suggest that growth hormone increases
protein balance by lowering body protein degradation in fed, growing pigs. Pair-fed, weight-
matched growing swine were treated with porcine growth hormone (150 μg/kg/d) or vehicle
for a week. Growth hormone treatment increased the efficiency with which the diet was used
for growth, but did not alter protein synthesis in skeletal muscles, liver, or jejunum. In the
absence of any changes in protein synthesis at these sites, the results suggest that in the fed
state, growth hormone treatment of growing swine increases protein deposition primarily
through a suppression of protein degradation.

6.3. Stress

Domestic animals are exposed to behavioural, environmental, and infectious stressors. One of
the metabolic hallmarks of these stresses is the catabolic response in skeletal muscle, mainly
reflecting increased protein breakdown, in particular myofibrillar protein breakdown
(reviewed by Hasselgren, 2002). Among different intracellular proteolytic pathways, the
energy-ubiquitin-dependent pathway is particularly important for the regulation of muscle
protein breakdown during acute infection. The gene expression of ubiquitin-conjugating
enzyme E214k, ubiquitin ligase E3α, and several components of the proteasome is up-regulated
and the activity of the proteasome is increased in muscle during infection. An increased
understanding of the molecular regulation of muscle wasting in stress may help in the future
to mitigate or prevent catabolic losses in animal production.

7. PROMOTION OF THE EFFICIENCY AND RATE OF GROWTH


BY MANIPULATION OF PROTEIN DEGRADATION
It is clearly important that proteins must be broken down; however, the physiological range
of protein catabolism rates is quite broad. It is of particular interest that some breeds of ani-
mals and physiological situations are associated with low breakdown rates. The breeds of
animals with lower rates of protein degradation appear to be those breeds adapted to harsh
environments and a low availability or quality of forage. Animal selection and breeding based
upon efficiency of protein deposition may potentially be used to develop this trait.
It has been known for about a decade that β-adrenergic agonists act in part through
suppression of protein degradation (Bardsley et al., 1992; Parr et al., 1992; Mills, 2002).
β-Adrenergic agonists are structurally similar to the catecholamines epinephrine and norepi-
nephrine and bind with high affinity to β-adrenergic receptors in adipose and muscle tissue.
This class of compound includes agents such as clenbuterol, cimaterol, and ractopamine.
Ractopamine was the first β-adrenergic receptor ligand to be cleared for use in pigs in the
USA, about 4 years ago. Ractopamine consistently increases muscle protein accretion in pigs
and while the mechanism responsible for increased protein accretion is not clear, cumulative
evidence points to a direct effect, possibly on both protein synthesis and degradation. One
reason why the role of protein catabolism is not entirely clear is that while β-agonists cause
78 V. E. Baracos

large differences in protein accretion, these are manifest over long periods of time, and the
marginal difference in degradation rates required to result in these changes would be quite
small. β-Agonist treatment does cause large increases in protease inhibitor activity and gene
expression, and these changes are suggestive. For example, cimaterol treatment of Friesian
steers (Parr et al., 1992) caused significant increases in muscle mass (+37%) and calpastatin
specific activity (+76%). Total RNA was unchanged, but there was a 96% overall increase in
calpastatin mRNA in muscle from treated animals.
If protein breakdown could be minimized, this would be an attractive way of promoting
protein deposition as well as a means of lowering the metabolic cost of maintaining any given
protein mass. Thus protein degradation would be an attractive target for growth promotants,
which should properly be called “anticatabolic” rather than “anabolic” factors. As we obtain
more details regarding the proteolytic processes and their regulation, the possibility of iden-
tifying molecular targets for anticatabolic agents seems tangible. A likely site of such targets
would be the ubiquitin–proteasome system and in particular the recently identified class of
muscle-specific ubiquitin ligases (Bodine et al., 2001). Two unique ubiquitin ligases, MuRF1,
a RING finger protein, and MAFbx (Bodine et al., 2001), also called Atrogin-1 (Gomes et al.,
2001), of the SCF family, have been reported to play a role in muscle atrophy. Unlike other
known ubiquitin ligases found in many tissues, these enzymes appear to be expressed mainly
in muscle cells, especially skeletal muscle. Multiple ubiquitin ligases may operate in skeletal
muscle, possibly to connect protein catabolism to different classes of external stimuli. This is
suggested by the reported findings (Bodine et al., 2001) that null mutation of either MAFbx
or MuRF1 in mice led to resistance to denervation-induced muscle atrophy.
The ubiquitin–proteasome system appears to be central in muscle protein degradation,
regardless of the humoral signal for the system’s activation. The intracellular signal trans-
duction from multiple factors converges upon a common proteolytic pathway, of which
ubiquitin ligases are likely to be a critical element. Further studies are required to better
understand the importance of the ubiquitin ligase family, including identifying the physio-
logical substrates for these enzymes in skeletal muscle, elucidating signalling events that
regulate their activity, and analysing the effects of specific inhibition through gene ablation
and/or the design of selective small molecule inhibitors. Ubiquitin ligases may be attractive
molecular targets for manipulation of proteolysis since there are isoforms specific to muscle.
These features may potentially allow for local suppression of muscle catabolism without
affecting the basal proteolytic processes in non-muscle tissues or associated with essential
functions.

8. PROTEIN DEGRADATION AND POST-MORTEM PROTEOLYSIS


Catabolic processes are modified, but not interrupted, at death. While the activity of ATP-
dependent processes (and proteases) would cease, the vast majority of proteolytic enzymes
can continue to express activity. This autolysis is prefaced by the pre-mortem level of proteo-
lytic activities, but thereafter evolves in a manner dissimilar to in vivo events, because of
changes in tissue temperature, pH, and the loss of structural integrity. The identity of all of
the active enzymes and the substrates to which they have access is only partly understood
(Ho et al., 1994).
It has long been believed that proteases play a key role in post-mortem tenderization
of meat (reviewed by Koohmaraie, 1992), and this concept is supported by the observation
that low levels of pre-mortem proteolysis are associated with reduced degradation during
meat maturation. Tissue growth made to be more efficient by reducing proteolytic activity
Whole animal and tissue proteolysis 79

(i.e. β-adrenergic agonist-induced muscle hypertrophy) manifested lowered rates of post-


mortem proteolysis. Koohmaraie et al. (1991) showed that the pattern of post-mortem
proteolysis was altered by β-adrenergic agonists. In β-agonist-treated lambs, post-mortem
storage was not associated with increased myofibril fragmentation index or degradation of
desmin and troponin-T. These results indicate that the ability of the muscle to undergo post-
mortem proteolysis has been dramatically reduced with β-adrenergic agonist feeding.
Similarly, enhanced muscle growth seems to be maintained in callipyge lambs by reduced
protein degradation, and Koohmaraie et al. (1995) suggested a causal relationship between
this effect and increased shear force in meat of calipyge lambs.
The relationship between pre-mortem proteolysis, post-mortem proteolysis, and meat qual-
ity is far from being fully explored. This would be a worthy target for future experimentation,
since an ability to modulate the structural integrity of tissue elements that confer toughness
to meat would have considerable value.

9. FUTURE PERSPECTIVES
A means of reducing protein degradation to its physiological minimum during growth holds
the potential to increase the efficiency of animal production by a large factor. A means of
activating protein degradation in the peri-mortem period holds the potential to increase the
quality of meat. Both of these outcomes could have a large economic impact. Given the
potential impact for animal growth and production, it is perhaps surprising that the research
momentum on proteolysis is not emanating from the agricultural research community.
A pharmaceutical industry strongly motivated to produce therapies for inappropriate degra-
dation, muscle atrophy, and wasting syndromes (i.e. Bodine et al., 2001) is presently making
large financial commitments in this area. Whatever the source, new developments in our
understanding of proteolysis are showing the way towards targets for intervention in these
processes, and animal agriculture may benefit greatly if it is poised to capture the relevant
information.

REFERENCES
Aberle, E.D., Reeves, E.S., Judge, M.D., Hunsley, R.E., Perry, T.W., 1981. Palatability and muscle
characteristics of cattle with controlled weight gain: time on a high energy diet. J. Anim. Sci. 52,
757–763.
Adegoke, O.A., McBurney, M.I., Samuels, S.E., Baracos, V.E., 1999. Luminal amino acids acutely
decrease intestinal mucosal protein synthesis and protease mRNA in piglets. J. Nutr. 129, 1871−1878.
Attaix, D., Aurousseau, E., Combaret, L., Kee, A., Larbaud, D., Rallière, C., Souweine, B., Taillandier, D.,
Tilignac, T., 1998. Ubiquitin-proteasome-dependent proteolysis in skeletal muscle. Reprod. Nutr. Dev.
38, 153−165.
Attaix, D., Combaret, L., Pouch, M.N., Taillandier, D., 2001. Regulation of proteolysis. (Review),
(49 refs). Curr. Opin. Clin. Nutr. Metab. Care 4, 45−49.
Attaix, D., Combaret, L., Pouch, M.N., Taillandier, D., 2002. Cellular control of ubiquitin-
proteasome-dependent proteolysis. J. Anim. Sci. 80, Suppl. 2, E56−E63.
Attaix, D., Taillandier, D., Combaret, L., Rallière, C., Larbaud, D., Aurousseau, E., Tanaka, K., 1997.
Expression of subunits of the 19S complex and of the PA28 activator in rat skeletal muscle (Review).
Mol. Biol. Rep. 24, 95−98.
Balcerzak, D., Querengesser, L., Dixon, W.T., Baracos, V.E., 2001. Coordinate expression of matrix-
degrading proteinases and their activators and inhibitors in bovine skeletal muscle. J. Anim. Sci.
79, 94−107.
Baracos, V.E., Brun-Bellut, J., Marie, M., 1991. Tissue protein-synthesis in lactating and dry goats.
Brit. J. Nutr. 66, 451−465.
80 V. E. Baracos

Baracos, V.E., Langman, M., Mak, A., 1989. An in vitro preparation of the extensor digitorum communis
muscle from the chick (Gallus domesticus) for studies of protein turnover. Comp. Biochem. Physiol.
A Comp. Physiol. 92, 555–563.
Bardsley, R.G., Allcock, S.M., Dawson, J.M., Dumelow, N.W., Higgins, J.A., Lasslett, Y.V., Lockley, A.K.,
Parr, T., Buttery, P.J., 1992. Effect of beta-agonists on expression of calpain and calpastatin activity
in skeletal muscle. Biochimie 74, 267−273.
Biolo, G., Gastaldelli, A., Zhang, X.J., Wolfe, R.R., 1994. Protein synthesis and breakdown in skin and
muscle: a leg model of amino acid kinetics. Amer. J. Physiol. 267, E467−E474.
Bodine, S.C., Latres, E., Baumhueter, S., Lai, V.K., Nunez, L., Clarke, B.A., Poueymirou, W.T., Panaro, F.J.,
Na, E., Dharmarajan, K., Pan, Z.Q., Valenzuela, D.M., DeChiara, T.M., Stitt, T.N., Yancopoulos, G.D.,
Glass, D.J., 2001. Identification of ubiquitin ligases required for skeletal muscle atrophy. Science
294, 1704−1708.
Boisclair, Y.R., Bell, A.W., Dunshea, F.R., Harkins, M., Bauman, D.E., 1993. Evaluation of the arteri-
ovenous difference technique to simultaneously estimate protein synthesis and degradation in the
hindlimb of fed and chronically underfed steers. J. Nutr. 123, 1076−1088.
Boyle, D.W., Denne, S.C., Moorehead, H., Lee, W.H., Bowsher, R.R., Liechty, E.A., 1998. Effect of
rhIGF-I infusion on whole fetal and fetal skeletal muscle protein metabolism in sheep.
Amer. J. Physiol. 275, E1082−E1091.
Breier, B.H., 1999. Regulation of protein and energy metabolism by the somatotropic axis. (Review, 28 refs.)
Domest. Anim. Endocrinol. 17, 209−218.
Castaneda, C., 2002. Muscle wasting and protein metabolism. J. Anim. Sci. 80, Suppl. 2,
E50−E55.
Crouse, J.D., Cross, H.R., Seideman, S.C., 1985. Effects of sex condition, genotype, diet and carcass
electrical stimulation on the collagen content and palatability of two bovine muscles. J. Anim. Sci. 60,
1228−1234.
Farges, M.C., Balcerzak, D., Fisher, B.D., Attaix, D., Bechet, D., Ferrara, M., Baracos, V.E., 2002.
Increased muscle proteolysis after local trauma mainly reflects macrophage-associated lysosomal
proteolysis. Amer. J. Physiol. 282, E326−E335.
Ferrando, A.A., Williams, B.D., Stuart, C.A., Lane, H.W., Wolfe, R.R., 1995. Oral branched-chain amino
acids decrease whole-body proteolysis. J. Parent. Enter. Nutr. 19, 47−54.
Funaba, M., Saito, S., Kagiyama, K., Iriki, T., Abe, M., 1996. Bone growth rather than myofibrillar
protein turnover is strongly affected by nutritional restriction at early weaning of calves. J. Nutr.
126, 898−905.
Gill, M., France, J., Summers, M., McBride, B.W., Milligan, L.P., 1989. Simulation of the energy
costs associated with protein turnover and Na+, K+-transport in growing lambs. J. Nutr. 119,
1287−1299.
Gomes, M.D., Lecker, S.H., Jagoe, R.T., Navon, A., Goldberg, A.L., 2001. Atrogin-1, a muscle-
specific F-box protein highly expressed during muscle atrophy. Proc. Natl. Acad. Sci. USA 98,
14440−14445.
Hasselgren, P.-O., 2002. Stress and muscle wasting J. Anim. Sci. 80, Suppl. 2, E98−E105.
Ho, C.Y., Stromer, M.H., Robson, R.M., 1994. Identification of the 30 kDa polypeptide in post mortem
skeletal muscle as a degradation product of troponin-T. Biochimie 76, 369−375.
Kelly, J.M., Vaage, A.S., Milligan, L.P., McBride, B.W.T., 1995. In vitro ouabain-sensitive respiration and
protein synthesis in rumen epithelial papillae of Hereford steers fed either timothy hay or timothy hay
supplemented with cracked corn once daily. J. Anim. Sci. 73, 3775−3784.
Killefer, J., Koohmaraie, M., 1994. Bovine skeletal muscle calpastatin: cloning, sequence analysis, and
steady-state mRNA expression. J. Anim. Sci. 72, 606−614.
Kinbara, K., Sorimachi, H., Ishiura, S., Suzuki, K., 1998. Skeletal muscle-specific calpain, p94: structure
and physiological function (Review). Biochem. Pharmacol. 56, 415−420.
Koohmaraie, M., 1992. The role of Ca2+-dependent proteases (calpains) in post mortem proteolysis and
meat tenderness (Review). Biochimie 74, 239−245.
Koohmaraie, M., Shackelford, S.D., Muggli-Cockett, N.E., Stone, R.T., 1991. Effect of the beta-adrenergic
agonist L644,969 on muscle growth, endogenous proteinase activities, and postmortem proteolysis in
wether lambs. J. Anim. Sci. 69, 4823−4835.
Koohmaraie, M., Shackelford, S.D., Wheeler, T.L., Lonergan, S.M., Doumit, M.E., 1995. A muscle
hypertrophy condition in lamb (callipyge): characterization of effects on muscle growth and meat
quality traits. J. Anim. Sci. 73, 3596−3607.
Whole animal and tissue proteolysis 81

Lapierre, H., Bernier, J.F., Dubreuil, P., Reynolds, C.K., Farmer, C., Ouellet, D.R., Lobley, G.E., 1999.
The effect of intake on protein metabolism across splanchnic tissues in growing beef steers. Brit. J. Nutr.
81, 457−466.
Larbaud, D., Balage, M., Taillandier, D., Combaret, L., Grizard, J., Attaix, D., 2001. Differential
regulation of the lysosomal, Ca2+-dependent and ubiquitin/proteasome-dependent proteolytic
pathways in fast-twitch and slow-twitch rat muscle following hyperinsulinaemia. Clin. Sci. 101,
551−558.
Larbaud, D., Debras, E., Taillandier, D., Samuels, S.E., Temparis, S., Champredon, C., Grizard, J., Attaix, D.,
1996. Euglycemic hyperinsulinemia and hyperaminoacidemia decrease skeletal muscle ubiquitin
mRNA in goats. Amer. J. Physiol. 271, E505−E512.
Lobley, G.E., Harris, P.M., Skene, P.A., Brown, D., Milne, E., Calder, A.G., Anderson, S.E., Garlick, P.J.,
Nevison, I., Connell, A., 1992. Response in tissue protein-synthesis to submaintenance and supra-
maintenance intake in young growing sheep: comparison of large-dose and continuous-infusion
techniques. Brit. J. Nutr. 68, 373−388.
Lobley, G.E., Sinclair, K.D., Grant, C.M., Miller, L., Mantle, D., Calder, A.G., Warkup, C.C.,
Maltin, C.A., 2000. The effects of breed and level of nutrition on whole-body and muscle
protein metabolism in pure-bred Aberdeen Angus and Charolais beef steers. Brit. J. Nutr. 84,
275−284.
Lohrke, B., Saggau, E., Schadereit, R., Beyer, M., Bellmann, O., Kuhla, S., Hagemeister, H., 2001.
Activation of skeletal muscle protein breakdown following consumption of soyabean protein in pigs.
Brit. J. Nutr. 85, 447−457.
McCormick, R.J., 1994. The flexibility of the collagen compartment of muscle. Meat Sci. 36, 79−91.
Miller, M.F., Cross, H.R., Crouse, J.D., Jenkins, T.G., 1987. Effect of feed energy intake on collagen
characteristics and muscle quality of mature cows. Meat Sci. 21, 287−294.
Mills, S.E., 2002. Biological basis of the ractopamine response. J. Anim. Sci. 80, Suppl. 2, E28−E32.
Mitch, W.E., Goldberg, A.L., 1996. Mechanisms of muscle wasting: the role of the ubiquitin-proteasome
pathway. N. Engl. J. Med. 335, 1897−1905.
Parr, T., Bardsley, R.G., Gilmour, R.S., Buttery, P.J., 1992. Changes in calpain and calpastatin
mRNA induced by beta-adrenergic stimulation of bovine skeletal muscle. Eur. J. Biochem. 208,
333−339.
Patterson, B.W., Zhang, X.J., Chen, Y., Klein, S., Wolfe, R.R., 1997. Measurement of very low stable isotope
enrichments by gas chromatography/mass spectrometry: application to measurement of muscle protein
synthesis. Metab. Clin. Exp. 46, 943−948.
Rathmacher, J.A., Nissen, S.L., 1998. Development and application of a compartmental model of
3-methylhistidine metabolism in humans and domestic animals. Adv. Exp. Med. Biol. 445, 303−324.
Reecy, J.M., Williams, J.E., Kerley, M.S., MacDonald, R.S., Thornton, W.H., Davis, J.L., 1996. The
effect of postruminal amino acid flow on muscle cell proliferation and protein turnover. J. Anim. Sci.
74, 2158−2169.
Reeds, P.J., Burrin, D.G., Davis, T.A., Stoll, B., 1998. Amino acid metabolism and the energetics of
growth. Arch. Anim. Nutr. 51, 187−197.
Samuels, S.E., Baracos, V.E., 1995. Tissue protein-turnover is altered during catch-up growth following
Escherichia coli infection in weaning rats. J. Nutr. 125, 520−530.
Samuels, S.E., Taillandier, D., Aurousseau, E., Cherel, Y., Le Maho, Y., Arnal, M., Attaix, D., 1996.
Gastrointestinal tract protein synthesis and mRNA levels for proteolytic systems in adult fasted rats.
Amer. J. Physiol. 271, E232−E238.
Savary, I.C., Hoskin, S.O., Dennison, N., Lobley, G.E., 2001. Lysine metabolism across the hindquarters
of sheep: effect of intake on transfers from plasma and red blood cells. Brit. J. Nutr. 85, 565−573.
Sorimachi, H., Ishiura, S., Suzuki, K., 1997. Structure and physiological function of calpains (Review).
Biochem. J. 328, 721−732.
Sylvestre, M.N., Balcerzak, D., Feidt, C., Baracos, V.E., Brun-Bellut, J., 2002. Elevated rate of collagen
solubilization and post-mortem degradation in muscles of lambs with high growth rates: possible
relationship with activity of matrix metalloproteinases. J. Anim. Sci. 80, 1871–1878.
Tan, F.C., Goll, D.E., Otsuka, Y., 1988. Some properties of the millimolar Ca2+-dependent proteinase
from bovine cardiac muscle. J. Mol. Cell. Cardiol. 20, 983–997.
Vann, R.C., Nguyen, H.V., Reeds, P.J., Burrin, D.G., Fiorotto, M.L., Steele, N.C., Deaver, D.R., Davis, T.A.,
2000. Somatotropin increases protein balance by lowering body protein degradation in fed, growing
pigs. Amer. J. Physiol. 278, E477−E783.
82 V. E. Baracos

Wheeler, T.L., Savell, J.W., Fiorotto, M.L., 2000. Protein kinetics in callipyge lambs. J. Anim. Sci.
78, 78−87.
Wing, S.S., Haas, A.L., Goldberg, A.L., 1995. Increase in ubiquitin-protein conjugates concomitant with
the increase in proteolysis in rat skeletal muscle during starvation and atrophy denervation. Biochem. J.
307, 639−645.
Zhang, X.J., Chinkes, D.L., Sakurai, Y., Wolfe, R.R., 1996a. An isotopic method for measurement of
muscle protein fractional breakdown rate in vivo. Amer. J. Physiol. 270, E759−E767.
Zhang, X.J., Sakurai, Y., Wolfe, R.R., 1996b. An animal model for measurement of protein metabolism
in the skin. Surgery 119, 326−332.
4 Cytokine regulation of protein accretion
in growing animals

R. W. Johnson and J. Escobar

Department of Animal Sciences, University of Illinois, Urbana, IL 61801, USA

Inflammatory cytokines secreted by activated leukocytes are the critical molecules that enable
the immune system to influence disparate physiological systems that are important for deter-
mining protein accretion in growing animals. The inflammatory cytokines, interleukin-1,
interleukin-6, and tumor necrosis factor α, reduce feed intake, interfere with the somatotropic
axis, reduce skeletal muscle protein synthesis, and enhance skeletal muscle protein degradation.
The purpose of this chapter is to discuss how the immune system regulates feed intake
and arbitrates the balance between skeletal muscle protein synthesis and degradation, so as to
provide a biological explanation for why sick animals do not grow well.

1. INTRODUCTION
Skeletal muscle protein accretion is the net result of both protein synthesis and degradation.
Both events occur constantly in normal skeletal muscle, but the mechanisms regulating
protein synthesis and degradation are distinct and therefore can be influenced independently.
It is now evident that the mechanisms that control skeletal muscle protein synthesis and
degradation are subject to regulation by the immune system (Johnson, 1997). Infectious
pathogens stimulate the immune system, and the immune system in turn actively suppresses
feed intake and skeletal muscle protein accretion. The general notion is that nutrients that
were allocated to support skeletal muscle protein accretion are reassigned to metabolic
processes that support the immune system, which at the time is a higher biological priority
(Klasing, 1988). This places the immune system at the interface of environmental pathogens
and animal growth (Broussard et al., 2001).
Inflammatory cytokines secreted by activated leukocytes are the critical molecules that
enable the immune system to regulate feed intake and nutrient allocation. Because inflam-
matory cytokines reduce voluntary feed intake, and thus the nutrients available to support
protein accretion, this issue will be briefly discussed. Feed intake alone, however, cannot
account for the decreased protein accretion witnessed in sick animals because inflammatory
cytokines also affect protein metabolism by several tissues, including skeletal muscle.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
83 © 2005 Elsevier Limited. All rights reserved.
84 R. W. Johnson and J. Escobar

Therefore, the effects of inflammatory cytokines on skeletal muscle protein synthesis will be
discussed. Some attention will be given to the somatotropic axis because inflammatory
cytokines regulate animals’ capacity to accrete skeletal muscle protein in part by reducing the
amount of growth hormone (GH) and insulin-like growth factor-I (IGF-I) available to skeletal
muscle and by reducing the sensitivity of receptors for GH-releasing hormone (GHRH), GH,
and IGF-I. Because the collective actions of cytokines lead to inhibition of mRNA translation
initiation – an obvious prerequisite for skeletal muscle protein accretion – this issue will be
briefly covered. Finally, the inflammatory cytokines that inhibit protein synthesis concomi-
tantly enhance skeletal muscle protein degradation. Therefore, the effects of infection and
inflammatory cytokines on the ATP-ubiquitin-dependent, calcium-dependent (calpains), and
lysosomal (cathepsins) proteolytic pathways will be discussed as well. The purpose of this
chapter is to discuss how the immune system regulates feed intake and arbitrates the balance
between skeletal muscle protein synthesis and degradation, so as to provide a biological
explanation for why sick animals do not grow well.

2. CYTOKINES ORCHESTRATE ANIMALS’ RESPONSES


TO INFECTIONS
Agricultural animals live surrounded by pathogens and routinely become infected, but
because of a well-developed defense system only occasionally do they show clinical signs of
illness. Still, they are constantly challenged by pathogens and must contend with subclinical
infections on a daily basis. This never-ending mêlée between the animal’s immune system
and pathogens is costly because there is a negative relationship between animal productivity
and the pathogenic environment: as pathogens in the environment increase, animal productivity
decreases (fig. 1). The animal’s immune system “senses” (Blalock, 1984) the pathogenic environ-
ment and biological functions, including feed intake and growth, are adjusted accordingly. To
appreciate how the pathogenic environment impinges upon skeletal muscle protein accretion,
a superficial understanding of the animal’s primary and secondary defenses is obligatory.
The body surfaces are made up of epithelial cells that provide a physical barrier between
the internal milieu and the external pathogen-containing environment. Epithelial cells form the
outer layer of skin and line gastrointestinal, respiratory, and genitourinary tracts. For infection
to occur, primary pathogens must penetrate one of these barriers. Surface epithelia provide
mechanical, chemical, and microbiological protection against infectious pathogens (table 1).
The immune system provides a secondary defense that deals with organisms once they
have entered the body proper. The innate immune response is the first secondary defense to
be mounted. For example, complement and certain acute-phase proteins bind and help destroy
some pathogens, and macrophages trap, engulf, and destroy others. Activated macrophages
also secrete cytokines that cause inflammation, which among other things attracts other
phagocytic cells (e.g. neutrophils and monocytes). These cytokines are collectively called
inflammatory cytokines. They include interleukin-1α/β (IL-1), interleukin-6 (IL-6), and tumor

Fig. 1. As pathogens in the environment


increase, animal productivity decreases.
Cytokine regulation of protein accretion 85

Table 1
Epithelial barriers to infection

Mechanical Epithelial cells joined by tight junctions


Ciliated epithelial cells and mucin that trap and remove pathogens
Chemical Bactericidal enzymes in saliva, sweat, tears, and gut
Low pH in stomach
Antibacterial peptides
Microbiological Normal flora in gastrointestinal tract produce antibacterial substances and
compete against pathogenic microorganisms

necrosis factor α (TNFα). Virus-infected cells produce interferon (IFN) α and γ cytokines that
interfere with viral replication. The interferons increase expression of major histocompability
(MHC) class I molecules on the surface of virally infected cells, thereby flagging these infected
cells for killing by cytotoxic T cells. They also activate natural killer (NK) cells that recognize
and kill virally infected cells. Often this set of responses suffices to eliminate or at least contain
the infection. If the infection cannot be contained it will spread to the lymphatic system where
macrophages and other specialized antigen-presenting cells (e.g. dendritic cells) present the anti-
gen to lymphocytes so they can initiate the second secondary defense – adaptive immunity.
The adaptive immune system is more efficient at eliminating pathogens from the host body as
compared to the innate immune system. When macrophages ingest and degrade pathogenic
microorganisms they process antigen from the pathogen and present it to lymphocytes. This is a
key step for proliferation of the B cells that in turn differentiate into plasma cells that will produce
antigen-specific antibody as well as forming memory B cells that provide long-lasting immunity.
This component of adaptive immunity is called humoral or antibody-mediated immunity.
An antigen also stimulates T cells to form cytotoxic T cells and memory T cells. This compo-
nent of adaptive immunity – cell-mediated immunity – is most effective against intracellular
pathogens such as viruses. Antigen binds a specific receptor on a T cell, stimulating that cell
to differentiate and proliferate. The result is formation of cytotoxic T cells and memory
T cells with receptors appropriate for the subject antigen.
Inflammatory cytokines have critical roles in orchestrating both innate and adaptive
immune responses (table 2). For example, macrophages secrete IL-1β and TNFα to cause
inflammation in order to facilitate movement of other effector cells to the infection site. IL-6
produced by macrophages stimulates hepatocytes to synthesize and secrete acute-phase
proteins, which bind and help remove certain bacteria. IL-6 also stimulates B cells to differ-
entiate into antibody-producing plasma cells. IFNα and γ inhibit virus replication and activate
NK cells which hunt down and kill virally infected cells. And IL-1β stimulates T lymphocytes
to express IL-2 and its receptor – a critical step for T-cell proliferation. A surprising finding
in the late 1970s and early 1980s was that cytokines produced by activated leukocytes affect
disparate physiological systems and orchestrate a systemic response that also helps protect
the host animal. The systemic response initiated by inflammatory cytokines has profound
effects on animal metabolism (table 2). How these cytokines might influence skeletal muscle
protein accretion in young animals is discussed herein.

3. CYTOKINES INHIBIT ANIMAL GROWTH


Animals with infections have reduced appetites, reduced growth rates, and convert feed to
product in an inefficient manner. Indeed, feed intake and growth are usually inversely related
86 R. W. Johnson and J. Escobar

Table 2
Immunological and metabolic effects of cytokines produced by macrophages

Cytokine Major immunological effects Major metabolic effects

Interleukin-1 Inflammation Muscle protein degradation


Activates lymphocytes Reduced muscle protein synthesis
T-cell proliferation Fever
Anorexia
Hypoferremia
Hypozincemia
Hypercupremia
Interleukin-6 Activates lymphocytes Muscle protein degradation
B-cell differentiation Reduced muscle protein synthesis
Antibody production Fever
Acute-phase protein synthesis Acute-phase protein synthesis
Tumor necrosis factor α Inflammation Muscle protein degradation
Reduced muscle protein synthesis
Fever
Anorexia
Lipolysis
Interferon α/γ Activates natural killer cells Not generally considered to have
Inhibits virus replication significant metabolic effects

to the level of interaction between the host immune system and pathogens (fig. 1). This is why
animals kept in poorly sanitized environments that afford a high degree of host–pathogen
interaction eat less and grow more slowly than their counterparts kept in cleaner environ-
ments. This indicates that the immune system “senses” the pathogenic environment and
interacts with the brain and other disparate physiological systems to regulate feed intake and
growth. Initial studies showed that activation of the hypothalamic–pituitary–adrenal (HPA)
axis, fever, and behavioral signs of illness (e.g. hypersomnia) could be induced by injecting
animals with cell-free supernatants collected from activated leukocytes (reviewed by Hart,
1988). The biologically active molecule in the conditioned supernatants was subsequently
determined to be endogenous pyrogen – a protein eventually renamed IL-1. Thus, the immune
system conveys its message to other physiological systems via inflammatory cytokines. It is now
dogma that inflammatory cytokines inhibit animal growth. Administration of inflammatory
stimuli that increase circulating levels of TNFα, IL-1, and IL-6 (e.g. lipopolysaccharide, LPS),
or injection of recombinant TNFα, IL-1, and IL-6, decrease skeletal muscle protein accretion.
These cytokines can reduce skeletal muscle protein accretion in several ways.

3.1. Inflammatory cytokines decrease appetite

A prolonged reduction in feed intake depletes protein and fat reserves. In AIDS and certain
neoplastic diseases, loss of lean body mass is correlated with increased morbidity and mor-
tality (Dewys et al., 1980; Delmore, 1997; Roubenoff, 2000). One thought is that decreased
nutrient intake reduces growth rate, or in adult animals perpetuates loss of skeletal muscle
mass, by limiting the supply of amino acids for protein synthesis. When decreased voluntary
feed intake induced by an inflammatory challenge is accounted for by covariate analysis or
by including a pair-fed control treatment, approximately 20–30% of the decreased growth can
be attributed to a reduction in nutrient intake (Ballinger et al., 2000). The inflammatory
Cytokine regulation of protein accretion 87

cytokines produced by activated mononuclear phagocytic cells, IL-1, IL-6, and TNFα, reduce
appetite and feed intake. The fact that IL-1, IL-6, and TNFα all reduce appetite illustrates an
important characteristic of this group of cytokines – redundancy. Indeed, blocking any one or
two of the three cytokines did not prevent LPS-induced anorexia (Swiergiel and Dunn, 1999).
Only when all three were antagonized simultaneously was LPS-induced anorexia prevented
(Swiergiel and Dunn, 1999). Still, based on dose-response studies, IL-1 appears to be most
potent at reducing appetite. Recombinant IL-1 injected peripherally (i.v., i.p., or s.c.) reduces
animals’ feed intake, similar to what occurs during an acute infection (Plata-Salaman et al.,
1988). The decrease in feed intake is due to a decrease in meal frequency and size (Langhans
et al., 1993). Cytokines can directly change the activity of hypothalamic neurons that medi-
ate feed intake, or affect neurochemicals and neuropeptides that are implicated in the control
of feed intake. For example, peripheral injection of IL-1 caused anorexia and increased the
steady-state level of corticotropin-releasing hormone (CRH) mRNA in the hypothalamus
(Suda et al., 1990). A CRH antagonist administered intracerebroventricularly (ICV) partially
blocked IL-1β-induced anorexia (Uehara et al., 1989). IL-1 also decreases hypothalamic
neuropeptide Y – a potent appetite-stimulating factor (Gayle et al., 1997). And LPS stimulates
the release of α-melanocyte-stimulating hormone, which has been shown to enhance
LPS-induced anorexia (Huang et al., 1999).

3.1.1. Reduced feed intake is an adaptive response to infection

Many explanations for why sick animals reduce feed intake are based on teleology. For exam-
ple, wild animals may expend considerable energy foraging or hunting for feed. Thus, feeding
behavior would enhance heat loss and thwart the beneficial fever response, place weakened,
vulnerable animals in harm’s way of predators, and perhaps facilitate disease transmission
within a group. However, there is tangible evidence that the loss of appetite benefits sick ani-
mals, too. In one study, researchers experimentally infected mice with Listeria monocytogenes
(LD50) and let some consume feed ad libitum, while others were intubated and force-fed to the
level of free-feeding, noninfected controls (Murray and Murray, 1979). Mice allowed to con-
sume feed ad libitum ate 58% of the controls and were much more likely to survive than those
force-fed: nearly 100% of infected, force-fed mice died, whereas only about 50% of infected,
ad libitum-fed mice died. Furthermore, there was a positive relationship between weight loss
and survival for the infected mice with ad libitum access to feed. In some cases,
survival appears to be positively related to anorexia and weight loss, provided it does not
persist too long. In general, the behavioral and metabolic responses to acute infection are ben-
eficial because they inhibit the pathogen and enhance animals’ immunological defenses (fig. 2).
Animals seem to employ their nutritional wisdom and simply eat what they can use. In
other words, feed intake in a growing animal might be determined by its capacity to accrete
protein. When rats were injected with LPS or IL-1 and allowed to self-select between
macronutrients during a 4 h meal period, they decreased total caloric intake by about 50% but
ingested relatively less protein and more carbohydrate; relative fat intake was unchanged
(Aubert et al., 1995). The fact that animals disproportionately reduced protein intake com-
pared to other macronutrients during an inflammatory challenge suggests a shift in metabolic
priorities and nutrient needs. Furthermore, increasing the diet concentration of limiting amino
acids to account for decreased appetite of chicks and pigs under immunological stress is not
effective for increasing whole-body protein accretion (Williams et al., 1997a,b,c; Webel et al.,
1998). Cytokines apparently reduce the animal’s capacity to accrete protein and feed intake
is adjusted accordingly.
88 R. W. Johnson and J. Escobar

Fig. 2. Mononuclear phagocytic cells produce inflammatory cytokines when activated by pathogens.
The cytokine molecules act in the brain to reorganize the animal’s behavioral priorities. The sickness behavior
syndrome that results is an adaptive response that enhances the animal’s immunological defenses and inhibits
proliferation of the pathogen. Thus, sickness behavior enhances disease resistance and promotes recovery
(Johnson, 2002).

3.1.2. How do cytokines affect feed intake regulatory centers?

Cytokines produced in the periphery can interact directly with central feed intake regulatory
centers by entering the circulatory system and moving from the blood into the brain (fig. 3).
Recombinant inflammatory cytokines administered directly into the brain via an indwelling
ICV cannula, for example, induce anorexia, suggesting that cytokines act centrally to reduce
feed intake (Plata-Salaman, 1988). Moreover infusing the IL-1 receptor antagonist ICV in
order to block IL-1 receptors in the brain inhibited anorexia caused by inflammation in the
periphery (Kent et al., 1992; McHugh et al., 1994). Because inflammatory cytokine proteins
are 17–26 kD in size, they are ordinarily too large to diffuse passively from the blood, across
the blood–brain barrier, into the brain. However, pathogens or cytokines might promote
passive movement of cytokine from the blood into the brain by increasing the permeability of
the blood–brain barrier (de Vries et al., 1996). There is also evidence that cytokines are
actively transported from the blood into the brain (fig. 3). For example, Banks and colleagues
injected radiolabeled cytokine (e.g. IL-1, IL-6, and TNFα) intravenously and were able to
recover from the brain a portion of what was injected (Banks et al., 1991, 1994a,b; Gutierrez
et al., 1993). The transport mechanism for each cytokine was saturable and the transport of
radiolabeled cytokine could be competitively blocked by intravenous injection of unlabeled
cytokine.
Peripheral cytokines may also access the brain through circumventricular organs, which
are devoid of blood–brain barrier (fig. 3). Here, peripherally produced cytokines diffuse
into the brain or stimulate glial cells, causing them to produce inflammatory molecules
(e.g. prostaglandins and cytokines), which diffuse into the brain. Based on extensive temporal
and spatial mapping of Fos expression (a marker for neural activity) and IL-1, Konsman et al.
(1999) proposed that cytokines produced in the periphery affect the brain according to the
principles of volume transmission. In this model, IL-1 or other inflammatory mediators in
the periphery induce IL-1 production in the choroid plexus and circumventricular organs. The
cytokine then slowly diffuses into the brain by volume transmission, along the way activating
neurons and neural pathways that result in anorexia (Konsman and Dantzer, 2001). Consistent
with this hypothesis, inflammatory stimuli in the periphery (e.g. LPS and inflammatory
Cytokine regulation of protein accretion 89

Fig. 3. Cytokines produced in the periphery can convey a message to the brain in several ways. Peripheral
cytokines may cross the blood–brain barrier by diffusion or active transport. In addition, peripheral cytokines
may activate the vagus nerve, which in turn induces cells in the brain (e.g. microglia) to produce cytokines.
Finally, peripheral cytokines may stimulate the release of hormones that are able to cross the blood–brain
barrier. Adapted from Johnson (2002).

cytokines) induce de novo synthesis of IL-1, IL-6, and TNFα in the brain of the mouse and
rat (Ban et al., 1992; Laye et al., 1994). For example, inflammatory stimuli in the periphery
induce perivascular microglial cells to express cytokines (van Dam et al., 1992). Moreover,
anorectic rats bearing prostate adenocarcinoma tumor cells had increased IL-1 mRNA in the
cerebellum, cortex, and hypothalamus.
Cytokines in the periphery can also convey a message to the brain via the vagus nerve (fig. 3).
After i.p. LPS challenge, dendritic cells and macrophages that are closely associated with the
abdominal vagus express IL-1 protein (Goehler et al., 1999). IL-1 binding sites are evident in
several regions of the vagus as well (Goehler et al., 1997). When activated by peripheral
cytokines the vagus can activate specific neural pathways that are involved in sickness behavior.
Activation of the vagus also appears to stimulate microglia in the brain to produce cytokines.
If the vagus nerve is severed just below the diaphragm in rats, the expression of cytokines in
the brain and the sickness behavior that normally occurs after intraperitoneal injection of LPS
is inhibited (Laye et al., 1995). Plasma levels of cytokines are elevated in LPS-injected vago-
tomized rats, indicating that the neural signal is needed for the induction of sickness. The
neural signal may be necessary for the induction of cytokines in the brain, or may sensitize
the brain to cytokines produced in the periphery. The neural pathways activated in the brain
by the vagus nerve for rapid immune-to-brain signaling have been recently described in some
detail (Dantzer, 2001a,b). These pathways appear to be responsible for activating the HPA
axis and depressing behavior in response to infection.
Cytokines originating in the periphery act on other peripheral targets as well, which in turn
reduce appetite (fig. 3). For instance, leptin is a 16 kD protein secreted by adipocytes. By
acting in the hypothalamus to reduce appetite and increase energy expenditure, leptin plays
an important role in long-term energy balance. Mice have increased circulating levels of
90 R. W. Johnson and J. Escobar

leptin when the immune system is stimulated with LPS (see Johnson and Finck, 2001). This
effect of LPS is cytokine-dependent because mice with a mutated toll-like receptor 4 gene –
a defect that prevents them from secreting cytokines in response to LPS – do not have
increased leptin when challenged with LPS. However, when mice with the mutation are
injected with recombinant TNFα, circulating leptin increases (Finck et al., 1998). TNFα acts
directly on adipocytes via the p55 TNF receptor to induce expression of leptin (Finck and
Johnson, 2000). It is reasonable to postulate that a cytokine-induced elevation in circulating
leptin is involved in the cytokine-induced anorexia. Accordingly, the anorectic response to
LPS is attenuated in mice lacking the leptin receptor (Faggioni et al., 1997). However, mice
with a mutated leptin gene reduce feed intake similar to mice with a fully functional leptin
gene after LPS injection (Faggioni et al., 1997), so this issue is not fully resolved.

3.2. Inflammatory cytokines decrease protein accretion in growing animals

Inflammatory cytokines are pleiotropic molecules that can either increase or decrease protein
synthesis, depending on the target tissue. In general, protein synthesis is decreased in skeletal
muscle and is increased in liver, lung, and heart. In the liver, for example, inflammatory
cytokines induce a marked increase in acute-phase protein synthesis. The weight of liver and
total liver protein is increased in animals chronically infused with IL-1, TNFα, or the two
cytokines together. However, animals infused with cytokines lose body weight because
muscle protein accretion is reduced due to a decrease in protein synthesis and an increase in
protein degradation. The liver represents roughly 3% of animals’ total body mass whereas
skeletal muscle represents 40–45%. Thus, the net effect of increased circulating cytokines in
a growing animal is a decrease in whole-body protein accretion.
Administration of IL-1 and TNFα – alone or in combination – results in increased urinary
nitrogen excretion and skeletal muscle catabolism accompanied by weight loss in rats (Flores
et al., 1989; Ling et al., 1997). The effects of cytokines on protein kinetics are vastly different
from those induced by fasting or feed restriction, where peripheral proteins are spared and
visceral proteins are degraded. Chronic treatment with either TNFα or IL-1 results in a redis-
tribution of body protein. Rats that were injected twice daily for 7 days with LPS, TNFα, or
IL-1 lost a comparable amount of weight to respective pair-fed animals. However, the LPS
and cytokine-treated animals had accelerated skeletal muscle protein degradation but pre-
served liver protein content, which was not the case for pair-fed animals (Fong et al., 1989).
The decrease in skeletal muscle protein under inflammatory conditions is associated with
decreases in steady-state levels of muscle mRNA for myofibrillar proteins myosin heavy
chain, myosin light chain, actin, and in 18S and 28S subunits of ribosomal RNA. Similarly,
the body weight of transgenic mice that overexpress IL-6 is comparable to that of wild-type
controls at 16 weeks of age. However, the transgenic mice have reduced gastrocnemius
muscle weights and suffer from severe muscle atrophy, which is prevented by treatment with
an antagonistic anti-mouse IL-6 receptor antibody (Tsujinaka et al., 1996). Furthermore,
implantation of a TNFα-secreting tumor in the hind leg muscles of nude mice led within
50 days to profound fat and protein loss (Tracey et al., 1990). In sepsis, protein synthesis
and translational efficiency are reduced in gastrocnemius muscle, and prior treatment with
TNF-binding protein (TNFBP) prevented these effects (Cooney et al., 1999). Prior treatment
with IL-1 receptor antagonist also prevented the inhibitory effects of sepsis on protein synthe-
sis (Vary et al., 1996). The release of one cytokine often initiates a cascade of cytokine synthesis
and release. For example, animals challenged with E. coli or endotoxin and treated with TNFBP
had reduced plasma levels of IL-1 and IL-6 (Roth et al., 1998; Solorzano et al., 1998).
Cytokine regulation of protein accretion 91

Consistent with these results, when pigs were challenged with LPS there was a marked
increase in circulating TNFα and IL-6 that preceded a 3-fold increase in plasma urea
nitrogen (PUN). Because pigs were fasted, the increase in PUN was interpreted to suggest an
increase in skeletal muscle protein degradation (Webel et al., 1997). The Porcine Reproductive
and Respiratory Syndrome Virus (PRRSV) preferentially infects and replicates within
mononuclear phagocytic cells (i.e. macrophages). Mononuclear phagocytic cells infected
by PRRSV produce copious amounts of inflammatory cytokines (van Reeth et al., 1999;
van Reeth and Nauwynck, 2000). Whole-body protein accretion is markedly reduced in nursery
pigs infected with the PRRSV. There is a high negative correlation between protein accretion
and circulating IL-1 and IL-6 (Escobar et al., 2002).
The influence of cytokines on protein synthesis and degradation seems to be dependent on
skeletal muscle fiber type. Slow-twitch, or Type I muscle fibers are designed to work repeti-
tively and generally use oxygen to fuel metabolic processes. Fast-twitch, or Type II muscle
fibers contract at a high rate of speed and work well in the absence of oxygen. Vary and
Kimball (1992) demonstrated that muscles consisting of fast-twitch fibers (gastrocnemius and
psoas) were subject to breakdown during sepsis whereas protein kinetics was unaffected by
sepsis in muscles consisting of slow-twitch fibers (soleus and heart). The specific effects of
inflammatory cytokines on fast-twitch and slow-twitch fibers may be important in domestic
food-producing animals. For example, the longissimus muscle in domestic pigs contains
fewer slow-twitch fibers and more fast-twitch fibers than wild boars of the same age (Essen-
Gustavsson and Lindholm, 1984). Pigs and chickens intended for meat production have been
selected for maximal lean growth rate and increased breast-meat yield, respectively. Because
pigs selected for maximal lean growth rate have a greater proportion of muscles containing
fast-twitch vs slow-twitch muscles (Rahelic and Puac, 1981), and chickens selected for
maximal breast-meat yield likewise have higher levels of fast-twitch fibers, the effects of
cytokines on muscle tissue growth are potentially more deleterious in leaner, more modern
genotypes.
It appears that a portion of the amino acids released by skeletal muscle as a result of
protein degradation are taken up by leukocytes to support cell proliferation and by the liver
to support acute-phase protein synthesis (fig. 4). For example, the inflammatory cytokines,
IL-1, IL-6, and TNFα, increase the rate of hepatic amino acid uptake (Argiles et al., 1989;
Argiles and Lopez-Soriano, 1990) and protein synthesis (Klasing and Austic, 1984; Geiger
et al., 1988; Ballmer et al., 1991). Reeds et al. (1994) proposed that a significant portion of
nitrogen excreted during an inflammatory response was the result of excessive demands for
the aromatic amino acids, phenylalanine, tyrosine, and tryptophan. Their conclusions were
based on a comparison of the amino acid profiles of the major acute-phase proteins produced
by humans and the amino acid profile of mixed muscle protein. Analysis of the acute-phase
proteins indicated that four of the six proteins contained high levels of phenylalanine, five
of the proteins were rich in tryptophan, and three contained high levels of tyrosine. By cal-
culating the quantity of amino acids incorporated into a typical acute-phase protein mixture
(850 mg/kg BW), they calculated that 1980 mg of muscle protein per kg body weight would
need to be liberated to supply an adequate quantity of phenylalanine for the increased hepatic
protein synthesis. The amino acids that are released in excess of the need for acute-phase pro-
tein production (1980–850 mg) are catabolized because they cannot be used for protein
resynthesis due to the phenylalanine limitation, with the end result being an excessive excre-
tion of nitrogen. Assuming that animals have a similar pattern and quantity of acute-phase
proteins, it is apparent that an infectious insult could result in a substantial amount of skeletal
muscle protein degradation and nitrogen excretion. For example, for a 100 kg pig there would
92 R. W. Johnson and J. Escobar

Fig. 4. Inflammatory cytokines inhibit skeletal muscle protein synthesis and enhance its degradation.
A portion of the freed amino acids (AA; e.g. glutamine) is taken up by leukocytes to support cell proliferation
and by the liver to support acute-phase protein (APP) synthesis and other metabolic processes.

be roughly 200 g of protein broken down to supply amino acids for acute-phase protein syn-
thesis, and approximately 13 g nitrogen would be excreted.
Glutamine is another example of how amino acids are repartitioned by cytokines (fig. 4).
Skeletal muscle is the major repository of glutamine. During infection, there is a 2-fold
increase in glutamine release from skeletal muscle. Despite a significant increase in endoge-
nous glutamine biosynthesis in skeletal muscle, intracellular glutamine becomes depleted.
At the same time there is an 8- to 10-fold increase in hepatic glutamine uptake where it can
be used for (1) biosynthesis of nonessential amino acids; (2) gluconeogensesis; (3) energy;
and (4) biosynthesis of urea, which is ultimately excreted.

4. CYTOKINES AND MUSCLE PROTEIN SYNTHESIS


AND DEGRADATION
When protein degradation remains constant, decreased protein accretion can occur when sub-
strates necessary for protein synthesis are limiting or when signals that promote protein
synthesis are thwarted. When protein synthesis remains constant, decreased protein accretion
can occur when signals that promote protein degradation are enhanced. Protein synthesis and
degradation can be influenced independently, so protein accretion is most profoundly affected
when there is a decrease in muscle protein synthesis and a concomitant increase in muscle
protein degradation. If the amount of degradation exceeds that of synthesis, muscle wasting
occurs. Inflammatory cytokines are uniquely qualified to manipulate protein accretion
because they have the ability to simultaneously influence both protein synthesis and degra-
dation. Thus, the immune system, by production of inflammatory cytokines, is able to adjust
animal growth according to the level of immunological challenge.

4.1. Inflammatory cytokines inhibit skeletal muscle protein synthesis

4.1.1. Inflammatory cytokines inhibit the GH–IGF-I axis

One way cytokines inhibit skeletal muscle protein synthesis is by adversely affecting the
somatotropic axis (fig. 5). Growth hormone induces IGF-I secretion, a potent growth factor
that is responsible for a wide range of anabolic processes. Insulin-like growth factor-I
increases skeletal muscle mass by binding the type I IGF-I receptor and initiating a cascade
of intracellular signaling events that ultimately initiate protein synthesis. For a detailed
Cytokine regulation of protein accretion 93

Fig. 5. Inflammatory cytokines can inhibit skeletal muscle protein synthesis by interfering with the secretion
of growth hormone (GH) and insulin-like growth factor (IGF)-I. Cytokines also lead to GH and IGF-I receptor
resistance.

description of IGF-I receptor signaling, the reader is referred to recent reviews (LeRoith,
2000; Broussard et al., 2001; Nakae et al., 2001). In short, the IGF-I receptor belongs to the
family of tyrosine kinase receptors. It is composed of two ligand-binding extracellular α subunits
and two transmembrane-spanning β subunits that have tyrosine kinase activity. Following
binding of the IGF-I to the extracellular receptor α subunit, the receptor dimerizes, forms a
heterotetramer (βααβ), and the tyrosine residues in the kinase domain of the β chains are
autophosphorylated. The tyrosine phosphorylated IGF-I receptor causes tyrosine phosphory-
lation of insulin receptor substrate (IRS)-1 and IRS-2 – docking molecules that can recruit
and bind the p85 regulatory subunit of phosphatidylinositol 3′-kinase (PI 3-kinase). The IRS
docking molecules can lead to a sustained activation of PI 3-kinase, which is key in connect-
ing several intracellular pathways that promote cell survival, differentiation, and protein
synthesis.
In general, inflammatory stimuli reduce IGF-I levels and reduce sensitivity of receptors for
GHRH, GH, and IGF-I. Receptors for IL-1 are present in the anterior pituitary and are localized
exclusively to somatotrophs – cells that produce GH (French et al., 1996). Stimulation of the
immune system with LPS causes species-specific changes in circulating GH levels. In humans
and sheep, GH is increased, but in other animals including cattle, chickens, and rats it is
decreased. The effects of LPS and cytokines on GH are discussed elsewhere and the general
conclusion is that the effects are variable and not consistent among species (Broussard et al.,
2001). Of seemingly greater importance is IGF-I, which is consistently depressed in immuno-
logically challenged animals. Pigs injected with LPS or infected with Salmonella typhimurium
showed a marked decrease in serum IGF-I levels, but little or no change in circulating GH
(Balaji et al., 2000; Wright et al., 2000). An uncoupling of GH and IGF-I secretion has been
reported in a number of species and is due to the impaired ability of GH to induce hepatic
IGF-I synthesis. TNFα and IL-1 decrease hepatic GH receptors and may inhibit post-receptor
signaling events necessary for IGF-I synthesis and release. Interleukin-6 also profoundly
affects IGF-I. Transgenic mice that overexpressed IL-6 and wild-type mice injected with
recombinant IL-6 had decreased IGF-I levels and stunted growth (De Benedetti et al., 1997).
94 R. W. Johnson and J. Escobar

Mice injected with IL-6 had decreased IGF-I levels even when feed intake was not depressed,
indicating that the decrease in IGF-I in sick animals is not necessarily due to reduced feed
intake.
Inflammatory cytokines can also act directly on skeletal muscle and induce IGF-I receptor
resistance. Bona fide receptors for IL-1, IL-6, TNFα, and IFNγ are present in skeletal muscle
(Zhang et al., 2000; Alvarez et al., 2002a) and a model for how cytokines might interfere with
IGF-I receptor signaling has been proposed (Broussard et al., 2001). TNFα completely
inhibits the IGF-I-induced increase in protein synthesis of human myoblasts (Frost et al.,
1997). Apparently, TNFα impairs the ability of IGF-I receptors to exert their biological effect
when IGF-I ligand binds. TNFα reduces IGF-I-induced tyrosine phosphorylation of the IGF-I
receptor and IRS (Venters et al., 1999). This decreases PI 3-kinase activity and thus inhibits
the ability of IGF-I to promote protein synthesis. The IGF-I receptor resistance partially
explains why administration of recombinant IGF-I to rats with colitis failed to restore linear
growth to that of control rats (Ballinger et al., 2000). However, most circulating IGF-I is
bound to IGF-binding protein 3 (IGFBP3). A recent study showed that administration of
IGF-I/IGFBP3 binary complex enhanced protein synthesis in septic rats, suggesting that the
decreased responsiveness of muscle to exogenous IGF-I may be due to both an induction
of IGF-I receptor resistance and decreased circulating levels of important IGF-I-binding
proteins (Svanberg et al., 2000).

4.1.2. Inflammatory cytokines inhibit the initiation of mRNA translation

The synthesis of new protein – an obvious prerequisite to protein accretion – begins with the
initiation of mRNA translation (i.e. translation initiation; fig. 6). In eukaryotic cells, transla-
tion initiation involves more than a dozen proteins referred to as eukaryotic initiation factors
(eIF). For a complete up-to-date description of translation initiation and how the process is

Fig. 6. A truncated schematic diagram of the translation initiation process. See text for details on how
cytokines inhibit the initiation of protein synthesis (eukaryotic initiation factor, eIF; binding protein, BP).
Cytokine regulation of protein accretion 95

influenced by nutrients and hormones, the reader is referred to excellent reviews by Kimball
(2002) and Shah et al. (2000). In short, translation initiation consists of four major steps
(Cooney et al., 1997): (1) dissociation of the 80S ribosomal complex into the 40S and 60S
ribosomal subunits; (2) binding of met-tRNAi (methyonil-tRNA initiator) to the 40S ribosomal
subunit to form the 43S pre-initiation complex; (3) binding of mRNA to the 43S pre-initiation
complex: and (4) association of the 60S ribosomal subunit to form the active ribosome
(fig. 6). Several recent reports indicate that steps 2 and 3 are affected most during infection,
so these are discussed below.
4.1.2.1. Formation of the 43S pre-initiation complex The primary function of eIF2 is to
bind met-tRNAi to the 40S ribosomal subunit. Thus, eIF2 can regulate translation initiation
in two ways. First, a decrease in the amount of intracellular eIF2 or its activity will reduce
translation initiation. However, to our best knowledge a reduction in eIF2 during infection
when skeletal muscle protein synthesis is reduced has not been reported. Second, eIF2 activity
is controlled by eIF2B, whose function is to exchange GDP for GTP in eIF2 (i.e., eIF2 . GDP
+ eIF2B + GTP → eIF2 . GTP + eIF2B + GDP). Only the eIF2 . GTP complex is able to bind
mRNA. Therefore, a reduction in eIF2 . GTP complex will decrease the translation initiation
process. Indeed, a decrease in eIF2B was accompanied by a decrease in protein synthesis and
translational efficiency in the gastrocnemius muscle of septic rats (Vary et al., 1994; Voisin
et al., 1996b). In the same study, treatment of septic rats with IL-1 receptor antagonist
returned eIF2B and protein synthesis and translational efficiency in the gastrocnemius muscle
to control levels. In a separate but similar study, injection of LPS did not affect eIF2B levels
in the gastrocnemius muscle, but eIF2B activity was decreased as was the rate of protein syn-
thesis and translational efficiency (Lang et al., 2000). Thus, a change in either the amount or
activity of eIF2B is sufficient to influence skeletal muscle protein synthesis. Similar results
were reported when rats were infused with TNFα (Lang et al., 2002). Administration of the
TNFα antagonist, TNFBP, restored the translational efficiency and protein synthesis in septic
rats to control levels (Cooney et al., 1999). Furthermore, TNFBP prevented the decrease in
eIF2B in gastrocnemius muscle of septic rats.

4.1.2.2. Binding of mRNA to the 43S pre-initiation complex The primary function of
eIF4E is to bind mRNA to form the eIF4E . mRNA complex, which binds to eIF4G and
eIF4A to form the active eIF4F (Gingras et al., 1999). The 43S pre-initiation complex directly
binds to eIF4F. The eIF4E is bound to its repressors 4E-binding protein (4E-BP1), 4E-BP2,
and 4E-BP3, where 4E-BP1 is the predominant form in skeletal muscle (Vary and Kimball,
2000). Kinases phosphorylate 4E-BP1 and free eIF4E. Thus, phosphorylation of 4E-BP1 and
the level and activity of eIF4E are important factors that can influence protein synthesis.
Interference of this signaling cascade by cytokines might partially explain the decreased pro-
tein synthesis in sick animals. However, it is not clear if the phosphorylation state of 4E-BP1
and the level and activity of eIF4E are affected during infection (Vary and Kimball, 2000;
Vary et al., 2001). For example, on the one hand, IGF-I stimulates protein synthesis in skele-
tal muscle by inducing an intracellular signaling cascade (reviewed by Broussard et al., 2001)
that ultimately results in the phosphorylation of 4E-BP1 and release of eIF4E. In septic rats,
phosphorylation of 4E-BP1 is markedly reduced in the gastrocnemius muscle so there is
more 4E-BP1 associated with eIF4E. Accordingly, both protein synthesis and translational
efficiency are reduced (Svanberg et al., 2000). Administration of IGF-I/IGFBP3 binary
complex to septic rats restored translational efficiency without affecting the phosphorylation
state of 4E-BP1 or the association of eIF4E with 4E-BP1 (Svanberg et al., 2000). Thus,
how IGF-I/IGFBP3 increases protein synthesis during sepsis is not yet known. On the
96 R. W. Johnson and J. Escobar

other hand, insulin induces hyperphosphorylation of 4E-BP1 during sepsis, which causes the
predicted dissociation of eIF4E . 4E-BP1 complex (Vary et al., 2001). However, formation of
eIF4E . eIF4G complex, which is necessary for protein synthesis, was drastically reduced in
septic rats despite the availability of eIF4E (Vary et al., 2001). The reduced binding of eIF4E
to eIF4G may be important for inhibition of protein synthesis during infection.
In summary, anti-cytokine treatment during sepsis appears to prevent the decrease in eIF2B
and enhance protein synthesis and translational efficiency. In contrast, administration of
IGF-I/IGFBP3 binary complex enhances protein synthesis and translational efficiency during
infection even though the amount of eIF2B is reduced. Thus, regulation of eIF2B during
sepsis might be independent of IGF-I but under direct cytokine control.

4.2. Inflammatory cytokines enhance skeletal muscle protein degradation

4.2.1. ATP-ubiquitin-dependent proteolytic pathway

The majority of proteolysis in skeletal muscle, including short-lived and long-lived myofib-
rillar proteins, is ATP-dependent and involves the cofactor ubiquitin (Ub) and the 26S
proteasome complex (Solomon and Goldberg, 1996; Hasselgren and Fischer, 2001; fig. 7).
The functionality, structural features, regulation, and modes of action of the 26S proteasome
have been extensively reviewed (Gorbea et al., 1999; Tanahashi et al., 1999; Voges et al., 1999).
In short, an enzyme (E1) activates Ub in the presence of ATP to produce Ub-adenylate, which
subsequently is transferred to one of several Ub-carrier enzymes (E2). Ubiquitin is attached
to the ε-amino group in a lysine residue of a target protein directly by Ub-carrier enzymes or
through the action of Ub-ligases (E3). Attachment of more than one Ub (poly-Ub) is generally
observed in E3-dependent reactions. Poly-Ub is the preferred signal for protein degradation
within the 26S proteasome. Protein degradation by the 26S proteasome results in many small

Fig. 7. A schematic diagram of the ATP–ubiquitin-dependent proteolytic pathway. See text for details on
how cytokines influence this proteolytic pathway (ubiquitin, Ub; enzyme, E).
Cytokine regulation of protein accretion 97

peptide chains that can be further degraded to free amino acids. During or after proteasome
degradation, Ub is released from the remnant of the target protein and is recycled for later use.
Not all Ub-tagged proteins are degraded because Ub can be freed by the action of specific
isopeptidases. The specificity of the ubiquitination process is thought to be primarily due to
the many different enzyme isoforms.
Because infection is associated with proteolysis, the role of the ATP-Ub-dependent pathway
in skeletal muscle has been investigated. The increase in skeletal muscle protein degradation
in septic or tumor-bearing rats is paralleled by an increase in ATP-Ub-dependent proteolysis
(Temparis et al., 1994; Tiao et al., 1994). For example, protein degradation was increased in
extensor digitorum longus muscle isolated from septic rats compare to controls (Tawa et al.,
1997; Hobler et al., 1998). Protein degradation was inhibited when the muscle was incubated
in the presence of specific proteasome inhibitors, MG101 (N-acetyl-Leu-Leu-norleucinal;
Hobler et al., 1998), lactacystin (Hobler et al., 1998), or MG132 (Tawa et al., 1997). Proteolysis
increases progressively in denervated muscle. Proteolysis in denervated soleus muscle was
inhibited 68% by the proteasome inhibitor, MG132 (Tawa et al., 1997). These results strongly
indicate the crucial role of the ATP-Ub-dependent pathway in skeletal muscle proteolysis.
As indicated earlier, skeletal muscle expresses receptors for IL-1, IL-6, TNFα, and IFNγ
(Zhang et al., 2000; Alvarez et al., 2002a). Inflammatory cytokines appear to play a key role
in activating the ATP-Ub-dependent proteolytic pathway (Llovera et al., 1998a); however, the
specific role of each cytokine is not clear. Tumor necrosis factor α but not IL-1 increased Ub
mRNA in gastrocnemius muscle of rats 3 h after bolus injection (Garcia-Martinez et al.,
1995). However, when examined 6 h after bolus injection, Ub mRNA in gastrocnemius
muscle was increased by TNFα, IL-1, and IFNγ, but not IL-6 (Llovera et al., 1998a). Tumor
necrosis factor α has been studied for its role in protein degradation more than any other
cytokine, having been initially referred to as cachectin. For example, acute treatment of rats
with TNFα increased protein ubiquitination (Garcia-Martinez et al., 1993) and enhanced
the degradation of both total and myofibrillar proteins (Zamir et al., 1992; Fischer et al.,
2001). And incubation of C2C12 murine myotubes with TNFα reduced total cellular protein
content (Li et al., 1998; Li and Reid, 2000; Alvarez et al., 2002b). The protein degradation
induced by TNFα is partially due to activation of the ATP-Ub-dependent proteolytic pathway.
During tumor growth, muscle wasting is associated with the activation of the ATP-Ub-dependent
proteolytic pathway, which is mediated via cytokines. Immunoneutralization of TNFα with
a polyclonal anti-TNF antibody blocked the increase in steady-state levels of Ub mRNA in
gastrocnemius muscle of tumor-bearing rats (Llovera et al., 1996). Xanthine derivatives
such as pentoxifylline have been used to inhibit TNFα production. A new xanthine derivative,
torbafylline, decreased plasma concentration of TNFα in rats injected with LPS and in
Yoshida sarcoma-bearing rats (Combaret et al., 2002). The decrease in circulating TNFα
was paralleled by decreases in muscle Ub mRNA, Ub-conjugated myofibrillar proteins,
proteasome mRNA (C2 subunit of 20S), and proteasome-dependent proteolysis (Combaret
et al., 2002).
The soluble TNF receptor can serve as a decoy for biologically active TNFα. Binding of
TNFα to the soluble receptor prevents the cytokine from binding cell membrane-bound
receptors, thus effectively eliminating the cytokine’s bioactivity. Implantation of Lewis
lung carcinoma cells caused a decrease in gastrocnemius muscle protein accumulation in
both wild-type mice and transgenic mice that overexpressed soluble TNF receptor (Llovera
et al., 1998b). However, the reduction in protein accumulation was substantially greater
in wild-type mice compared to transgenic mice. In both cases, when protein accumulation
was decreased, Ub mRNA in gastrocnemius muscle was increased. Tumor necrosis factor
98 R. W. Johnson and J. Escobar

receptor type I knockout mice implanted with Lewis lung carcinoma cells had reduced gas-
trocnemius muscle wasting compared to wild-type controls (Llovera et al., 1998c).
Further evidence that IL-6 does not activate the ATP-Ub-dependent proteolytic pathway
comes from studies with IL-6 knockout mice. Sepsis induced by cecal ligation and puncture
induced a similar increase in total and myofibrillar protein degradation in both wild-type and
IL-6 knockout mice, suggesting that IL-6 was not necessary for protein degradation (Williams
et al., 1998). Interestingly, in wild-type mice sepsis was accompanied by a marked increase
in muscle Ub mRNA, whereas Ub mRNA was only moderately increased in muscle of IL-6
knockout mice. In contrast, administration of IL-6 to rats reportedly increased total and
myofibrillar protein degradation (Goodman, 1994). The expression of Ub mRNA, however,
is not always indicative of the proteolytic activity of the ATP-Ub-dependent pathway
(Hasselgren, 2000). In contrast, transgenic mice that overexpressed IL-6 had severe muscle
atrophy and increased levels of Ub mRNA in gastrocnemius muscle. The increase in Ub
mRNA levels was reduced when mice were treated with an anti-IL-6 antibody, which also
restored muscle mass (Tsujinaka et al., 1996). These contradictory results exemplify the
potential difficulties of interpreting results from genetically altered animals. For example,
TNFα, which induces proteolysis, was reportedly 3-fold higher in septic IL-6 knockout mice
compared to wild-type controls (Fattori et al., 1994). Manipulating a single cytokine often
disrupts the normal cytokine cascade. In addition, the ability of the ATP-Ub-dependent pathway
to degrade protein may depend on the activity of other proteolytic systems (e.g. Ca2+-dependent).
Thus, it might be important to examine the effects of multiple cytokines and multiple proteolytic
systems simultaneously.

4.2.2. Calcium (Ca2+)-dependent proteolytic pathway

The Ca2+-dependent proteolytic pathway involves the proteases μ-calpain, m-calpain, and
calpain 3 (muscle-specific calpain p94). μ-Calpain and m-calpain are so named because they
are activated by micro- and millimolar concentrations of Ca2+, respectively. A concentration
of 1–20 μM of Ca2+ is required to activate μ-calpain, which exceeds the normal physiological
intracellular concentration. Therefore, the calpains are generally inactive, so their role in normal
cellular function is largely unknown (Stracher, 1999). Mitochondria and the sarcoplasmic
reticulum release large amounts of Ca2+ postmortem and activate the calpains. The calpains
are most noted for the proteolytic changes in postmortem muscle, which are important for
improving tenderness (Koohmaraie, 1992).
The overall contribution of calpains to skeletal muscle protein degradation during infection
when protein accretion is decreased is relatively small compared to ATP-Ub-dependent
proteolysis. The calpains, however, are important for degradation of certain muscle proteins
including the sarcomeric proteins (Huang and Forsberg, 1998). Even though the total muscle
protein degraded by calpains is relatively small, it is important because degradation of
sarcomeric proteins may facilitate degradation of myofilaments. In other words, the activity
of the Ca2+-dependent pathway may facilitate the activity of the ATP-Ub-dependent pathway.
It has been proposed that the Ca2+-dependent release of myofilaments from the sarcomere
of myofibrils is the rate-limiting event for ATP-Ub-dependent proteasome degradation of
the myofilaments (Hasselgren et al., 2002). Indeed, it appears that the activity of the Ca2+-
dependent proteolytic pathway is increased during infection and that blockade of this
pathway markedly reduces total skeletal muscle protein degradation.
Skeletal muscle m-calpain was increased in septic rats experiencing accelerated muscle
protein degradation (Voisin et al., 1996a). In a recent study, steady-state mRNA levels
Cytokine regulation of protein accretion 99

of μ-calpain, m-calpain, and calpain 3 were increased and the Z-disks disrupted in extensor
digitorum longus muscle of septic rats (Williams et al., 1999). The addition of dantrolene, an
inhibitor of the release of Ca2+ from intracellular stores to the cytoplasm, significantly
reduced myofilament release from the sarcomere in septic rats, suggesting involvement of a
Ca2+-dependent proteolytic pathway – probably calpain (Williams et al., 1999). In support of
this, dantrolene prevented the sepsis-induced increase in muscle Ca2+ levels, mRNA levels for
m-calpain, μ-calpain, and calpain 3, the increase in release of myofilaments, and total protein
degradation in extensor digitorum longus muscle (Fischer et al., 2001). Interestingly, dantro-
lene reduced serum TNFα as well, suggesting an important but yet to be defined relationship
between cytokines, Ca2+-dependent proteolysis, and muscle wasting.

4.2.3. Lysosomal proteolytic pathway

Skeletal muscle contains relatively few lysosomes, but the involvement of the lysosomal
proteolytic pathway in skeletal muscle protein degradation during disease has been investi-
gated nonetheless. The main lysosomal proteases are cathepsins B, H, L, and D, which play
a major role in the degradation of long-lived, soluble, and integral membrane proteins.
Cathepsins, however, are unable to degrade myofibrillar proteins (Furuno et al., 1990), so the
lysosomal proteolytic pathway’s contribution to disease-induced muscle wasting is consid-
ered to be relatively small. Inconsistent findings have been reported for the activity of the
lysosomal proteolytic pathway in skeletal muscle of wasting animals (Llovera et al., 1994,
1995; Temparis et al., 1994; Baracos et al., 1995; Voisin et al., 1996a). For example, muscle
levels of cathepsin B were increased in septic rats compared to pair-fed controls (Voisin et al.,
1996a), whereas cathepsin B and B + L activities and cathepsin B mRNA were unchanged in
wasting tumor-bearing rats (Temparis et al., 1994). Transgenic mice that overexpressed IL-6
suffered from muscle atrophy and had increased steady-state mRNA levels of cathepsins B
and L as well as ubiquitin. Injecting IL-6 transgenic mice with an IL-6 receptor-blocking anti-
body decreased cathepsin (B and L) and Ub mRNA. In injured muscle, lysosomal proteolysis
appears to be mediated by enzymes produced primarily by infiltrating macrophages (Farges
et al., 2002). Enzymes from macrophages may not be a prerequisite, however. Incubation of
C2C12 myotubes with IL-6 reduced the half-life of long-lived proteins and increased cathep-
sin B + L activity (Ebisui et al., 1995). It is possible that the lysosomal pathway participates
in muscle wasting indirectly by controlling degradation of intracellular regulatory proteins.

5. FUTURE PERSPECTIVES
Most of what is known about the in vivo effects of cytokines on skeletal muscle protein
synthesis and degradation has been derived from adult animal models of sepsis or cancer –
diseases that are often characterized by severe muscle wasting. These are important models
for human medicine and have served to amplify the host’s responses to acute immune system
activation, which is useful for elucidating the underlying mechanisms involved. It must be
noted, however, that these diseases are only marginally significant to growing domestic food-
producing animals. In general, the most significant diseases in growing agricultural animals
do not cause muscle wasting, but instead cause a significant and chronic decrease in skeletal
muscle protein accretion. Thus, despite the presence of infectious disease, farm animals often
continue to grow. In any case, it is reasonable to postulate that the same machinery that results
in muscle wasting in sepsis, for example, is involved in decreasing protein accretion in
infected slow-growing animals. Therefore, results of studies on muscle wasting might be
100 R. W. Johnson and J. Escobar

germane to increasing protein synthesis and reducing protein degradation in slow-growing


animals. However, it is not clear if the “markers” of protein synthesis and degradation evident
in animals undergoing severe muscle wasting are detectable in farm animals that are merely
accreting skeletal muscle protein at a less than maximal rate due to infection.
To defend the host animal from infectious pathogens, the immune system and liver use
some of the amino acids made available by the effects of cytokines on skeletal muscle. Thus,
in growing animals we believe that the repartitioning of nutrients is a constructive adaptation
that enables the animal to contend against the pathogen and continue to accrete protein, albeit
at a decreased rate. It should be possible to develop novel nutrition programs that better match
the animal’s metabolic state(s) during an infection, so that protein accretion is maintained
while the needs of the defense systems are met. However, the nutrient requirements for
agriculture animals are, for the most part, based on experiments conducted in laboratory
situations where exposure to infectious pathogens and other stresses is minimized. Therefore,
the estimated nutrient requirements for animals have been established to maximize production
of healthy animals. If the nutritional requirements of slow-growing infected animals can be
precisely defined, it will be possible to formulate cost-effective diets that maximize protein
accretion under the given circumstance.
Of course the goal should be to minimize infectious disease in animal production systems.
However, because many infectious pathogens are endemic, it will be necessary to understand
how and why the immune system regulates protein accretion in growing animals. This is
particularly important because even animals with subclinical infections have reduced growth.
How cytokines simultaneously influence systems involved in protein synthesis and degradation
in growing animals is needed because certainly the whole is greater than the sum of their effects.

REFERENCES
Alvarez, B., Quinn, L.S., Busquets, S., Lopez-Soriano, F.J., Argiles, J.M., 2002a. TNF-α modulates
cytokine and cytokine receptors in C2C12 myotubes. Cancer Lett. 175, 181–185.
Alvarez, B., Quinn, L.S., Busquets, S., Quiles, M.T., Lopez-Soriano, F.J., Argiles, J.M., 2002b. Tumor
necrosis factor-α exerts interleukin-6-dependent and -independent effects on cultured skeletal muscle
cells. Biochim. Biophys. Acta 1542, 66–72.
Argiles, J.M., Lopez-Soriano, F.J., 1990. The effects of tumour necrosis factor-α (cachectin) and tumour
growth on hepatic amino acid utilization in the rat. Biochem. J. 266, 123–126.
Argiles, J.M., Lopez-Soriano, F.J., Wiggins, D., Williamson, D.H., 1989. Comparative effects of tumour
necrosis factor-α (cachectin), interleukin-1β and tumour growth on amino acid metabolism in the rat
in vivo: absorption and tissue uptake of α-amino[1-14C]isobutyrate. Biochem. J. 261, 357–362.
Aubert, A., Goodall, G., Dantzer, R., 1995. Compared effects of cold ambient temperature and cytokines
on macronutrient intake in rats. Physiol. Behav. 57, 869–873.
Balaji, R., Wright, K.J., Hill, C.M., Dritz, S.S., Knoppel, E.L., Minton, J.E., 2000. Acute phase responses
of pigs challenged orally with Salmonella typhimurium. J. Anim. Sci. 78, 1885–1891.
Ballinger, A.B., Azooz, O., El-Haj, T., Poole, S., Farthing, M.J., 2000. Growth failure occurs through a
decrease in insulin-like growth factor 1 which is independent of undernutrition in a rat model of colitis.
Gut 46, 694–700.
Ballmer, P.E., McNurlan, M.A., Southorn, B.G., Grant, I., Garlick, P.J., 1991. Effects of human recombi-
nant interleukin-1β on protein synthesis in rat tissues compared with a classical acute-phase reaction
induced by turpentine: rapid response of muscle to interleukin-1β. Biochem. J. 279, 683–688.
Ban, E., Haour, F., Lenstra, R., 1992. Brain interleukin 1 gene expression induced by peripheral
lipopolysaccharide administration. Cytokine 4, 48–54.
Banks, W.A., Kastin, A.J., Ehrensing, C.A., 1994a. Blood-borne interleukin-1α is transported across the
endothelial blood-spinal cord barrier of mice. J. Physiol. 479, 257–264.
Banks, W.A., Kastin, A.J., Gutierrez, E.G., 1994b. Penetration of interleukin-6 across the murine
blood-brain barrier. Neurosci. Lett. 179, 53–56.
Cytokine regulation of protein accretion 101

Banks, W.A., Ortiz, L., Plotkin, S.R., Kastin, A.J., 1991. Human interleukin (IL) 1α, murine IL-1α and
murine IL-1β are transported from blood to brain in the mouse by a shared saturable mechanism.
J. Pharmacol. Exp. Ther. 259, 988–996.
Baracos, V.E., DeVivo, C., Hoyle, D.H., Goldberg, A.L., 1995. Activation of the ATP-ubiquitin-
proteasome pathway in skeletal muscle of cachectic rats bearing a hepatoma. Amer. J. Physiol. 268,
E996–E1006.
Blalock, J.E., 1984. The immune system as a sensory organ. J. Immunol. 132, 1067–1070.
Broussard, S., Zhou, J.H., Venters, H.D., Bluthé, R.M., Johnson, R.W., Dantzer, R., Kelley, K.W., 2001.
At the interface of environment-immune interactions: cytokines and growth factor receptors. J. Anim. Sci.
79, E268–E284.
Combaret, L., Tilignac, T., Claustre, A., Voisin, L., Taillandier, D., Obled, C., Tanaka, K., Attaix, D.,
2002. Torbafylline (HWA 448) inhibits enhanced skeletal muscle ubiquitin-proteasome-dependent
proteolysis in cancer and septic rats. Biochem. J. 361, 185–192.
Cooney, R., Kimball, S.R., Eckman, R., Maish, G. 3rd, Shumate, M., Vary, T.C., 1999. TNF-binding
protein ameliorates inhibition of skeletal muscle protein synthesis during sepsis. Amer. J. Physiol.
276, E611–E619.
Cooney, R.N., Kimball, S.R., Vary, T.C., 1997. Regulation of skeletal muscle protein turnover during
sepsis: mechanisms and mediators. Shock 7, 1–16.
Dantzer, R., 2001a. Cytokine-induced sickness behavior: mechanisms and implications. Ann. N.Y. Acad.
Sci. 933, 222–234.
Dantzer, R., 2001b. Cytokine-induced sickness behavior: where do we stand? Brain Behav. Immun.
15, 7–24.
De Benedetti, F., Alonzi, T., Moretta, A., Lazzaro, D., Costa, P., Poli, V., Martini, A., Ciliberto, G., Fattori, E.,
1997. Interleukin 6 causes growth impairment in transgenic mice through a decrease in insulin-like
growth factor-I: a model for stunted growth in children with chronic inflammation. J. Clin. Invest. 99,
643–650.
Delmore, G., 1997. Assessment of nutritional status in cancer patients: widely neglected? Support. Care
Cancer. 5, 376–380.
de Vries, H.E., Blom-Roosemalen, M.C., van Oosten, M., de Boer, A.G., van Berkel, T.J., Breimer, D.D.,
Kuiper, J., 1996. The influence of cytokines on the integrity of the blood-brain barrier in vitro.
J. Neuroimmunol. 64, 37–43.
Dewys, W.D., Begg, C., Lavin, P.T., Band, P.R., Bennett, J.M., Bertino, J.R., Cohen, M.H., Douglass, H.O.
Jr., Engstrom, P.F., Ezdinli, E.Z., Horton, J., Johnson, G.J., Moertel, C.G., Oken, M.M., Perlia, C.,
Rosenbaum, C., Silverstein, M.N., Skeel, R.T., Sponzo, R.W., Tormey, D.C., 1980. Prognostic effect
of weight loss prior to chemotherapy in cancer patients. Eastern Cooperative Oncology Group. Amer.
J. Med. 69, 491–497.
Ebisui, C., Tsujinaka, T., Morimoto, T., Kan, K., Iijima, S., Yano, M., Kominami, E., Tanaka, K., Monden, M.,
1995. Interleukin-6 induces proteolysis by activating intracellular proteases (cathepsins B and L,
proteasome) in C2C12 myotubes. Clin. Sci. (Colch.) 89, 431–439.
Escobar, J., Toepfer, T.L., Van Alstine, W.G., Baker, D.H., Johnson, R.W., 2002. Porcine Reproductive
and Respiratory Syndrome Virus (PRRSV) but not Mycoplasma hyopneumoniae (Mh) decreases
protein accretion and markedly increases skeletal muscle myostatin (MSTN) gene expression in
young pigs. FASEB J. 16, A618–A619.
Essen-Gustavsson, B., Lindholm, A., 1984. Fiber types and metabolic characteristics in muscles of
wild boars, normal and halothane sensitive Swedish landrace pigs. Comp. Biochem. Physiol. A.
78, 67–71.
Faggioni, R., Fuller, J., Moser, A., Feingold, K.R., Grunfeld, C., 1997. LPS-induced anorexia in
leptin-deficient (ob/ob) and leptin receptor-deficient (db/db) mice. Amer. J. Physiol. 273,
R181–R186.
Farges, M.C., Balcerzak, D., Fisher, B.D., Attaix, D., Bechet, D., Ferrara, M., Baracos, V.E., 2002.
Increased muscle proteolysis after local trauma mainly reflects macrophage-associated lysosomal
proteolysis. Amer. J. Physiol. Endocrinol. Metab. 282, E326–E335.
Fattori, E., Cappelletti, M., Costa, P., Sellitto, C., Cantoni, L., Carelli, M., Faggioni, R., Fantuzzi, G.,
Ghezzi, P., Poli, V., 1994. Defective inflammatory response in interleukin 6-deficient mice. J. Exp.
Med. 180, 1243–1250.
Finck, B.N., Johnson, R.W., 2000. Tumor necrosis factor (TNF)-α induces leptin production through the
p55 TNF receptor. Amer. J. Physiol. Regul. Integr. Comp. Physiol. 278, R537–R543.
102 R. W. Johnson and J. Escobar

Finck, B.N., Kelley, K.W., Dantzer, R., Johnson, R.W., 1998. In vivo and in vitro evidence for the
involvement of tumor necrosis factor-α in the induction of leptin by lipopolysaccharide.
Endocrinology 139, 2278–2283.
Fischer, D.R., Sun, X., Williams, A.B., Gang, G., Pritts, T.A., James, J.H., Molloy, M., Fischer, J.E., Paul, R.J.,
Hasselgren, P.O., 2001. Dantrolene reduces serum TNFα and corticosterone levels and muscle
calcium, calpain gene expression, and protein breakdown in septic rats. Shock 15, 200–207.
Flores, E.A., Bistrian, B.R., Pomposelli, J.J., Dinarello, C.A., Blackburn, G.L., Istfan, N.W., 1989.
Infusion of tumor necrosis factor/cachectin promotes muscle catabolism in the rat: a synergistic effect
with interleukin 1. J. Clin. Invest. 83, 1614–1622.
Fong, Y., Moldawer, L.L., Marano, M., Wei, H., Barber, A., Manogue, K., Tracey, K.J., Kuo, G.,
Fischman, D.A., Cerami, A., et al., 1989. Cachectin/TNF or IL-1α induces cachexia with redistribu-
tion of body proteins. Amer. J. Physiol. 256, R659–R665.
French, R.A., Zachary, J.F., Dantzer, R., Frawley, L.S., Chizzonite, R., Parnet, P., Kelley, K.W., 1996.
Dual expression of p80 type I and p68 type II interleukin-I receptors on anterior pituitary cells
synthesizing growth hormone. Endocrinology 137, 4027–4036.
Frost, R.A., Lang, C.H., Gelato, M.C., 1997. Transient exposure of human myoblasts to tumor necrosis
factor-α inhibits serum and insulin-like growth factor-I stimulated protein synthesis. Endocrinology
138, 4153–4159.
Furuno, K., Goodman, M.N., Goldberg, A.L., 1990. Role of different proteolytic systems in the
degradation of muscle proteins during denervation atrophy. J. Biol. Chem. 265, 8550–8557.
Garcia-Martinez, C., Agell, N., Llovera, M., Lopez-Soriano, F.J., Argiles, J.M., 1993. Tumour necrosis
factor-α increases the ubiquitinization of rat skeletal muscle proteins. FEBS Lett. 323, 211–214.
Garcia-Martinez, C., Llovera, M., Agell, N., Lopez-Soriano, F.J., Argiles, J.M., 1995. Ubiquitin gene
expression in skeletal muscle is increased during sepsis: involvement of TNF-α but not IL-1.
Biochem. Biophys. Res. Commun. 217, 839–844.
Gayle, D., Ilyin, S.E., Plata-Salaman, C.R., 1997. Central nervous system IL-1β system and neuropep-
tide Y mRNAs during IL-1β-induced anorexia in rats. Brain Res. Bull. 44, 311–317.
Geiger, T., Andus, T., Klapproth, J., Hirano, T., Kishimoto, T., Heinrich, P.C., 1988. Induction of rat
acute-phase proteins by interleukin 6 in vivo. Eur. J. Immunol. 18, 717–721.
Gingras, A.C., Raught, B., Sonenberg, N., 1999. eIF4 initiation factors: effectors of mRNA recruitment
to ribosomes and regulators of translation. Annu. Rev. Biochem. 68, 913–963.
Goehler, L.E., Gaykema, R.P., Nguyen, K.T., Lee, J.E., Tilders, F.J., Maier, S.F., Watkins, L.R., 1999.
Interleukin-1β in immune cells of the abdominal vagus nerve: a link between the immune and
nervous systems? J. Neurosci. 19, 2799–2806.
Goehler, L.E., Relton, J.K., Dripps, D., Kiechle, R., Tartaglia, N., Maier, S.F., Watkins, L.R., 1997. Vagal
paraganglia bind biotinylated interleukin-1 receptor antagonist: a possible mechanism for immune-
to-brain communication. Brain Res. Bull. 43, 357–364.
Goodman, M.N., 1994. Interleukin-6 induces skeletal muscle protein breakdown in rats. Proc. Soc.
Exp. Biol. Med. 205, 182–185.
Gorbea, C., Taillandier, D., Rechsteiner, M., 1999. Assembly of the regulatory complex of the 26S
proteasome. Mol. Biol. Rep. 26, 15–19.
Gutierrez, E.G., Banks, W.A., Kastin, A.J., 1993. Murine tumor necrosis factor α is transported from
blood to brain in the mouse. J. Neuroimmunol. 47, 169–176.
Hart, B.L., 1988. Biological basis of the behavior of sick animals. Neurosci. Biobehav. Rev. 12,
123–137.
Hasselgren, P.O., 2000. Catabolic response to stress and injury: implications for regulation. World J. Surg.
24, 1452–1459.
Hasselgren, P.O., Fischer, J.E., 2001. Muscle cachexia: current concepts of intracellular mechanisms and
molecular regulation. Ann. Surg. 233, 9–17.
Hasselgren, P.O., Wray, C., Mammen, J., 2002. Molecular regulation of muscle cachexia: it may be more
than the proteasome. Biochem. Biophys. Res. Commun. 290, 1–10.
Hobler, S.C., Tiao, G., Fischer, J.E., Monaco, J., Hasselgren, P.O., 1998. Sepsis-induced increase in
muscle proteolysis is blocked by specific proteasome inhibitors. Amer. J. Physiol. 274, R30–R37.
Huang, J., Forsberg, N.E., 1998. Role of calpain in skeletal-muscle protein degradation. Proc. Natl. Acad.
Sci. USA 95, 12100–12105.
Huang, Q.H., Hruby, V.J., Tatro, J.B., 1999. Role of central melanocortins in endotoxin-induced anorexia.
Amer. J. Physiol. 276, R864–R871.
Cytokine regulation of protein accretion 103

Johnson, R.W., 1997. Inhibition of growth by pro-inflammatory cytokines: an integrated view. J. Anim.
Sci. 75, 1244–1255.
Johnson, R.W., 2002. The concept of sickness behavior: a brief chronological account of four key
discoveries. Vet. Immunol. Immunopathol. 87, 443–450.
Johnson, R.W., Finck, B.N., 2001. Tumor necrosis factor α and leptin: two players in animals’ metabolic
and immunologic responses to infection. J. Anim. Sci. 79, E118–E127.
Kent, S., Bluthe, R.M., Kelley, K.W., Dantzer, R., 1992. Sickness behavior as a new target for drug devel-
opment. Trends Pharmacol. Sci. 13, 24–28.
Kimball, S.R., 2002. Regulation of global and specific mRNA translation by amino acids. J. Nutr.
132, 883–886.
Klasing, K.C., 1988. Nutritional aspects of leukocytic cytokines. J. Nutr. 118, 1436–1446.
Klasing, K.C., Austic, R.E., 1984. Changes in protein synthesis due to an inflammatory challenge. Proc.
Soc. Exp. Biol. Med. 176, 285–291.
Konsman, J.P., Dantzer, R., 2001. How the immune and nervous systems interact during disease-associated
anorexia. Nutrition 17, 664–668.
Konsman, J.P., Kelley, K., Dantzer, R., 1999. Temporal and spatial relationships between lipopolysaccharide-
induced expression of Fos, interleukin-1β and inducible nitric oxide synthase in rat brain.
Neuroscience 89, 535–548.
Koohmaraie, M., 1992. The role of Ca2+-dependent proteases (calpains) in post mortem proteolysis and
meat tenderness. Biochimie. 74, 239–245.
Lang, C.H., Frost, R.A., Jefferson, L.S., Kimball, S.R., Vary, T.C., 2000. Endotoxin-induced decrease in
muscle protein synthesis is associated with changes in eIF2B, eIF4E, and IGF-I. Amer. J. Physiol.
Endocrinol. Metab. 278, E1133–E1143.
Lang, C.H., Frost, R.A., Nairn, A.C., MacLean, D.A., Vary, T.C., 2002. TNF-α impairs heart and skeletal
muscle protein synthesis by altering translation initiation. Amer. J. Physiol. Endocrinol. Metab. 282,
E336–E347.
Langhans, W., Savoldelli, D., Weingarten, S., 1993. Comparison of the feeding responses to bacterial
lipopolysaccharide and interleukin-1 β. Physiol. Behav. 53, 643–649.
Laye, S., Bluthe, R.M., Kent, S., Combe, C., Medina, C., Parnet, P., Kelley, K., Dantzer, R., 1995.
Subdiaphragmatic vagotomy blocks induction of IL-1β mRNA in mice brain in response to peripheral
LPS. Amer. J. Physiol. 268, R1327–R1331.
Laye, S., Parnet, P., Goujon, E., Dantzer, R., 1994. Peripheral administration of lipopolysaccharide induces
the expression of cytokine transcripts in the brain and pituitary of mice. Brain Res. Mol. Brain Res.
27, 157–162.
LeRoith, D., 2000. Insulin-like growth factor I receptor signaling: overlapping or redundant pathways?
Endocrinology 141, 1287–1288.
Li, Y.P., Reid, M.B., 2000. NF-κB mediates the protein loss induced by TNF-α in differentiated skeletal
muscle myotubes. Amer. J. Physiol. Regul. Integr. Comp. Physiol. 279, R1165–R1170.
Li, Y.P., Schwartz, R.J., Waddell, I.D., Holloway, B.R., Reid, M.B., 1998. Skeletal muscle myocytes
undergo protein loss and reactive oxygen-mediated NF-κB activation in response to tumor necrosis
factor α. FASEB J. 12, 871–880.
Ling, P.R., Schwartz, J.H., Bistrian, B.R., 1997. Mechanisms of host wasting induced by administration
of cytokines in rats. Amer. J. Physiol. 272, E333–E339.
Llovera, M., Carbo, N., Garcia-Martinez, C., Costelli, P., Tessitore, L., Baccino, F.M., Agell, N.,
Bagby, G.J., Lopez-Soriano, F.J., Argiles, J.M., 1996. Anti-TNF treatment reverts increased muscle
ubiquitin gene expression in tumour-bearing rats. Biochem. Biophys. Res. Commun. 221, 653–655.
Llovera, M., Carbo, N., Lopez-Soriano, J., Garcia-Martinez, C., Busquets, S., Alvarez, B., Agell, N.,
Costelli, P., Lopez-Soriano, F.J., Celada, A., Argiles, J.M., 1998a. Different cytokines modulate ubiq-
uitin gene expression in rat skeletal muscle. Cancer Lett. 133, 83–87.
Llovera, M., Garcia-Martinez, C., Agell, N., Lopez-Soriano, F.J., Argiles, J.M., 1995. Muscle wasting
associated with cancer cachexia is linked to an important activation of the ATP-dependent ubiquitin-
mediated proteolysis. Int. J. Cancer. 61, 138–141.
Llovera, M., Garcia-Martinez, C., Agell, N., Marzabal, M., Lopez-Soriano, F.J., Argiles, J.M., 1994. Ubiquitin
gene expression is increased in skeletal muscle of tumour-bearing rats. FEBS Lett. 338, 311–318.
Llovera, M., Garcia-Martinez, C., Lopez-Soriano, J., Agell, N., Lopez-Soriano, F.J., Garcia, I., Argiles, J.M.,
1998b. Protein turnover in skeletal muscle of tumour-bearing transgenic mice overexpressing the sol-
uble TNF receptor-1. Cancer Lett. 130, 19–27.
104 R. W. Johnson and J. Escobar

Llovera, M., Garcia-Martinez, C., Lopez-Soriano, J., Carbo, N., Agell, N., Lopez-Soriano, F.J., Argiles, J.M.,
1998c. Role of TNF receptor 1 in protein turnover during cancer cachexia using gene knockout mice.
Mol. Cell. Endocrinol. 142, 183–189.
McHugh, K.J., Collins, S.M., Weingarten, H.P., 1994. Central interleukin-1 receptors contribute to
suppression of feeding after acute colitis in the rat. Amer. J. Physiol. 266, R1659–R1663.
Murray, M.J., Murray, A.B., 1979. Anorexia of infection as a mechanism of host defense. Amer. J. Clin.
Nutr. 32, 593–596.
Nakae, J., Kido, Y., Accili, D., 2001. Distinct and overlapping functions of insulin and IGF-I receptors.
Endocr. Rev. 22, 818–835.
Plata-Salaman, C.R., 1988. Food intake suppression by growth factors and platelet peptides by direct
action in the central nervous system. Neurosci. Lett. 94, 161–166.
Plata-Salaman, C.R., Oomura, Y., Kai, Y., 1988. Tumor necrosis factor and interleukin-1β: suppression
of food intake by direct action in the central nervous system. Brain Res. 448, 106–114.
Rahelic, S., Puac, S., 1981. Fiber types in longissimus dorsi from wild and highly selected pig breeds.
Meat Sci. 5, 439.
Reeds, P.J., Fjeld, C.R., Jahoor, F., 1994. Do the differences between the amino acid compositions
of acute-phase and muscle proteins have a bearing on nitrogen loss in traumatic states? J. Nutr.
124, 906–910.
Roth, J., Martin, D., Storr, B., Zeisberger, E., 1998. Neutralization of pyrogen-induced tumour necrosis
factor by its type 1 soluble receptor in guinea-pigs: effects on fever and interleukin-6 release.
J. Physiol. 509, 267–275.
Roubenoff, R., 2000. Acquired immunodeficiency syndrome wasting, functional performance, and quality
of life. Amer. J. Manag. Care 6, 1003–1016.
Shah, O.J., Anthony, J.C., Kimball, S.R., Jefferson, L.S., 2000. 4E-BP1 and S6K1: translational integra-
tion sites for nutritional and hormonal information in muscle. Amer. J. Physiol. Endocrinol. Metab.
279, E715–E729.
Solomon, V., Goldberg, A.L., 1996. Importance of the ATP-ubiquitin-proteasome pathway in the degra-
dation of soluble and myofibrillar proteins in rabbit muscle extracts. J. Biol. Chem. 271,
26690–26697.
Solorzano, C.C., Kaibara, A., Hess, P.J., Edwards, P.D., Ksontini, R., Abouhamze, A., McDaniel, S.,
Frazier, J., Trujillo, D., Kieft, G., Seely, J., Kohno, T., Cosenza, M.E., Clare-Salzler, M., MacKay, S.L.,
Martin, S.W., Moldawer, L.L., Edwards, C.K. 3rd, 1998. Pharmacokinetics, immunogenicity, and
efficacy of dimeric TNFR binding proteins in healthy and bacteremic baboon. J. Appl. Physiol.
84, 1119–1130.
Stracher, A., 1999. Calpain inhibitors as therapeutic agents in nerve and muscle degeneration. Ann.
N.Y. Acad. Sci. 884, 52–59.
Suda, T., Tozawa, F., Ushiyama, T., Sumitomo, T., Yamada, M., Demura, H., 1990. Interleukin-1
stimulates corticotropin-releasing factor gene expression in rat hypothalamus. Endocrinology
126, 1223–1228.
Svanberg, E., Frost, R.A., Lang, C.H., Isgaard, J., Jefferson, L.S., Kimball, S.R., Vary, T.C., 2000.
IGF-I/IGFBP-3 binary complex modulates sepsis-induced inhibition of protein synthesis in skeletal
muscle. Amer. J. Physiol. Endocrinol. Metab. 279, E1145–E1158.
Swiergiel, A.H., Dunn, A.J., 1999. The roles of IL-1, IL-6, and TNFα in the feeding responses to
endotoxin and influenza virus infection in mice. Brain Behav. Immun. 13, 252–265.
Tanahashi, N., Kawahara, H., Murakami, Y., Tanaka, K., 1999. The proteasome-dependent proteolytic
system. Mol. Biol. Rep. 26, 3–9.
Tawa, N.E. Jr., Odessey, R., Goldberg, A.L., 1997. Inhibitors of the proteasome reduce the accelerated
proteolysis in atrophying rat skeletal muscles. J. Clin. Invest. 100, 197–203.
Temparis, S., Asensi, M., Taillandier, D., Aurousseau, E., Larbaud, D., Obled, A., Bechet, D., Ferrara, M.,
Estrela, J.M., Attaix, D., 1994. Increased ATP-ubiquitin-dependent proteolysis in skeletal muscles of
tumor-bearing rats. Cancer Res. 54, 5568–5573.
Tiao, G., Fagan, J.M., Samuels, N., James, J.H., Hudson, K., Lieberman, M., Fischer, J.E., Hasselgren, P.O.,
1994. Sepsis stimulates nonlysosomal, energy-dependent proteolysis and increases ubiquitin mRNA
levels in rat skeletal muscle. J. Clin. Invest. 94, 2255–2264.
Tracey, K.J., Morgello, S., Koplin, B., Fahey, T.J. 3rd, Fox, J., Aledo, A., Manogue, K.R., Cerami, A.,
1990. Metabolic effects of cachectin/tumor necrosis factor are modified by site of production:
cachectin/tumor necrosis factor-secreting tumor in skeletal muscle induces chronic cachexia,
Cytokine regulation of protein accretion 105

while implantation in brain induces predominantly acute anorexia. J. Clin. Invest. 86,
2014–2024.
Tsujinaka, T., Fujita, J., Ebisui, C., Yano, M., Kominami, E., Suzuki, K., Tanaka, K., Katsume, A.,
Ohsugi, Y., Shiozaki, H., Monden, M., 1996. Interleukin 6 receptor antibody inhibits muscle atrophy
and modulates proteolytic systems in interleukin 6 transgenic mice. J. Clin. Invest. 97, 244–249.
Uehara, A., Sekiya, C., Takasugi, Y., Namiki, M., Arimura, A., 1989. Anorexia induced by interleukin 1:
involvement of corticotropin-releasing factor. Amer. J. Physiol. 257, R613–R617.
van Dam, A.M., Brouns, M., Louisse, S., Berkenbosch, F., 1992. Appearance of interleukin-1 in
macrophages and in ramified microglia in the brain of endotoxin-treated rats: a pathway for the induc-
tion of non-specific symptoms of sickness? Brain Res. 588, 291–296.
van Reeth, K., Nauwynck, H., 2000. Proinflammatory cytokines and viral respiratory disease in pigs.
Vet. Res. 31, 187–213.
van Reeth, K., Labarque, G., Nauwynck, H., Pensaert, M., 1999. Differential production of proinflam-
matory cytokines in the pig lung during different respiratory virus infections: correlations with
pathogenicity. Res. Vet. Sci. 67, 47–52.
Vary, T.C., Kimball, S.R., 1992. Sepsis-induced changes in protein synthesis: differential effects on
fast- and slow-twitch muscles. Amer. J. Physiol. 262, C1513–C1519.
Vary, T.C., Kimball, S.R., 2000. Effect of sepsis on eIE4E availability in skeletal muscle. Amer.
J. Physiol. Endocrinol. Metab. 279, E1178–E1184.
Vary, T.C., Jefferson, L.S., Kimball, S.R., 2001. Insulin fails to stimulate muscle protein synthesis in
sepsis despite unimpaired signaling to 4E-BP1 and S6K1. Amer. J. Physiol. Endocrinol. Metab. 281,
E1045–E1053.
Vary, T.C., Jurasinski, C.V., Karinch, A.M., Kimball, S.R., 1994. Regulation of eukaryotic initiation
factor-2 expression during sepsis. Amer. J. Physiol. 266, E193–E201.
Vary, T.C., Owens, E.L., Beers, J.K., Verner, K., Cooney, R.N., 1996. Sepsis inhibits synthesis of
myofibrillar and sarcoplasmic proteins: modulation by interleukin-1 receptor antagonist. Shock
6, 13–18.
Venters, H.D., Tang, Q., Liu, Q., VanHoy, R.W., Dantzer, R., Kelley, K.W., 1999. A new mechanism of
neurodegeneration: a proinflammatory cytokine inhibits receptor signaling by a survival peptide.
Proc. Natl. Acad. Sci. USA 96, 9879–9884.
Voges, D., Zwickl, P., Baumeister, W., 1999. The 26S proteasome: a molecular machine designed for
controlled proteolysis. Annu. Rev. Biochem. 68, 1015–1068.
Voisin, L., Breuille, D., Combaret, L., Pouyet, C., Taillandier, D., Aurousseau, E., Obled, C. Attaix, D.,
1996a. Muscle wasting in a rat model of long-lasting sepsis results from the activation of lysosomal,
Ca2+-activated, and ubiquitin-proteasome proteolytic pathways. J. Clin. Invest. 97, 1610–1617.
Voisin, L., Gray, K., Flowers, K.M., Kimball, S.R., Jefferson, L.S., Vary, T.C., 1996b. Altered expres-
sion of eukaryotic initiation factor 2B in skeletal muscle during sepsis. Amer. J. Physiol. 270,
E43–E50.
Webel, D.M., Finck, B.N., Baker, D.H., Johnson, R.W., 1997. Time course of increased plasma cytokines,
cortisol, and urea nitrogen in pigs following intraperitoneal injection of lipopolysaccharide. J. Anim. Sci.
75, 1514–1520.
Webel, D.M., Johnson, R.W., Baker, D.H., 1998. Lipopolysaccharide-induced reductions in food intake
do not decrease the efficiency of lysine and threonine utilization for protein accretion in chickens.
J. Nutr. 128, 1760–1766.
Williams, A., Wang, J.J., Wang, L., Sun, X., Fischer, J.E., Hasselgren, P.O., 1998. Sepsis in mice stimu-
lates muscle proteolysis in the absence of IL-6. Amer. J. Physiol. 275, R1983–R1991.
Williams, A.B., Decourten-Myers, G.M., Fischer, J.E., Luo, G., Sun, X., Hasselgren, P.O., 1999. Sepsis
stimulates release of myofilaments in skeletal muscle by a calcium-dependent mechanism. FASEB J.
13, 1435–1443.
Williams, N.H., Stahly, T.S., Zimmerman, D.R., 1997a. Effect of chronic immune system activation on
body nitrogen retention, partial efficiency of lysine utilization, and lysine needs of pigs. J. Anim. Sci.
75, 2472–2480.
Williams, N.H., Stahly, T.S., Zimmerman, D.R., 1997b. Effect of chronic immune system activation on
the rate, efficiency, and composition of growth and lysine needs of pigs fed from 6 to 27 kg. J. Anim. Sci.
75, 2463–2471.
Williams, N.H., Stahly, T.S., Zimmerman, D.R., 1997c. Effect of level of chronic immune system activation
on the growth and dietary lysine needs of pigs fed from 6 to 112 kg. J. Anim. Sci. 75, 2481–2496.
106 R. W. Johnson and J. Escobar

Wright, K.J., Balaji, R., Hill, C.M., Dritz, S.S., Knoppel, E.L., Minton, J.E., 2000. Integrated adrenal,
somatotropic, and immune responses of growing pigs to treatment with lipopolysaccharide. J. Anim.
Sci. 78, 1892–1899.
Zamir, O., Hasselgren, P.O., Kunkel, S.L., Frederick, J., Higashiguchi, T., Fischer, J.E., 1992. Evidence
that tumor necrosis factor participates in the regulation of muscle proteolysis during sepsis. Arch.
Surg. 127, 170–174.
Zhang, Y., Pilon, G., Marette, A., Baracos, V.E., 2000. Cytokines and endotoxin induce cytokine recep-
tors in skeletal muscle. Amer. J. Physiol. Endocrinol. Metab. 279, E196–E205.
5 Amino acid metabolism in the
small intestine: biochemical bases
and nutritional significance1

G. Wua, D. A. Knabea, and N. E. Flynnb

aDepartment of Animal Science and Faculty of Nutrition,

Texas A & M University, College Station, Texas, TX 77843-2471, USA


bDepartment of Chemistry and Biochemistry, Angelo State University,

San Angelo, TX 76909, USA

The small intestine is a highly differentiated and complex organ, which is not only responsi-
ble for the terminal digestion and absorption of nutrients, but also plays an important role in
amino acid metabolism. Most of glutamine and almost all of glutamate and aspartate in the
diet are catabolized by the small intestinal mucosa in the first pass. The small intestinal
mucosa also degrades enteral arginine, ornithine, proline, branched-chain amino acids and
lysine, and perhaps enteral methionine, phenylalanine, threonine, glycine, and serine, such
that 30−50% of these dietary amino acids do not enter the portal circulation. In the post-
absorptive state, the small intestine actively takes up arterial glutamine and releases ammonia,
alanine, citrulline, and proline as the major nitrogenous products. The intestine-derived citrulline
is effectively utilized for arginine synthesis by extrahepatic cells and organs (e.g. the kidneys).
This is of nutritional significance, particularly for suckling neonates because the milk of most
species, including the pig, cattle, sheep, rat, and human, is remarkably deficient in arginine.
In addition to hepatic gluconeogenesis, the alanine released by the small intestine plays a key
role in the extensive recycling of nitrogen between the liver and the gut. Because dietary
amino acids are major fuels for the small intestinal mucosa, and are essential precursors for
intestinal synthesis of proteins, glutathione, polyamines, nitric oxide, purine, and pyrimidine
nucleotides, intestinal amino acid metabolism is obligatory for maintaining intestinal mucosal
mass, function, and integrity. However, the extensive catabolism of enteral amino acids by the

1We thank our students, technicians, and collaborators who have contributed to the work cited here. Our research on
intestinal amino acid metabolism was supported, in part, by USDA National Research Initiative competitive grants
No. 92-37206-8004, No. 94-37206-1100, No. 97-35206-5096, No. 2001-35203-11247 and No. 2003-35206-13694
(GW), by Hatch projects No. 8200 (GW) and No. 6601 (DAK) from the Texas Agricultural Experiment Station, by
grants from the Houston Livestock Show and Rodeo (GW and DAK), and by a Texas A & M University Faculty
Fellowship (GW). This paper is dedicated to the memory of Dr. Peter J. Reeds, our dear friend and mentor.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
107 © 2005 Elsevier Limited. All rights reserved.
108 G. Wu et al.

small intestine substantially reduces their availability to extraintestinal tissues and selectively
alters the patterns of amino acids that enter the systemic circulation. This has important practical
implications for the utilization efficiency and recommended requirements of dietary protein
and amino acids by animals, including humans.

1. INTRODUCTION
The small intestine is a highly differentiated and complex organ, which is responsible for the
terminal digestion and absorption of dietary nutrients and therefore is essential to health,
growth, development, reproduction, and sustaining life of the organism (Madara, 1991).
Enterocytes (epithelial absorptive cells of the small intestine) constitute >80% of the mucosal
epithelial cell population (Cheng and Leblond, 1974; Klein and McKenzie, 1983) and have
high rates of intracellular protein turnover and cell proliferation (Smith and Jarvis, 1978;
Burrin and Reeds, 1997). Interestingly, the apical and basolateral membranes of each entero-
cyte are chemically, biochemically, and physically distinct (Madara, 1991). Such a polar
organization of the enterocyte allows it to selectively receive nutrients from two sources: the
arterial blood across its basolateral membrane and the intestinal lumen across its brush border
membrane. This has important practical implications for choosing the route of feeding (e.g.
enteral vs parenteral) for nutrient delivery to animals.
The gut is also the barrier separating the internal milieu of the organism from the external
environment, therefore excluding food-borne pathogens and preventing the translocation of
luminal microorganisms into the circulation. As the largest lymphoid organ in the body, the
small intestine participates in immune surveillance of the intestinal epithelial layer and regu-
lation of the mucosal response to foreign antigens (Mowat, 1987). The pioneering studies of
Windmueller and coworkers in the 1970s have demonstrated extensive intestinal catabolism
of glutamine, glutamate, and aspartate (see Windmueller, 1982, for review). In recent years,
there has been growing recognition that the small intestinal mucosa also degrades enteral
arginine, ornithine, proline, branched-chain amino acids (BCAA), and lysine, and perhaps
enteral methionine, phenylalanine, threonine, glycine, and serine, such that 30−50% of these
dietary amino acids do not enter the portal circulation (Wu, 1998a; Reeds and Burrin, 2001).
The major objective of this chapter is to review recent work on intestinal amino acid metab-
olism, with an emphasis on its biochemical bases and nutritional significance.

2. AMINO ACID METABOLISM IN THE SMALL INTESTINE


Amino acids can be classified into four major groups on the basis of their metabolic fates in
the small intestinal mucosa of rats, pigs, and ruminants (sheep and cattle): (1) amino acids
that are neither synthesized nor degraded; (2) amino acids that are synthesized but not
degraded; (3) amino acids that are degraded but not synthesized; and (4) amino acids that are
both degraded and synthesized (table 1). Note that there are species differences in intestinal
amino acid metabolism (Wu, 1998b).

2.1. Amino acids that are neither synthesized nor degraded by the intestinal mucosa

2.1.1. Asparagine

There is no synthesis of asparagine from glutamine and aspartate in enterocytes of rats, pigs
(Wu et al., 1995b; Wu, 1998b), and ruminants (sheep and cattle) (Wu, unpublished data).
Amino acid metablism in the small intestine 109

Table 1
Metabolic fates of amino acid in the small intestine

Metabolic fates Amino acids

Neither synthesized nor degraded Asparagine, cysteine, histidine, tryptophan


Synthesized but not degraded Tyrosine
Degraded but not synthesized Branched-chain amino acids (isoleucine, leucine, valine),
lysine, methionine, phenylalanine, threonine
Both degraded and synthesized Alanine, arginine, aspartate, citrulline, glutamate, glutamine,
glycine, ornithine, proline, serine

Asparagine is not degraded by the rat small intestine, because of the absence of asparaginase;
thus in rats all of the asparagine absorbed by enterocytes from the intestinal lumen appears in
the intestinal venous blood intact (Windmueller, 1982). There is no release of asparagine by
the postabsorptive small intestine of pigs (Wu et al., 1994a) and rats (Brosnan et al., 1983),
indicating the absence of asparagine synthesis by the gut. Interestingly, asparaginase activity
is detectable in the canine small intestine, and asparagine is hydrolyzed to aspartate plus
ammonia in the guinea pig small intestine (Windmueller, 1982), indicating species differ-
ences in intestinal asparagine metabolism. However, in both dogs and guinea pigs, intestinal
mucosal asparagine catabolism is quantitatively low.

2.1.2. Cysteine, tryptophan, and histidine

Cysteine, tryptophan, and histidine are neither synthesized nor degraded by enterocytes of
rats, pigs, sheep, and cattle (Wu, unpublished data). However, cysteine is used for glutathione
synthesis in enterocytes (Reeds et al., 1997). Because there are mixed cell populations in the
small intestine, the infiltrating mast cells can decarboxylate histidine to produce histamine in
response to immunological activation (Wu, 1998b), thereby contributing to intestinal histidine
utilization by intestinal mucosal cells and the portal-drained viscera (PDV).

2.2. Amino acids that are synthesized but not degraded by the intestinal mucosa

2.2.1. Tyrosine

Tyrosine is synthesized from phenylalanine by phenylalanine hydroxylase, a tetrahydro-


biopterin-dependent enzyme. This enzyme is restricted primarily to the liver but is also
expressed in the kidney and pancreas (Tourian et al., 1969). Phenylalanine hydroxylase was
previously found to be absent from the small intestine (Tourian et al., 1969). However, in
these earlier studies, protease inhibitors were not used for preparing intestinal extracts or the
enzyme assay, and thus intestinal activity of phenylalanine hydroxylase should be reexam-
ined. We have recently found that phenylalanine was converted into tyrosine in enterocytes of
pigs, rats, sheep, and cattle (Wu, unpublished data), indicating the presence of intestinal
phenylalanine hydroxylation. In support of this view, there is significant output of tyrosine
(167% of dietary intake) by the PDV of the milk protein-fed pig (Stoll et al., 1998) and of
tyrosine (28%) by the small intestine of sheep fed a 20% crude protein diet (Tagari and
Bergman, 1978).
110 G. Wu et al.

2.3. Amino acids that are degraded but not synthesized by the intestinal mucosa

2.3.1. BCAA

Studies from a number of species have documented BCAA catabolism by the small intestine
of nonruminant animals. For example, 30% of the total ingested dietary leucine is extracted
by the dog small intestine in the first pass, 55% of which enters the transamination reaction
(Yu et al., 1990). In adult humans, 20−30% of enterally delivered leucine is taken up in the
first pass within the splanchnic region (mainly the small intestine) during the postabsorptive
state or feeding (Hoerr et al., 1993). Similarly, in growing pigs fed a milk protein-based diet,
40% of leucine, 30% of isoleucine, and 40% of valine in the diet are sequestered by the PDV
in the first pass, and <20% of the sequestered BCAA are utilized for mucosal protein syn-
thesis (Stoll et al., 1998). These results suggest substantial catabolism of dietary BCAA by
the small intestine of humans, dogs, and pigs. In more recent studies, we found that in pig
enterocytes, most of the transaminated BCAA were released as branched-chain α-ketoacid
(BCKA) (Wu, unpublished data), indicating a low rate of oxidative decarboxylation of BCKA
by the small intestinal mucosa. In contrast to nonruminant animals, BCAA catabolism is
negligible in the small intestinal mucosa of fed or fasted sheep (Pell et al., 1986; Cappelli
et al., 1997).
Both BCAA transaminase, which initiates BCAA degradation, and BCKA dehydrogenase,
which oxidatively decarboxylates BCKA, are present in the nonruminant small intestinal
mucosa (Khatra et al., 1977; Harper et al., 1984). The specific activities of BCAA transaminase
and BCKA dehydrogenase are relatively low in the small intestine, compared with skeletal
muscle and liver, respectively. However, this should not negate a quantitatively important role
of the small intestine in BCAA catabolism in the whole animal, partly due to a relatively large
mass of the gut. On the basis of the intestinal activities of BCAA transaminase and BCKA
dehydrogenase, most of the BCKA produced by enterocytes likely enters the portal circulation
and are then utilized by the liver for complete oxidation and/or gluconeogenesis.

2.3.2. Lysine, methionine, phenylalanine, and threonine

The degradation of these four amino acids has recently been demonstrated in the small intes-
tine, such that ~50% of dietary lysine and methionine, 45% of dietary phenylalanine, and
60% of dietary threonine are extracted in the first pass by the PDV of milk protein-fed pigs
(Stoll et al., 1998). Only <20% of the extracted amino acids are utilized for mucosal protein
synthesis, whereas one-third of the extracted amino acids are catabolized by the small intes-
tinal mucosa such that intestinal metabolism dominates the splanchnic extraction of lysine,
methionine, phenylalanine, and threonine in pigs (Stoll et al., 1997, 1998). Using stable iso-
topes, van Goudoever et al. (2000) showed that intestinal oxidation of enteral lysine
contributed one-third of total body lysine oxidation in growing pigs fed a high-protein diet,
but was virtually absent in pigs fed a low-protein diet. These results indicate adaptive regula-
tion of intestinal lysine metabolism. In support of intestinal lysine oxidation, Pink et al.
(2002) recently reported the production of CO2 from [14C]lysine in mitochondria of pig ente-
rocytes. In adult humans, 30% and 58% of enterally delivered lysine and phenylalanine are
extracted in the first pass, respectively, within the splanchnic bed (Biolo et al., 1992; Hoerr
et al., 1993). Remarkably, in sheep fed 16−20% crude protein diets, 30−40% of lysine,
37−43% of phenylalanine, 55−73% of threonine, and 69−71% of methionine that disappeared
from the small intestinal lumen did not enter the portal circulation (Tagari and Bergman,
1978), suggesting extensive catabolism of these amino acids by the ovine small intestine.
Amino acid metablism in the small intestine 111

There is evidence for the presence of key enzymes responsible for the degradation of
lysine, methionine, phenylalanine, and threonine in the intestinal mucosa. For example, Pink
et al. (2002) recently detected lysine α-ketoglutarate reductase activity in the pig small intes-
tinal mucosa when the enzyme assay was conducted in the presence of protease inhibitors.
Also, both methionine transamination (Mitchell and Benevenga, 1978) and the transsulfura-
tion pathway (Luk et al., 1980) are present in the rat small intestinal mucosa. In addition,
glutamine transaminase K, whose major substrates include glutamine, phenylalanine, and
methionine, is widespread in mammalian tissues, including the small intestine of rats (Cooper
and Meister, 1977) and pigs (Wu, unpublished data). Furthermore, there is an inherent phenyl-
alanine transaminase activity in porcine aspartate transaminase isoenzymes (Shrawder and
Martinez-Carrion, 1972). More recently, we were able to detect threonine dehydrogenase
activity in mitochondria of pig enterocytes when protease inhibitors were used for preparing
cell extracts and the enzyme assay (Wu, unpublished data).

2.4. Amino acids that are both degraded and synthesized by the intestinal mucosa

2.4.1. Glutamine, glutamate, aspartate, and alanine

The pioneering studies of Windmueller and coworkers have clearly demonstrated that the small
intestine extensively catabolizes enteral glutamine, glutamate, and aspartate as well as arterial
glutamine, and releases large amounts of alanine, citrulline, and proline (Windmueller, 1982). For
example, the small intestine of postabsorptive rats extracts 30% of arterial glutamine in a single
pass, which accounts for 30% of whole-body glutamine utilization (Windmueller and Spaeth,
1975). Interestingly, intestinal utilization of arterial glutamine appears to be lower in ruminants,
compared with nonruminants (Gate et al., 1999). Approximately 55%, 66%, and almost 100%
of enterally delivered glutamine are sequestered in the first pass by the small intestine of adult
humans (Matthews et al., 1993), adult rats (Windmueller and Spaeth, 1975), and growing pigs
(Stoll et al., 1998), respectively. In contrast, there is no significant uptake of arterial glutamate
and aspartate by the small intestine. Remarkably, 98% and >99% of luminal glutamate and
aspartate (6 mM) are catabolized in a single pass by the rat jejunum, respectively (Windmueller,
1982). Similarly, 96% and 95% of enterally delivered glutamate is extracted in the first pass by
the human splanchnic bed (Battezzati et al., 1995) and by the porcine PDV (Reeds et al., 1996),
respectively. Likewise, there is negligible appearance of intra-abomasum infused glutamate in
the portal circulation of sheep (Tagari and Bergman, 1978). Furthermore, the PDV of growing beef
steers extracts virtually all diet- and rumen-derived glutamate, glutamine, and aspartate (Lapierre
et al., 2000). Thus, most of glutamine and almost all glutamate and aspartate in the diet do not
enter portal circulation in both ruminant and nonruminant animals.
The metabolic fate of glutamine, glutamate, and aspartate has been quantified in the small
intestine. Ammonia, citrulline, alanine, and proline released by the rat jejunum account for
38%, 28%, 24%, and 7% of the metabolized glutamine nitrogen, respectively. In postabsorp-
tive pigs, the small intestine takes up arterial glutamine and releases not only ammonia,
citrulline, alanine, and proline but also arginine, glutamate, and aspartate (Wu et al., 1994a).
In vitro studies have also shown that pig enterocytes extensively utilize glutamine and
produce ammonia, glutamate, alanine, aspartate, CO2, ornithine, citrulline, arginine, and
proline, and rates of glutamine catabolism are greater in cells from newborn pigs compared with
suckling and postweanling pigs (Wu et al., 1995b). Ammonia, glutamate, alanine, aspartate, CO2,
ornithine, citrulline, arginine, and proline are also produced from glutamine in enterocytes of
sheep and cattle (Wu, unpublished data). Interestingly, there is no production of ammonia
112 G. Wu et al.

from glutamate and aspartate in the small intestine (Windmueller, 1982), suggesting a dominant
role of transamination in their catabolism. In the rat small intestine, CO2, lactate, alanine, and
glucose account for 56−64%, 16−20%, 4−8%, and 2−10% of the total catabolized carbons of
luminal glutamine, glutamate, and aspartate, respectively. Under conditions similar to a meal,
oxidation of arterial glutamine, luminal glutamine plus glutamate plus aspartate, and luminal
glucose contribute 38%, 39%, and 6% of the CO2 produced by the rat small intestine, respec-
tively (Windmueller, 1982). Similarly, the oxidation of enteral glutamate accounts for 36% of
the total CO2 production by the PDV of fully fed growing pigs, and is the most important single
contributor to mucosal oxidative ATP production (Stoll et al., 1999b). Approximately 80% and
65% of enterally delivered glutamate is oxidized to CO2 in the first pass by the human splanch-
nic bed (Battezzati et al., 1995) and by the porcine PDV (Reeds et al., 1996), respectively,
indicating that oxidation dominates intestinal glutamate metabolism. These results demonstrate
that amino acids, rather than glucose, are major fuels for the small intestinal mucosa.
Syntheses of purine and pyrimidine nucleotides and of glutathione represent physiologically
important pathways for intestinal utilization of glutamine and aspartate and of glutamate,
respectively (Burrin and Reeds, 1997; Wu, 1998b). Reeds et al. (1997) have shown that lumi-
nal glutamate, rather than the glutamate derived from intracellular glutamine degradation, is
the preferential source of glutamate for glutathione synthesis in the porcine intestinal mucosa,
suggesting that intestinal glutamate utilization is highly compartmentalized.
Enzymes for intestinal glutamine degradation have been identified, which include phosphate-
dependent glutaminase (PDG), carbamoylphosphate synthase II (glutamine) (CPS-II),
glutamate-oxaloacetate transaminase (GOT), glutamate-pyruvate transaminase (GPT), pyrroline-
5-carboxylate (P5C) synthase, ornithine aminotransferase (OAT), P5C reductase, ornithine
carbamoyltransferase (OCT), carbamoylphosphate synthase I (ammonia) (CPS-I), arginino-
succinate synthase (ASS), argininosuccinate lyase (ASL), ornithine decarboxylase (ODC), and
Krebs cycle enzymes (Wu and Morris, 1998; Bush et al., 2002; Morris, 2002). In animals, P5C
synthase is expressed primarily in enterocytes, indicating a unique role of the small intestine in
citrulline production and thus endogenous synthesis of arginine. The intestinal mucosa also con-
tains N-acetylglutamate (NAG) synthase, which synthesizes NAG (an allosteric activator of
CPS-I) from glutamate and acetyl-CoA (Wakabayashi et al., 1991). All of the glutamine-metabolic
enzymes, except for CPS II, P5C reductase, ASS, and ASL (cytosolic enzymes), are located in
mitochondria, whereas GPT and GOT are expressed in both the cytosol and mitochondria of the
intestinal mucosa.
Glutamine can be synthesized from glutamate and ammonia by glutamine synthetase in
avian and mammalian small intestines. For example, chick enterocytes are capable of syn-
thesizing glutamine in the presence of glutamate and ammonia (Porteous, 1980), which may
explain the net release of glutamine by the chick small intestine in vivo (Windmueller, 1982).
The intestinal synthesis of glutamine, coupled with a low rate of intestinal glutamine degra-
dation, helps explain a relatively high plasma concentration of free glutamine (1.1 mM) in
chicks (Wu et al., 1995a). In addition, a small amount of [14C]glutamine (1.2% of infused
[14C]glutamate) appeared in the sheep portal circulation when abomasum was infused with
[14C]glutamate (Tagari and Bergman, 1978). Both immunological and in situ hybridization
studies have shown that glutamine synthetase protein and mRNA are located primarily in the
intestinal crypt (Roig et al., 1995). Using the IEC-6 cell line (a well-characterized rat small
intestinal epithelial cell line), DeMarco et al. (1999) have shown that an inhibition of glutamine
synthetase reduces cell proliferation. This result suggests that when extracellular glutamine is
absent, the cytosolic synthesis of glutamine from glutamate and NH4+ may play a role in
endogenous provision of glutamine for supporting DNA and protein synthesis. However, the
Amino acid metablism in the small intestine 113

nutritional significance of such an observation is not entirely clear because the small intestine
constantly receives a supply of arterial glutamine. Moreover, the activity of glutamine syn-
thetase in the small intestine is generally very low, compared with PDG (Wu et al., 1994b,
1995b), and some of the glutamine synthesized in the cytosol likely enters mitochondria for
extensive catabolism. Thus, although there may be an intracellular (between the cytosol and
mitochondrion) or perhaps an intercellular glutamine–glutamate cycle (between crypt and
villus cells) in the small intestine, it is unlikely that net synthesis of glutamine in a nutrition-
ally significant quantity occurs in the mammalian gut in vivo.

2.4.2. Arginine, citrulline, ornithine, and proline

These four closely related amino acids are interconverted in the small intestinal mucosa.
Studies over the past 25 years have established that the production of citrulline by enterocytes
of the small intestine plays a crucial role in the endogenous synthesis of arginine (Wu and
Morris, 1998). Glutamine/glutamate and proline (abundant amino acids in milk) are the major
precursors for citrulline and arginine synthesis in incubated enterocytes from pigs (Wu et al.,
1994a; Wu, 1997), sheep, and cattle (Wu, unpublished data). In vivo studies have also demon-
strated citrulline and arginine synthesis from enteral proline and glutamate in pigs (Murphy
et al., 1996; Brunton et al., 1999) and the release of citrulline by the ruminant small intestine
(Bergman and Heitmann, 1978). The citrulline released by the small intestine is not taken up
by the liver, and is utilized for arginine synthesis primarily in kidneys (Dhanakoti et al.,
1990). Also, uptake of physiological concentrations of arginine by the liver is low due to a
low activity of the amino acid transport system y+ in hepatocytes (Wu and Morris, 1998).
Importantly, almost all extrahepatic cells are capable of synthesizing arginine from citrulline
(Wu and Morris, 1998). Thus, intestine-derived citrulline and arginine are equally effective as
sources of arginine for the whole body. The release of citrulline into the portal circulation by
the small intestine and the uptake of arterial citrulline by the kidneys for arginine production
is referred to as the intestinal–renal axis for endogenous synthesis of arginine.
Recent studies with pigs have demonstrated marked developmental changes in intestinal
arginine metabolism. First, in 1- to 7-day-old pigs, most of the citrulline synthesized from
glutamine and proline in enterocytes is converted locally into arginine because of high activ-
ities of both ASS and ASL (Wu and Knabe, 1995). However, in older piglets (14- to
21-day-old), enterocytes release most of the synthesized citrulline due to a low ASS activity
(Wu et al., 1994a). Thus, the small intestine shifts from the major site of net arginine synthe-
sis in 1-week-old pigs to the major site of net citrulline synthesis in 2- to 3-week-old pigs.
Second, intestinal synthesis of citrulline and arginine decreases by 60−75% in 7-day-old suck-
ling pigs compared with newborn pigs, and declines further in 14- to 21-day-old suckling pigs
(Wu, 1997). The metabolic basis for the marked decrease in citrulline and arginine synthesis by
enterocytes of 7- to 21-day-old pigs is not known. Because the ratios of small intestinal weight
(or mucosal protein weight) to body weight do not change substantially in newborn and suckling
piglets (28−32 g small intestine per kg body wt from 1 to 21 days of age), intestinal synthesis of
citrulline and arginine per kg body weight remains strikingly low in 7- to 21-day-old piglets com-
pared with 1- to 3-day-old piglets. Consequently, plasma concentrations of arginine and its
immediate precursors (ornithine and citrulline) decrease progressively by 20−41%, as the age of
sow-reared piglets increases from 3 to 14 days (Flynn et al., 2000). It should be borne in mind
that plasma arginine concentration is regulated by a number of factors, including arginine syn-
thesis and degradation, dietary arginine intake, degradation of plasma proteins and peptides, and
intracellular protein turnover (protein synthesis and degradation) (Young and El-Khoury, 1995).
114 G. Wu et al.

Thus, in suckling piglets, plasma arginine concentration would not be expected to decrease to the
same extent as the intestinal synthesis of citrulline and arginine.
In contrast to pigs, there are no postnatal decreases in plasma citrulline and arginine concen-
trations or in rates of intestinal citrulline and arginine synthesis from glutamine between 2- and
7-day-old suckling calves (Wu, unpublished data). For examples, rates of citrulline production
from 2 mM glutamine are 149 ±13 and 132 ± 16 nmol/mg DNA per 30 min (means ± SEM,
n = 5), respectively, in jejunal enterocytes of 2- and 7-day-old suckling calves. In addition, rates
of arginine synthesis from 2 mM glutamine are 185 ± 20 and 169 ± 22 nmol/mg DNA per 30 min
(means ± SEM, n = 5), respectively, in jejunal enterocytes of 2- and 7-day-old suckling calves.
The marked postnatal decline in intestinal citrulline and arginine synthesis represents an intrigu-
ing, and perhaps unique, aspect of amino acid metabolism in neonatal pigs.
Third, arginine catabolism in pig enterocytes is limited at birth and during the suckling
period due to a negligible arginase activity, but is markedly enhanced at weaning (Wu et al.,
1996a) due to the cortisol-mediated induction of arginase (Flynn and Wu, 1997a,b).
Consequently, urea is synthesized from extracellular and intramitochondrially generated
ammonia and from arginine in enterocytes of weaned pigs (Wu, 1995). This novel finding
challenges the textbook view that ureagenesis occurs only in the mammalian liver. The induc-
tion of arginase also makes possible the synthesis of proline and polyamines from arginine in
enterocytes of postweaning pigs (Wu, 1995; Wu et al., 2000a,b). Because of a relatively high
activity of arginase in the small intestinal mucosa of postweaning mammals, 40% of the
arginine absorbed by enterocytes from the intestinal lumen is degraded in a single pass in adult
rats (Windmueller and Spaeth, 1976). Similarly, in adult humans, 38% of dietary arginine is
removed in the first pass within the splanchnic region, and most of the arginine uptake is
accounted for by the small intestine (Castillo et al., 1993a). In sheep fed 16−20% crude protein
diets, 42−68% of arginine that disappears from the small intestinal lumen does not enter the
portal circulation (Tagari and Bergman, 1978).
In addition to ornithine production, there is nitric oxide (NO) synthesis from arginine by
NO synthase in enterocytes (Wu et al., 1996a) or the small intestine (Alican and Kubes,
1996). This pathway, however, is quantitatively low in enterocytes of newborn, suckling, and
weaned pigs (Wu et al., 1996a). Similarly, in healthy adult humans, only 0.34% of the dietary
arginine taken up in the first pass within the splanchnic region is utilized for NO production
(Castillo et al., 1993b). Despite the presence of arginine decarboxylase in a number of animal
tissues (e.g., brain, liver, and kidney) for agmatine synthesis from arginine, this enzyme is
absent from pig enterocytes (Wu et al., 1996a).
The findings that proline is actively catabolized by porcine enterocytes to produce ornithine,
citrulline, arginine, and polyamines (Wu, 1997; Wu et al., 2000a,b) are some of the most excit-
ing developments in intestinal amino acid metabolism in recent years. In pigs, the activities of
proline oxidase and OAT are greatest in the small intestinal mucosa (Wu, 1997). Proline oxidase
activity is also present in enterocyte mitochondria of rats (Wu, 1997), sheep, and cattle (Wu,
unpublished data). Considering the relatively large mass of the small intestine as compared with
the liver and kidneys, the intestinal mucosa likely plays a major role in initiating proline degra-
dation in the body. Consistent with this suggestion, Stoll et al. (1998) demonstrated that 38% of
dietary proline was extracted in the first pass by the PDV of milk protein-fed piglets. In sheep
fed 16−20% crude protein diets, 54−71% of proline that disappeared from the small intestinal
lumen did not enter the portal circulation (Tagari and Bergman, 1978).
By bridging the urea cycle with the Krebs cycle, arginine and proline metabolism,
and polyamine synthesis, OAT plays a central role in intestinal nitrogen and carbon
metabolism (fig. 1). We have recently suggested that the intestinal OAT reaction proceeds in
Fig. 1. Intestinal mucosal amino acid metabolism. Enzymes that catalyze the indicated reactions are: (1) phosphate-dependent glutaminase; (2) glutamine synthetase; (3) pyrroline-5-
carboxylate synthase; (4) ornithine aminotransferase; (5) ornithine carbamoyltransferase; (6) argininosuccinate synthase; (7) argininosuccinate lyase; (8) arginase; (9) ornithine decarboxy-
lase; (10) spermidine synthase; (11) spermine synthase; (12) S-adenosylmethionine synthase; (13) S-adenosylmethionine decarboxylase; (14) nitric oxide synthase; (15) N-acetylglutamate
synthase; (16) carbamoylphosphate synthase-I ; (17) proline oxidase; (18) pyrroline-5-carboxylate reductase; (19) branched-chain amino acid transaminase; (20) branched-chain α-ketoacid
dehydrogenase; (21) alanine transaminase; (22) aspartate transaminase; (23) purine- and pyrimidine-synthesizing enzymes; (24) aspartate decarboxylase; (25) histidine decarboxylase;
(26) α-ketoglutarate dehydrogenase; (27) asparaginase; (28) via aspartate transaminase and Krebs cycle enzymes; (29) possibly via NADP-linked malic enzyme, phosphoenolpyruvate
carboxykinase/pyruvate kinase, and oxaloacetate decarboxylase; (30) pyruvate dehydrogenase; (31) via Krebs cycle enzymes; (32) phenylalanine hydroxylase; (33) phenylalanine transaminase;
(34) lysine:α-ketoglutarate reductase; (35) serine transaminase; (36) serine dehydratase; (37) via enzymes of gluconeogenesis; (38) serine hydroxymethyltransferase; (39) glutathione-
synthesizing enzymes; and (40) glycine cleavage system. The symbol “?” denotes unknown reactions in intestinal mucosa. Abbreviations: Ala, alanine; Asp, aspartate; BCAA, branched-chain
amino acids; BCKA, branched-chain α-ketoacid; CoA, coenzyme A; CP, carbamoyl phosphate; DCAM, decarboxylated S-adenosylmethionine; α-KG, α-ketoglutarate; Met, methionine;
MTA, methylthioadenosine; MTF, N5,N10-methylenetetrahydrofolate; NAG, N-acetylglutamate; OAA, oxaloacetate; PPYR, phenylpyruvate; Pyr, pyruvate; TF, tetrahydrofolate.
116 G. Wu et al.

the direction of net synthesis of either ornithine or P5C, depending on intramitochondrial con-
centrations of ornithine and P5C (Wu, 1998b; Dekaney et al., 2000). For example, when there
is a high concentration of P5C in mitochondria due to the oxidation of large amounts of proline
by proline oxidase, the synthesis of ornithine and therefore of citrulline from P5C is favored in
enterocytes (Wu, 1997). On the other hand, when there is a high concentration of mitochondr-
ial ornithine due to the hydrolysis of large amounts of arginine by arginase II, the synthesis of
P5C and therefore of proline from ornithine is favored in enterocytes (Wu et al., 1996a).
Arginine-metabolic enzymes have been identified in the small intestine, and arginase is the
major enzyme initiating arginine catabolism in enterocytes (Wu et al., 1996a). Both enzymolog-
ical and metabolic evidence have established that P5C synthase is a key regulatory enzyme in
intestinal synthesis of citrulline from glutamine/glutamate (Wakabayashi et al., 1991; Wu and
Morris, 1998). By synthesizing NAG, NAG synthase may be another key regulatory enzyme in
intestinal synthesis of citrulline and arginine from glutamine/glutamate and proline. Given the
high activity of OCT in the small intestine, it is surprising that the major product of the degrada-
tion of extracellular ornithine in enterocytes is proline but not citrulline (Wu et al., 1996a). This
may be explained by the following reasons. First, enterocytes have an exceedingly high activity
of mitochondrial OAT, but a relatively low activity of mitochondrial CPS I for yielding carbamoyl
phosphate (a substrate for OCT) from NH3, HCO−3, and ATP. Thus, intramitochondrial ornithine
is preferentially utilized by OAT to form P5C instead of citrulline. Second, enterocytes have a
high activity of cytosolic P5C reductase. Therefore, when P5C enters the cytosol from mito-
chondria, P5C is readily converted into proline by P5C reductase. Thus, dietary or arterial blood
ornithine is a poor precursor for intestinal synthesis of citrulline, and does not contribute signif-
icantly to maintaining arginine homeostasis in animals (Wu and Morris, 1998).
There are marked species differences in intestinal citrulline and proline synthesis, both of which
are NADPH-dependent (Wu, 1996). For example, P5C synthase is absent and OAT activity is very
low in chick enterocytes (Wu et al., 1995a). This provides the metabolic basis for explaining the
absence of citrulline synthesis from glutamate and glutamine in the avian small intestine and thus
the nutritional essentiality of arginine for birds. Similarly, the near absence of P5C synthase in the
intestinal mucosa of the cat explains the lack of intestinal production of citrulline and thus little
endogenous synthesis of arginine in this species (Rogers and Phang, 1985). Strikingly, the rate of
conversion of arginine or ornithine into proline in chick enterocytes is only about 4% of that in pig
enterocytes owing to low activities of both arginase and OAT (Wu et al., 1995a). These data help
explain why ornithine is an ineffective replacement of proline in the chicken diet and why proline
is an essential amino acid for the chick (Graber and Baker, 1971, 1973).

2.4.3. Serine and glycine


Although the small intestine was not traditionally considered as a major organ for the catab-
olism of serine and glycine, Stoll et al. (1998) have recently reported that 40% and 50% of
dietary serine and glycine are extracted in the first pass by the PDV of the milk protein-fed
pig, respectively. Interestingly, <20% of the extracted serine and glycine are utilized for intes-
tinal protein synthesis, and the majority of the extracted amino acids are catabolized (Stoll
et al., 1998). In sheep fed a 20% crude protein diet, 28% of serine and 36% of glycine
that disappear from the small intestinal lumen do not enter the portal circulation (Tagari and
Bergman, 1978). These results suggest that dietary serine and glycine may be substantially
catabolized by the small intestine of both monogastric animals and ruminants.
There is some published evidence for the presence of key enzymes for intestinal metabo-
lism of serine and glycine. For example, the rat small intestine has been reported to contain
Amino acid metablism in the small intestine 117

the activities of serine dehydratase, serine aminotransferase, and serine hydroxymethyltrans-


ferase (Kikuchi et al., 1980). The latter interconverts serine into glycine and generates
N5, N10-methylenetetrahydrofolate for purine and pyrimidine synthesis. This reaction is
important for supporting the high rates of protein synthesis and epithelial cell proliferation in
the small intestinal mucosa. Glutathione synthesis also represents a physiologically important
pathway for glycine utilization by the small intestinal mucosa (Reeds et al., 1997). Furthermore,
the transsulfuration pathway occurs in the small intestinal mucosa (Luk et al., 1980), which can
contribute to serine metabolism. Although previous studies could not show the presence of the
glycine cleavage system in the rat small intestine (Kikuchi et al., 1980), likely due to the lack
of protease inhibitors in the enzyme assay system, we recently detected the activity of the
glycine cleavage system for the tetrahydrofolate-dependent production of ammonia and CO2
from glycine in enterocytes of rats, pigs, sheep, and cattle (Wu, unpublished data).

3. SIGNIFICANCE OF INTESTINAL AMINO ACID METABOLISM


3.1. Intestinal integrity and function

A characteristic of the small intestine physiology is high rates of intracellular protein turnover
and protein secretion by mucosal epithelial cells (Burrin and Reeds, 1997). Although protein
in the small intestine accounts for only 5% of whole-body protein, the amount of protein syn-
thesized by the small intestine daily contributes 15−20% of the whole-body protein synthesis
(McNurlan and Garlick, 1980; Reeds et al., 1993). This necessitates an adequate supply of
both amino acids and energy to the small intestinal mucosa. In support of this notion, the
available evidence shows that enteral feeding is the primary source of amino acids for the
intestinal mucosa because uptake of amino acids other than glutamine from arterial blood is
either low or insignificant (Windmueller, 1982; Wu et al., 1994a).
Amino acid metabolism is crucial for intestinal integrity and function through the following
mechanisms. First, dietary glutamine, glutamate and aspartate, and arterial blood glutamine are
major fuels for the small intestinal mucosa, and are responsible for providing energy required for
intestinal ATP-dependent metabolic processes, including active nutrient transport, intracellular
protein turnover, as well as epithelial cell proliferation and migration (Burrin and Reeds, 1997;
van der Schoor et al., 2001). On the basis of both experimental and clinical evidence, the impor-
tance of glutamine for supporting the metabolic function of intestinal mucosa has now generally
been accepted (Reeds and Burrin, 2001). Second, ornithine (a product of arginine, glutamine, and
proline metabolism) is the immediate precursor for the synthesis of polyamines in the enterocyte
(Wu et al., 2000a,b), which are essential to DNA and protein synthesis, as well as to the prolif-
eration, differentiation, and repair of intestinal epithelial cells (Luk et al., 1980; Wu, 1998b). In
addition, glutamine, asparagine, and glycine are potent stimulators of intestinal ODC (Kandil
et al., 1995), thereby enhancing polyamine synthesis from arginine- and proline-derived ornithine.
Third, arginine is the physiological precursor of NO, which plays an important role in regulating
intestinal blood flow, integrity, secretion, and epithelial cell migration (Alican and Kubes, 1996),
the relaxation of gastrointestinal smooth muscle cells, and feed intake (Morley and Flood, 1991).
Fourth, glutamate, glycine, and cysteine are precursors for the synthesis of glutathione, a tripep-
tide critical for defending the intestinal mucosa against toxic and oxidative damage (Reeds et al.,
1997). Because there is compartmentation of glutathione synthesis in enterocytes, and because
there is no significant uptake of arterial glutamate, glycine, and cysteine by the small intestine,
adequate dietary supply of these amino acids plays a vital role in glutathione synthesis by the
intestinal mucosa (Reeds et al., 2000a; Reeds and Burrin, 2001). Fifth, villus enterocytes derive
amino acids for protein synthesis and cell proliferation preferentially from the intestinal lumen
118 G. Wu et al.

rather than from the arterial blood (Alpers, 1972; Stoll et al., 1999a, 2000; Burrin et al., 2000).
This helps explain why parenteral nutrition selectively decreases protein synthesis in the small
intestinal mucosa and results in intestinal atrophy (Dudley et al., 1998; Stoll et al., 2000), and fur-
ther supports the view that enteral feeding of amino acids is obligatory for maintaining intestinal
mucosal mass and integrity (Wu, 1998a; Reeds et al., 2000a,b).

3.2. Endogenous synthesis of amino acids

3.2.1. Citrulline and arginine

Although arginine is formed via the urea cycle in the mammalian liver, there is no net syn-
thesis of arginine by this organ due to an exceedingly high activity of arginase in hepatocytes
(Wu and Morris, 1998). A unique, important aspect of intestinal amino acid metabolism is the
synthesis of citrulline and arginine from glutamate, glutamine, and proline in ruminant and
nonruminant animals (Bergman and Heitmann, 1978; Wu, 1998a). In both neonates and
adults, enterocytes are almost the exclusive source of citrulline for endogenous synthesis of
arginine. The near absence of arginase from neonatal enterocytes helps maximize the output
of arginine by the small intestine. This is of nutritional significance because arginine is
remarkably deficient in the milk of most mammals studied, including primates (human, chim-
panzee, gorilla, and rhesus), ruminants (cow, goat, sheep), and other nonprimates (elephant,
llama, pig, and rat) (Davis et al., 1994; Wu and Knabe, 1994; Reeds et al., 2000a), owing to
extensive catabolism of arginine by lactating mammary tissue (O’Quinn et al., 2002). For
example, concentrations of amino acids in sow’s whole milk (containing 18.6% dry matter)
on day 7 to day 28 of lactation are as follows (g/L): alanine, 1.89; arginine, 1.43; aspartate plus
asparagine, 5.02; cysteine, 0.72; glutamate plus glutamine, 9.42; glycine, 1.16; histidine, 0.92;
isoleucine, 2.12; leucine, 4.23; lysine, 3.87; methionine, 0.98; phenylalanine, 1.95; proline, 5.56;
serine, 2.30; threonine, 2.08; tryptophan, 0.64; tyrosine, 1.92; and valine, 2.40 (N = 10 sows;
Wu and Knabe, unpublished data). Note that the ratio of arginine/lysine in the sow’s milk (0.37)
is only 40% of that in the pig body (0.92) (Davis et al., 1993).
Arginine requirements by young mammals are particularly high (Rogers et al., 1970; Wu
et al., 2000a). The relative contribution of milk vs endogenous synthesis to meeting arginine
requirements by the suckling neonate can be estimated on the basis of arginine intake and
arginine accretion plus catabolism in the body. For example, for a 7-day-old pig (2.5 kg)
which gains 200 g body weight (27.2 g crude protein or 1.88 g arginine) per day (Flynn et al.,
2000), catabolizes 0.65 g arginine daily via arginase and NOS pathways (Murch et al., 1996),
and utilizes 0.17 g arginine daily for creatine synthesis [calculated on the basis of urinary cre-
atinine excretion (0.38 mmol/kg body wt/day)] (Weiler et al., 1997), the arginine requirement
is at least 2.7 g/day (table 2). On the basis of milk consumption by the suckling 7-day-old pig
(0.78 L milk/day), arginine content in sow’s milk (1.43 g/L of whole milk) (Wu and Knabe,
1994), and digestibility of arginine in milk protein (90.4%) (Mavromichalis et al., 2001), the
intake of sow’s milk provides only 1.01 g arginine/day, or at most 37% of the daily arginine
requirement (table 2). Thus, endogenous synthesis of arginine must provide at least 63% of
arginine for the suckling piglet.
The crucial role of the small intestine in endogenous arginine synthesis is further supported by
the following lines of direct evidence. First, an inhibition of intestinal synthesis of citrulline
retards the growth of young rats fed an arginine-deficient diet (Hoogenraad et al., 1985). Second,
resection of the rat small intestine (removal of 80% of the gut) results in (1) marked deficiencies
of citrulline and arginine in plasma and of arginine in skeletal muscle, (2) reduced animal
growth, (3) negative nitrogen balance, and (4) hypertension due to a deficiency of arginine for
Amino acid metablism in the small intestine 119

Table 2
Relative contributions of milk vs endogenous synthesis of arginine to arginine
requirements by the 7-day-old (2.5 kg) sow-reared piglet

Arginine requirements and sources Amounts of arginine (g/day)

Arginine requirements ≥2.7


Body weight gain (200 g/day; 27.2 g crude protein)a 1.88
Arginine catabolism via arginase and NO synthaseb 0.65
Creatine synthesis (0.38 mmol/kg body wt/day)c 0.17
Arginine supply from sow’s milk ≤1.01
Milk consumption (0.78 L/day; 1.43 g/L whole milk)d 1.12
Undigestible arginine in sow’s milk (9.6%)e −0.11
Arginine supply from endogenous synthesis ≥1.69
a Wu et al. (2000c), b Murch et al. (1996), c Weiler et al. (1997), d Wu and Knabe (1994), e Mavromichalis et al. (2001).

NO synthesis (Wakabayashi et al., 1994a,b). Third, hypoornithinemia, hypocitrullinemia,


hypoargininemia, and hyperammonemia occur in patients with short bowel syndrome and
chronic renal failure (Yokoyama et al., 1996), due to reduced intestinal production of citrulline
and impaired synthesis of arginine from citrulline in kidneys. Fourth, a deficiency of intestinal
OAT results in hypoornithinemia, hypocitrullinemia, hypoargininemia, hyperammonemia, and
even death in mice and humans during the neonatal period (Wang et al., 1995). Fifth, a recently
recognized deficiency of intestinal P5C synthase in humans causes hypoornithinemia, hypo-
citrullinemia, hypoargininemia, and hyperammonemia, retarded mental development, and death
(Kamoun et al., 1998). Sixth, in the sparse-fur mutant mouse, the inborn X-linked deficiency
of OCT limits intestinal citrulline synthesis, leads to impaired maturation of intestinal epithe-
lial cells (Malo et al., 1986), and causes retarded postnatal growth and death (DeMars et al.,
1976). Finally, an inhibition of intestinal citrulline synthesis for 12 h results in decreased
plasma concentrations of ornithine, citrulline, and arginine by 59%, 52%, and 76%, respec-
tively, in 4-day-old neonatal pigs nursed by sows (Flynn and Wu, 1996).
On the basis of decreases in plasma concentrations of arginine, ornithine, and citrulline as
well as nitrite and nitrate (stable end products of NO oxidation), and a concomitant increase
in plasma ammonia concentration, we have suggested that arginine is deficient in 7- to
21-day-old sow-reared piglets (Flynn et al., 2000). Although sow-reared piglets continue to
grow during the 21-day suckling period, this does not necessarily mean that arginine supply
from milk plus endogenous synthesis meets arginine requirements for maximal growth, as
exemplified by submaximal growth and impaired NO synthesis in arginine-deficient young
rats (Wu et al., 1999). Indeed, recent artificial rearing data show that the biological potential
for neonatal pig growth (from birth to day 21 of age) is at least 74% greater than that for sow-
reared piglets (Boyd et al., 1995). Both metabolic and growth data indicate that arginine
deficiency represents a major obstacle to maximal growth in milk-fed piglets (Kim et al.,
2004). In view of reduced arginine supply from endogenous synthesis in suckling compared
with newborn piglets and a great potential of arginine to enhance neonatal pig growth, it is of
crucial importance to identify an effective means for enhancing intestinal synthesis of
citrulline, thereby improving arginine nutrition and growth of sow-reared piglets.
Endogenous synthesis of arginine also plays an important role in maintaining arginine
homeostasis in postweaning growing animals. In 75-day-old pigs fed a conventional diet con-
taining 0.98% arginine (2.5 times the recommended National Research Council requirement for
dietary arginine), an inhibition of intestinal citrulline synthesis decreases plasma concentrations
120 G. Wu et al.

of citrulline and arginine by 26% and 22%, respectively (Wu et al., 1997). On the basis of
dietary arginine intake, ileal digestibility of dietary arginine, tissue protein accretion, and
oxidation of plasma arginine, we have estimated that endogenous synthesis of arginine provides
~50% of the total daily arginine requirement in the postweaning growing pig (Wu et al., 1997).

3.2.2. Proline

Proline is synthesized from arginine, ornithine, glutamine, and glutamate in enterocytes of


most mammals (Wu, 1998a). The finding that there is no synthesis of proline from intra-
venously infused glutamate in both pigs and humans (Matthews et al., 1993; Murphy et al.,
1996) not only suggests the complex compartmentation of intestinal glutamate metabolism and
proline synthesis, but also underscores the essential role for the small intestine in synthesizing
proline from enteral glutamate. The dietary essentiality of proline for birds results from (1) a
low rate of endogenous synthesis of proline from arginine in the small intestine because of a
low activity of intestinal OAT in birds, and (2) the lack of synthesis of proline from glutamate
or glutamine in the small intestine because of the absence of P5C synthase (Wu et al., 1995a).
The synthesis of proline from arginine and glutamine is low in enterocytes of suckling pigs
because of a negligible activity of arginase and P5C synthase, but is markedly increased in
cells from postweaning pigs owing to the induction of both enzymes (Wu et al., 1996a). This
may also explain, in part, (1) why proline is an essential amino acid for neonatal pigs (2.5 kg
of body weight) (Ball et al., 1986), and (2) why proline is a nonessential amino acid for post-
weaning pigs (5−15 kg of body weight) (Chung and Baker, 1993).

3.2.3. Alanine

In nonruminant animals, alanine is an important nitrogenous product of the intestinal catabolism


of glutamate, aspartate, and BCAA (Windmueller, 1982; Brosnan et al., 1983), and its carbon
skeleton (pyruvate) is derived partially from enteral glucose (Stoll et al., 1999b). Large amounts
of alanine are also released by the ruminant small intestine (Bergman and Heitmann, 1978).
Thus, alanine transaminase functions primarily for alanine synthesis in the small intestinal
mucosa of both monogastric animals and ruminants. Because alanine is a major amino acid for
hepatic gluconeogenesis, the intestine-derived alanine plays an important role in maintaining glu-
cose homeostasis. Alanine released by the small intestine also helps transport the nitrogen and
perhaps the carbons of dietary amino acids and arterial glutamine from the small intestine to the
liver. Because the liver actively takes up alanine and ammonia from portal and arterial blood and
releases glutamine, and because the small intestine substantially utilizes enteral and arterial blood
glutamine and releases large amounts of alanine and ammonia during both the postabsorptive
state and feeding, there appears to be a “glutamine–alanine cycle” involving the small intestine
and the liver. This cycle seems to be analogous to the glucose–lactate cycle (the Cori cycle) which
spans the skeletal muscle and liver. However, it should be pointed out that the carbons and nitro-
gen of the alanine released by the small intestine are utilized preferentially for the synthesis of
glucose and urea, respectively, rather than for glutamine synthesis, in the liver. Thus, the splanch-
nic “glutamine–alanine” cycle is indeed not a true metabolic cycle, but illustrates a key role of
alanine in the extensive recycling of nitrogen between the liver and the gut (Wu, 1998a).

3.3. Availability of dietary amino acids to extraintestinal tissues

A theme that has emerged from this review is that intestinal mucosal amino acid catabo-
lism plays an important role in regulating the availability of dietary amino acids to
Amino acid metablism in the small intestine 121

extraintestinal tissues. This novel concept has important implications for protein and amino
acid nutrition in animals. First, the extensive catabolism of dietary essential amino acids by
the small intestine results in a decrease in their nutritional efficiency. For example, the par-
ticularly high requirement of dietary arginine by the postweaning growing animal results, in
part, from the extensive hydrolysis of the absorbed dietary arginine by enterocytes (Wu et al.,
1997). Also, the extensive degradation of dietary glutamine limits the role for enteral gluta-
mine feeding to increase its plasma concentrations in many monogastric mammals, including
the rat and pig (Wu et al., 1996b; Wu, 1998b). In addition, infection with Trichostrongylus col-
ubriformis in 6- to 9-month-old lambs increases the catabolism of luminal leucine by the
gastrointestinal tract and reduces the availability of diet- and rumen-derived leucine for other
tissues, which provides a metabolic basis for the decreases in nitrogen retention and growth
rates under conditions of subclinical nematode infection (Yu et al., 2000). In support of the
concept of first-pass intestinal catabolism of essential amino acids, recent in vivo studies have
shown that methionine and threonine requirements are 35% and 45% higher, respectively,
during oral compared with parenteral feeding in neonatal pigs (Bertolo et al., 1998; Shoveller
et al., 2000). Most recently, Elango et al. (2002) reported that the parenteral requirement for
total BCAA is only 56% of the enteral requirement in neonatal pigs, indicating that 44% of
total BCAA is extracted by the first-pass metabolism in the gut.
There is a positive correlation between first-pass intestinal catabolism of dietary amino
acids and mucosal mass (Stoll et al., 1998). Thus, factors that affect intestinal mass (e.g.
antibiotics, growth hormone, insulin-like growth factor-I, and diabetes) may have an impor-
tant impact on the requirements of dietary amino acids (Wu, 1998b). For example, in pigs fed
antimicrobial agents (antibiotics and chemotherapeutics), a decrease in the small intestinal
mucosal mass is associated with an increase in whole-body growth rate (Yen et al., 1985;
Cromwell, 2001). This raises a possibility that reduced catabolism of dietary amino acids by the
small intestine is a mechanism responsible for the growth-promoting effect of antimicrobial
agents.
Second, intestinal amino acid metabolism modulates the entry of absorbed dietary amino
acids into portal circulation. Therefore, the pattern of amino acids in the diet differs remark-
ably from that in the intestinal tissue, portal circulation, and the extraintestinal organs (Le
Floc’h and Seve, 2000; Daenzer et al., 2001). Also, the pattern of amino acids in tissue pro-
teins or animal products (e.g. egg, milk, wool, and meat) is not necessarily similar to the ideal
pattern of dietary amino acid requirements in animals. In addition, there are profound differ-
ences in organ or plasma amino acid concentrations between enteral and parenteral feeding
in neonates, including infants and piglets (Bertolo et al., 2000; Wu et al., 2001).
Third, there are developmental changes, disease-associated alterations, and species differ-
ences in intestinal amino acid catabolism. Thus, these factors should be taken into
consideration in recommending dietary amino acid requirements and in refining in vivo
models of amino acid and protein metabolism. This is graphically demonstrated, for example,
by the dynamic changes in dietary requirements of arginine and proline by developmental pigs
(Wu, 1998b; Wu et al., 2000c), by increased requirements of arginine and proline for wound
healing in burned patients (Young and El-Khoury, 1995), and by the nutritional essentiality
of arginine and proline for birds (Wu and Morris, 1998).

4. FUTURE PERSPECTIVES
Much has been learned in recent years regarding intestinal amino acid metabolism. However,
there are many important and yet challenging questions in this rapidly growing and fruitful
122 G. Wu et al.

area of investigation. First, recent intriguing findings of the extensive first-pass extraction of
dietary essential amino acids and cysteine by the porcine PDV and human splanchnic bed raise
an important question regarding their catabolism in the small intestinal mucosa. This novel
concept should be firmly established by biochemical studies with isolated enterocytes. In addi-
tion, microorganisms in the intestinal lumen may substantially contribute to the catabolism of
enteral essential amino acids and cysteine and such microbial pathways should be quantified.
Second, in view of the recently recognized deficiency of arginine (an essential amino acid for
neonates) in sow-reared piglets (Flynn et al., 2000), further studies are necessary to elucidate
the mechanisms responsible for the marked decline in intestinal synthesis of citrulline and argi-
nine in suckling piglets. This new knowledge will undoubtedly help design new, effective
means to enhance arginine supply to the piglets and therefore improve their postnatal growth.
Third, much work is required to define hormonal and nutritional regulation of intestinal amino
acid metabolism at molecular, cellular, and whole-body levels. This will be facilitated by the
recent availability of porcine small intestinal epithelial cells (Lu et al., 2002) and of mam-
malian cDNAs for key regulatory enzymes [e.g. arginase II (Morris, 2002), P5C synthase (Aral
et al., 1996), and NAG synthase (Caldovic et al., 2002)], and by recent biochemical and molec-
ular biology techniques (e.g. proteomics, metabolomics, and microarrays) (Phelps et al., 2002).
We predict that exciting new knowledge on the regulation of intestinal amino acid metabolism
will be discovered in the coming years, which will help design new means to improve the effi-
ciency of protein and amino acid utilization by animals, including humans.

REFERENCES
Alican, I., Kubes, P., 1996. A critical role for nitric oxide in intestinal barrier function and dysfunction.
Amer. J. Physiol. 270, G225−G237.
Alpers, D.H., 1972. Protein synthesis in intestinal mucosal. J. Clin. Invest. 51, 167−173.
Aral, B., Schlenzig, J.-S., Liu, G., Kamoun, P., 1996. Database cloning human Δ1-pyrroline-5-carboxylate
synthetase (P5CS) cDNA: a bifunctional enzyme catalyzing the first 2 steps in proline biosynthesis.
C.R. Acad. Sci. Paris 319, 171−178.
Ball, R.O., Atkinson, J.L., Bayley, S.H., 1986. Proline as an essential amino acid for the young pig. Brit.
J. Nutr. 55, 659−668.
Battezzati, A., Brillon, D.J., Matthews, D.E., 1995. Oxidation of glutamic acid by the splanchnic bed in
humans. Amer. J. Physiol. 269, E269−E276.
Bergman, E.N., Heitmann, R.N., 1978. Metabolism of amino acids by the gut, liver, kidneys, and periph-
eral tissues. Fed. Proc. 37, 1228−1232.
Bertolo, R.F.P., Chen, C.Z.L., Law, G., Pencharz, P.B., Ball, R.O., 1998. Threonine requirement of neonatal
piglets receiving total parenteral nutrition is considerably lower than that of piglets receiving an iden-
tical diet intragastrically. J. Nutr. 128, 1752−1759.
Bertolo, R.F.P., Pencharz, P.B., Ball, R.O., 2000. Organ and plasma amino acid concentrations are
profoundly different in piglets fed identical diets via gastric, central venous or portal venous route.
J. Nutr. 130, 1261−1266.
Biolo, G., Tessari, P., Inchiostro, S., Bruttomesso, D., Fongher, C., Sabadin, L., Fratton, M.G., Valerio, A.,
Tiengo, A., 1992. Leucine and phenylalanine kinetics during mixed meal ingestion: a multiple tracer
approach. Amer. J. Physiol. 262, E455−E463.
Boyd, R.D., Kensinger, R.S., Harrell, R.J., Bauman, D.E., 1995. Nutrient uptake and endocrine regula-
tion of milk synthesis by mammary tissue of lactating sows. J. Anim. Sci. 73, Suppl. 2, 36−56.
Brosnan, J.T., Man, K.C., Hall, D.E., Colbourne, S.A., Brosnan, M.E., 1983. Interorgan metabolism of
amino acids in streptozotocin-diabetic ketoacidotic rat. Amer. J. Physiol. 244, E151−E158.
Brunton, J.A., Bertolo, R.F.P., Pencharz, P.B., Ball, R.O., 1999. Proline ameliorates arginine deficiency
during enteral but not parenteral feeding in neonatal pigs. Amer. J. Physiol. 277, E223−E231.
Burrin, D.G., Reeds, P.J., 1997. Alternative fuels in the gastrointestinal tract. Curr. Opin. Gastroenterol.
13, 165−170.
Amino acid metablism in the small intestine 123

Burrin, D.G., Stoll, B., Jiang, R.H., Chang, X.Y., Hartmann, B., Holst, J.J., Greeley, G.H., Reeds, P.J.,
2000. Minimal enteral nutrient requirements for intestinal growth in neonatal piglets: how much is
enough? Amer. J. Clin. Nutr. 71, 1603−1610.
Bush, J.A., Wu, G., Suryawan, A., Nguyen, H.V., Davis, T.A., 2002. Somatotropin-induced amino acid
conservation in pigs involves differential regulation of liver and gut urea cycle enzyme activity.
J. Nutr. 132, 59−67.
Caldovic, L., Morizono, H., Yu, X., Thompson, M., Shi, D., Gallegos, R., Allewell, N.M., Malamy, M.H.,
Tuchman, M., 2002. Identification, cloning and expression of the mouse N-acetylglutamate synthase
gene. Biochem. J. 364, 825−831.
Cappelli, F.P., Seal, C.J., Parker, D.S., 1997. Glucose and [13C]leucine metabolism by the portal-drained
viscera of sheep fed on dried grass with acute intravenous and intraduodenal infusion of glucose.
Brit. J. Nutr. 78, 931−946.
Castillo, L., Chapman, T.E., Yu, Y.M., Ajami, A., Burke, J.F., Young, V.R., 1993a. Dietary arginine uptake
by the splanchnic region in adult humans. Amer. J. Physiol. 265, E532−E539.
Castillo, L., DeRojas, T.C., Chapman, T.E., Vogt, J., Burke, J.F., Tannenbaum, S.R., Young, V.R., 1993b.
Splanchnic metabolism of dietary arginine in relation to nitric oxide synthesis in normal adult man.
Proc. Natl. Acad. Sci. USA 90, 193−197.
Cheng, H., Leblond, C.P., 1974. Origin, differentiation and renewal of the four main epithelial cell types
in the mouse small intestine. V. Unitarian theory of the origin of the four epithelial cell types.
Amer. J. Anat. 141, 537−548.
Chung, T.K., Baker, D.H., 1993. A note on the dispensability of proline for weanling pigs. Anim. Prod.
56, 407−408.
Cooper, A.J., Meister, A., 1977. The glutamine transaminase–ω-amidase pathway. CRC Crit. Rev.
Biochem. 4, 281−303.
Cromwell, G.L., 2001. Antimicrobial and promicrobial agents. In: Lewis, A.J., Southern, L.L. (Eds.),
Swine Nutrition. CRC Press, New York, pp. 401−426.
Daenzer, M., Petzke, K.J., Bequette, B.J., Metges, C.C., 2001. Whole-body nitrogen and splanchnic
amino acid metabolism differ in rats fed mixed diets containing casein or its corresponding amino
acid mixture. J. Nutr. 131, 1965−1972.
Davis, T.A., Fiorotto, M.L., Reeds, P.J., 1993. Amino acid compositions of body and milk protein change
during the suckling period in rats. J. Nutr. 123, 947−956.
Davis, T.A., Nguyen, H.V., Garcia-Bravo, R., Fiorotto, M.L., Jackson, E.M., Lewis, D.S., Lee, D.R.,
Reeds, P.J., 1994. Amino acid composition of human milk is not unique. J. Nutr. 124, 1126−1132.
Dekaney, C.M., Wu, G., Jaeger, L.A., 2000. Regulation and function of ornithine aminotransferase in
animals. Trends Comp. Biochem. Physiol. 6, 175−183.
DeMarco, V., Dyess, K., Strauss, D., West, C.M., Neu, J., 1999. Inhibition of glutamine synthetase
decreases proliferation of cultured rat intestinal epithelial cells. J. Nutr. 129, 57−62.
DeMars, R., LeVan, S.L., Trend, B.L., Russell, L.B., 1976. Abnormal ornithine carbamoyltransferase in
mice having the sparse-fur mutation. Proc. Natl. Acad. Sci. USA 73, 1693−1697.
Dhanakoti, S.N., Brosnan, J.T., Herzberg, G.R., Brosnan, M.E., 1990. Renal arginine synthesis: studies
in vitro and in vivo. Amer. J. Physiol. 259, E437−E442.
Dudley, M.A., Wykes, L.J., Dudley, A.W. Jr., Burrin, D.G., Nichols, B.L., Rosenberger, J., Jahoor, F.,
Heird, W.C., Reeds, P.J., 1998. Parenteral nutrition selectively decreases protein synthesis in the small
intestine. Amer. J. Physiol. 274, G131−G137.
Elango, R., Pencharz, P.B., Ball, R.O., 2002. The branched-chain amino acid requirement of parenterally
fed neonatal piglets is less than the enteral requirement. J. Nutr. 132, 3123−3129.
Flynn, N.E., Wu, G., 1996. An important role for endogenous synthesis of arginine in maintaining
arginine homeostasis in neonatal pigs. Amer. J. Physiol. 271, R1149−R1155.
Flynn, N.E., Wu, G., 1997a. Glucocorticoids play an important role in mediating the enhanced meta-
bolism of arginine and glutamine in enterocytes of postweaning pigs. J. Nutr. 127, 732−737.
Flynn, N.E., Wu, G., 1997b. Enhanced metabolism of arginine and glutamine in enterocytes of cortisol-
treated pigs. Amer. J. Physiol. 272, G474−G480.
Flynn, N.E., Knabe, D.A., Mallick, B.K., Wu, G., 2000. Postnatal changes of plasma amino acids in suckling
pigs. J. Anim. Sci. 78, 2369−2375.
Gate, J.J., Parker, D.S., Lobley, G.E., 1999. The metabolic fate of the amino-N group of glutamine in the
tissues of the gastrointestinal tract in 24 h-fasted sheep. Brit. J. Nutr. 81, 297−306.
Graber, G., Baker, D.H., 1971. Ornithine utilization by the chick. Proc. Soc. Exp. Biol. Med. 138, 585−588.
124 G. Wu et al.

Graber, G., Baker, D.H., 1973. The essential nature of glycine and proline for growing chickens. Poultry
Sci. 52, 892−896.
Harper, A.E., Miller, R.H., Block, K.P., 1984. Branched-chain amino acid metabolism. Annu. Rev. Nutr.
4, 409−454.
Hoerr, R.A., Matthews, D.E., Bier, D.M., Young, V.R., 1993. Effects of protein restriction and acute
refeeding on leucine and lysine kinetics in young men. Amer. J. Physiol. 264, E567−E575.
Hoogenraad, N., Totino, N., Elmer, H., Wraight, C., Alewood, P., Johns, R.B., 1985. Inhibition of intes-
tinal citrulline synthesis causes growth retardation in rats. Amer. J. Physiol. 249, G792−G799.
Kamoun, P., Aral, B., Saudubray, J.M., 1998. A new inherited metabolic disease: pyrroline-5-carboxylate
synthetase deficiency. Bull. Acad. Natle. Med. Paris 182, 131−139.
Kandil, H.M., Argenzio, R.A., Chen, W., Berschneider, H.M., Stiles, A.D., Westwick, J.K., Rippe, R.A.,
Brenner, D.A., Rhoads, J.M., 1995. L-Glutamine and L-asparagine stimulate ODC activity and
proliferation in a porcine jejunal enterocytes line. Amer. J. Physiol. 269, G591−G599.
Khatra, B.S., Chawla, R.K., Sewell, C.W., Rudman, D., 1977. Distribution of branched-chain α-keto acid
dehydrogenase in primate tissues. J. Clin. Invest. 59, 558−564.
Kikuchi, G., Hiraga, K., Yoshida, T., 1980. Role of the glycine-cleavage system in glycine and serine
metabolism in various organs. Biochem. Soc. Trans. 8, 504−506.
Kim, S.W., McPherson, R.L., Wu, G., 2004. Dietary arginine supplementation enhances the growth of
milk-fed young pigs. J. Nutr. 134, 625–630.
Klein, R.M., McKenzie, J.C., 1983. The role of cell renewal in the ontogeny of the intestine. I. Cell
proliferation patterns in adult, fetal, and neonatal intestine. J. Pediat. Gastroenterol. Nutr. 2, 10−43.
Lapierre, H., Bernier, J.F., Dubreuil, P., Reynolds, C.K., Farmer, C., Ouellet, D.R., Lobley, G.E., 2000. The
effect of feed intake level on splanchnic metabolism in growing beef steers. J. Anim. Sci. 78, 1084−1099.
Le Floc’h, N., Seve, B., 2000. Protein and amino acid metabolism in the intestine of the pig: from digestion
to appearance in the portal vein. Prod. Anim. 13, 303−314.
Lu, S., Yao, Y., Meng, S., Cheng, X., Black, D.D., 2002. Overexpression of apolipoprotein A-IV enhances
lipid transport in newborn swine intestinal epithelial cells. J. Biol. Chem. 277, 31929−31937.
Luk, G.D., Marton, L.J., Baylin, S.B., 1980. Ornithine decarboxylase is important in intestinal mucosal
maturation and recovery from injury in rats. Science 210, 195−198.
Madara, J.L., 1991. Functional morphology of epithelium of the small intestine. In: Shultz, S.G. (Ed.),
Handbook of Physiology: The Gastrointestinal System. American Physiological Society, Bethesda,
MD, pp. 83−120.
Malo, C., Qureshi, I.A., Letarte, J., 1986. Postnatal maturation of enterocytes in sparse-fur mutant mice.
Amer. J. Physiol. 250, G177−G184.
Matthews, D.E., Marano, M.A., Campbell, R.G., 1993. Splanchnic bed utilization of glutamine and
glutamic acid in humans. Amer. J. Physiol. 264, E848−E854.
Mavromichalis, I., Parr, T.M., Gabert, V.M., Baker, D.H., 2001. True ileal digestibility of amino acids in
sow’s milk for 17-day-old pigs. J. Anim. Sci. 79, 707−713.
McNurlan, M.A., Garlick, P.J., 1980. Contribution of rat liver and gastrointestinal tract to whole body
protein synthesis in the rat. Biochem. J. 186, 381−383.
Mitchell, A.D., Benevenga, N.J., 1978. The role of transamination in methionine oxidation in the rat.
J. Nutr. 108, 67−78.
Morley, J.E., Flood, J.F., 1991. Evidence that nitric oxide modulates food intake in mice. Life Sci. 49, 707−711.
Morris, S.M. Jr., 2002. Regulation of enzymes of the urea cycle and arginine metabolism. Annu. Rev.
Nutr. 22, 87−105.
Mowat, A.M., 1987. The cellular basis of gastrointestinal immunity. In: Marsh, M.N. (Ed.),
Immunopathology of the Small Intestine. John Wiley, London, pp. 41−72.
Murch, S.J., Wilson, R.L., Murphy, J.M., Ball, R.O., 1996. Proline is synthesized from intravenously
infused arginine by piglets consuming low proline diets. Can. J. Anim. Sci. 76, 435−441.
Murphy, J.M., Murch, S.J., Ball, R.O., 1996. Proline is synthesized from glutamate during intragastric
infusion but not during intravenous infusion in neonatal piglets. J. Nutr. 126, 878−886.
O’Quinn, P.R., Knabe, D.A., Wu, G., 2002. Arginine catabolism in lactating porcine mammary tissue.
J. Anim. Sci. 80, 467−474.
Pell, J.M., Caldarone, E.M., Bergman, E.N., 1986. Leucine and α-ketoisocaproate metabolism and
interconversion in fed and fasted sheep. Metabolism 35, 1005−1016.
Phelps, T.J., Palumbo, A.V., Beliaev, A.S., 2002. Metabolomics and microarrays for improved under-
standing of phenotypic characteristics controlled by both genomics and environmental constraints.
Curr. Opin. Biotechnol. 13, 20−24.
Amino acid metablism in the small intestine 125

Pink, D.B.S., Dixon, W.T., Ball, R.O., 2002. Lysine catabolism in swine: an enzymatic approach. FASEB J.
16, A258.
Porteous, J.W., 1980. Glutamate, glutamine, aspartate, asparagines, glucose and ketone-body metabolism
in chick intestinal brush-border cells. Biochem. J. 188, 619−632.
Reeds, P.J., Burrin, D.G., 2001. Glutamine and the bowel. J. Nutr. 131, 2505S−2508S.
Reeds, P.J., Burrin, D.G., Davis, T.A., Fiorotto, M.L., 1993. Postnatal growth of gut and muscle: com-
petitors or collaborators. Proc. Nutr. Soc. 52, 57−67.
Reeds, P.J., Burrin, D.G., Davis, T.A., Fiorotto, M.L., Stoll, B., van Goudoever, J.B., 2000a. Protein nutrition
in the neonate. Proc. Nutr. Soc. 59, 87−97.
Reeds, P.J., Burrin, D.G., Jahoor, F., Wykes, L., Henry, J., Frazer, E.M., 1996. Enteral glutamate is almost
completely metabolized in first pass by the gastrointestinal tract of infant pigs. Amer. J. Physiol. 270,
E413−E418.
Reeds, P.J., Burrin, D.G., Stoll, B., Jahoor, F., 2000b. Intestinal glutamate metabolism. J. Nutr. 130, 978S−982S.
Reeds, P.J., Burrin, D.G., Stoll, B., Jahoor, F., Wykes, L., Henry, J., Frazer, M.E., 1997. Enteral glutamate is
the preferential source for mucosal glutathione synthesis in pigs. Amer. J. Physiol. 273, E408−E415.
Roig, J.C., Shenoy, V.B., Chakrabarti, R., Lau, J.Y.N., Neu, J., 1995. Localization of rat small intestine gluta-
mine synthetase using immunofluorescence and in situ hybridization. J. Parent. Enter. Nutr. 19, 179−181.
Rogers, Q.R., Chen, D.M., Harper, A.E., 1970. The importance of dispensable amino acids for maximal
growth in rats. Proc. Soc. Exp. Biol. Med. 134, 517−522.
Rogers, Q.R., Phang, J.M., 1985. Deficiency of pyrroline-5-carboxylate synthase in the intestinal mucosa
of the cat. J. Nutr. 115, 146−150.
Shoveller, A.K., Ball, R.O., Pencharz, P.B., 2000. Methionine requirement is 35% lower during parenteral
versus oral feeding in neonatal piglets. FASEB J. 14, A558.
Shrawder, E., Martinez-Carrion, M., 1972. Evidence of phenylalanine transaminase activity in the isoen-
zymes of aspartate transaminase. J. Biol. Chem. 247, 2486−2492.
Smith, M.W., Jarvis, L.G., 1978. Growth and cell replacement in the new-born pig intestine. Proc. R. Soc.
Lond. B 203, 69−89.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1997. Phenylalanine utilization by the gut and
liver measured with intravenous and intragastric tracers in pigs. Amer. J. Physiol. 273, G1208−G1217.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1999a. Utilization of enteral and arterial pheny-
lalanine for mucosal and hepatic constitutive protein synthesis in pigs. Amer. J. Physiol. 276, G49−G57.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1999b. Substrate oxidation by the portal
drained viscera of fed pigs. Amer. J. Physiol. 277, E168−E175.
Stoll, B., Chang, X.Y., Fan, M.Z., Reeds, P.J., Burrin, D.G., 2000. Enteral nutrient intake level determines
intestinal protein synthesis and accretion rates in neonatal pigs. Amer. J. Physiol. 279, G288−G294.
Stoll, B., Henry, J., Reeds, P.J., Yu, H., Jahoor, F., Burrin, D.G., 1998. Catabolism dominates the first-pass
intestinal metabolism of dietary essential amino acids in milk-protein-fed piglets. J. Nutr. 128, 606−614.
Tagari, H., Bergman, E.N., 1978. Intestinal disappearance and portal blood appearance of amino acids in
sheep. J. Nutr. 108, 790−803.
Tourian, A., Goddard, J., Puck, T.T., 1969. Phenylalanine hydroxylase activity in mammalian cells.
J. Cell. Physiol. 73, 59−170.
van der Schoor, S.R.D., van Goudoever, J.B., Stoll, B., Henry, J.F., Rosenberger, J.R., Burrin, D.G.,
Reeds, P.J., 2001. The pattern of intestinal substrate oxidation is altered by protein restriction in pigs.
Gastroenterology 121, 1167−1175.
van Goudoever, J.B., Stoll, B., Henry, J.F., Burrin, D.G., Reeds, P.J., 2000. Adaptive regulation of intes-
tinal lysine metabolism. Proc. Natl Acad. Sci. USA 97, 11620−11625.
Wakabayashi, Y., Yamada, E., Hasegawa, T., Yamada, R., 1991. Enzymological evidence for the indispens-
ability of small intestine in the synthesis of arginine from glutamate. Arch. Biochem. Biophys. 291, 1−8.
Wakabayashi, Y., Yamada, E., Yoshida, T., Takahashi, H., 1994a. Arginine becomes an essential amino
acid after massive resection of rat small intestine. J. Biol. Chem. 269, 32667−32671.
Wakabayashi, Y., Yamada, E., Yoshida, T., Takahashi, H., 1994b. Deficiency of endogenous arginine
synthesis provokes hypertension by exhausting substrate arginine for nitric oxide synthesis. Biochem.
Biophys. Res. Commun. 205, 1391−1398.
Wang, T., Lawler, A.M., Steel, G., Sipik, I., Milam, A.H., Valle, D., 1995. Mice lacking ornithine-δ-
aminotransferase have paradoxical neonatal hypo-ornithinaemia and retinal degeneration. Nature
Genet. 11, 185−190.
Weiler, H.A., Wang, Z., Atkinson, S.A., 1997. Whole body lean mass is altered by dexamethasone
treatment through reduction in protein and energy utilization in piglets. Biol. Neonate 71, 53−59.
126 G. Wu et al.

Windmueller, H.G., 1982. Glutamine utilization by the small intestine. Adv. Enzymol. 53, 201−237.
Windmueller, H.G., Spaeth, A.E., 1975. Intestinal metabolism of glutamine and glutamate from the
lumen as compared to glutamine from blood. Arch. Biochem. Biophys. 171, 662−672.
Windmueller, H.G., Spaeth, A.E., 1976. Metabolism of absorbed aspartate, asparagine, and arginine by
rat small intestine in vivo. Arch. Biochem. Biophys. 175, 670−676.
Wu, G., 1995. Urea synthesis in enterocytes of developing pigs. Biochem. J. 312, 717−723.
Wu, G., 1996. An important role for pentose cycle in the synthesis of citrulline and proline from glutamine
in porcine enterocytes. Arch. Biochem. Biophys. 336, 224−230.
Wu, G., 1997. Synthesis of citrulline and arginine from proline in enterocytes of postnatal pigs. Amer.
J. Physiol. 272, G1382−G1390.
Wu, G., 1998a. Intestinal mucosal amino acid catabolism. J. Nutr. 128, 1249−1252.
Wu, G., 1998b. Amino acid metabolism in the small intestine. Trends Comp. Biochem. Physiol. 4, 39−74.
Wu, G., Becker, R.M., Bose, C.L., Rhoads, J.M., 2001. Serum amino acid concentrations in preterm
infants. J. Pediat. 139, 334−337.
Wu, G., Borbolla, A.G., Knabe, D.A., 1994a. The uptake of glutamine and release of arginine, citrulline
and proline by the small intestine of developing pigs. J. Nutr. 124, 2437−2444.
Wu, G., Davis, P.K., Flynn, N.E., Knabe, D.A., Davidson, J.T., 1997. Endogenous synthesis of arginine
plays an important role in maintaining arginine homeostasis in postweaning growing pigs. J. Nutr.
127, 2342−2349.
Wu, G., Flynn, N.E., Flynn, S.P., Jolly, C.A., Davis, P.K., 1999. Dietary protein or arginine deficiency
impairs constitutive and inducible nitric oxide synthesis in young rats. J. Nutr. 129, 1347−1354.
Wu, G., Flynn, N.E., Knabe, D.A., 2000a. Enhanced intestinal synthesis of polyamines from proline in
cortisol-treated pigs. Amer. J. Physiol. 279, E395−E402.
Wu, G., Flynn, N.E., Knabe, D.A., Jaeger, L.A., 2000b. A cortisol surge mediates the enhanced
polyamine synthesis in porcine enterocytes during weaning. Amer. J. Physiol. 279, R554−R559.
Wu, G., Flynn, N.E., Yan, W., Barstow, D.G. Jr., 1995a. Glutamine metabolism in chick enterocytes:
absence of pyrroline-5-carboxylate synthase and citrulline synthesis. Biochem. J. 306, 717−721.
Wu, G., Knabe, D.A., 1994. Free and protein-bound amino acids in sow’s colostrums and milk. J. Nutr.
124, 415−424.
Wu, G., Knabe, D.A., 1995. Arginine synthesis in enterocytes of neonatal pigs. Amer. J. Physiol. 269,
R621−R629.
Wu, G., Knabe, D.A., Flynn, N.E., 1994b. Synthesis of citrulline from glutamine in pig enterocytes.
Biochem. J. 299, 115−121.
Wu, G., Knabe, D.A., Flynn, N.E., Yan, W., Flynn, S.P., 1996a. Arginine degradation in developing
porcine enterocytes. Amer. J. Physiol. 271, G913−G919.
Wu, G., Knabe, D.A., Yan, W., Flynn, N.E., 1995b. Glutamine and glucose metabolism in enterocytes of
the neonatal pig. Amer. J. Physiol. 268, R334−R342.
Wu, G., Meier, S.A., Knabe, D.A., 1996b. Dietary glutamine supplementation prevents jejunal atrophy in
weaned pigs. J. Nutr. 126, 2578−2584.
Wu, G., Meininger, C.J., Knabe, D.A., Bazer, F.W., Rhoads, J.M., 2000c. Arginine nutrition in develop-
ment, health and disease. Curr. Opin. Clin. Nutr. Metab. Care 3, 59−66.
Wu, G., Morris, S.M. Jr., 1998. Arginine metabolism: nitric oxide and beyond. Biochem. J. 336, 1−17.
Yen, J.T., Nienaber, J.A., Pond, W.G., Varel, V.H., 1985. Effect of carbadox on growth, fasting metabo-
lism, thyroid function and gastrointestinal tract in young pigs. J. Nutr. 115, 970−979.
Yokoyama, K., Ogura, Y., Kawabata, M., Hinoshita, F., Susuki, Y., Hara, S., Yamada, A., Mimura, N.,
Nakayama, M., Kawaguchi, Y., Sakai, O., 1996. Hyperammonemia in a patient with short bowel
syndrome and chronic renal failure. Nephron 72, 693−695.
Young, V.R., El-Khoury, A.E., 1995. The notion of the nutritional essentiality of amino acids, revisited,
with a note on the indispensable amino acid requirements in adults. In: Cynober, L.A. (Ed.), Amino
Acid Metabolism and Therapy in Health and Nutritional Disease, CRC Press, New York, pp. 191−232.
Yu, F., Bruce, L.A., Calder, A.G., Milne, E., Coop, R.L., Jackson, F., Horgan, G.W., MacRae, J.C., 2000.
Subclinical infection with the nematode Trichostrongylus colubriformis increases gastrointestinal tract
leucine metabolism and reduces availability of leucine for other tissues. J. Anim. Sci. 78, 380−390.
Yu, Y.-M., Wagner, D.A., Tredget, E.E., Walaszewski, J.A., Burke, J.F., Young, V.R., 1990. Quantitative
role of splanchnic region in leucine metabolism: L-[1-13C,15N]leucine and substrate balance studies.
Amer. J. Physiol. 259, E36−E51.
6 Role of intestinal first-pass metabolism
on whole-body amino acid requirements

R. F. P. Bertoloa, P. B. Pencharzc,d and R. O. Ballb,d

a Department of Biochemistry, Memorial University of Newfoundland,


St. John’s, NewfoundLand, Canada A1B 3X9
b Department of Agricultural, Food and Nutritional Science, University of Alberta,

Edmonton, Alberta, Canada T6G 2P5


c Department of Paediatrics, University of Toronto, Toronto, Ontario,

Canada M5G 1X8


d The Research Institute, The Hospital for Sick Children, Toronto,

Department of Nutritional Sciences, University of Toronto,


Toronto, Ontario, Canada

The small intestine utilizes a different profile of amino acids compared to whole-body
requirements. Quantifying the gut requirements for amino acids is critical to understand the
limiting availability of these amino acids during periods of rapid growth in animals. Many
methods have been employed to determine amino acid requirements in man and animals
including growth assays, nitrogen balance and amino acid oxidation methods. The most versa-
tile approach is the indicator amino acid oxidation technique which can be safely employed
in many vulnerable populations. The amino acid requirements of the gut have been estimated
using this technique in the parenterally fed piglet, which is a model of a gut-deficient animal,
and comparing requirements to enterally fed controls. The gut’s requirement for threonine is
proportionately greatest of the amino acids tested due to its role in mucin synthesis. The sulphur
and branched-chain amino acids are also significantly utilized by the gut. Tryptophan, lysine,
phenylalanine and tyrosine utilization by the gut is not significant. The availability of threonine
and sulphur amino acids may be limiting for growth in situations of gut stress or disease due to
the higher maintenance requirements during such gut challenges.

1. INTRODUCTION
The small intestine has classically been regarded as a digestive organ responsible for the
absorption of nutrients from foods. Only recently has the gastrointestinal tract, whose metab-
olism is dominated by the small intestine, been studied as a significant metabolic tissue with

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
127 © 2005 Elsevier Limited. All rights reserved.
128 R. F. P. Bertolo et al.

great impact on whole-body metabolism. Through co-ordinated inter-organ pathways, the gas-
trointestinal tract is involved in the synthesis, conversion and catabolism of amino acids to be
used by other tissues in the body. In addition to this critical role in whole-body nutriture, the gut
also requires vast amounts of particular amino acids for maintenance and growth. The profile of
these amino acid requirements does not seem to parallel those for growth and maintenance of
the rest of the body. Rather, because of the gut’s specific functions in digestion, absorption and
immunity, the gut requires a different profile of amino acids. Quantifying these requirements
has become an important goal in understanding the role of the gut in amino acid availability for
the rest of the body. In particular, this availability becomes of paramount concern in situations
of gut disease or stress where increased maintenance requirements can limit the availability of
certain amino acids for whole-body growth and physiological functions.

2. METHODS TO MEASURE AMINO ACID REQUIREMENTS


Many methods have been employed in humans and animals to determine amino acid require-
ments. The advantages and disadvantages of many of these techniques have been extensively
reviewed by others (Lewis, 1992; Fuller and Garlick, 1994; Young and el-Khoury, 1995; Zello
et al., 1995; Waterlow, 1999). Many of these methods were originally developed in animals and
then modified for humans. However, it is important to note that most animal research on amino
acid requirements has primarily focused on the growing phase for economic reasons, whereas
human research has almost exclusively focused on the adult phase which comprises most of the
lifespan. The basic strategy employed by almost all studies involved with the determination of
amino acid requirements includes the feeding of graded levels of the test amino acid and the
measurement of a specific biological response. The choice of biological response depends on
many factors including species, age, health status, sample availability as well as ethical
considerations, analytical equipment availability, financial constraints and practicality.
In choosing the biological response, the most important aspect of amino acid metabolism
to consider is the need for an amino acid for incorporation into protein. If all other essential
nutrients, especially energy and other amino acids, are at or above requirement levels, then
whole-body protein synthesis will occur at a level determined by the intake of the most
limiting amino acid (i.e. the test amino acid). If the intake of this test amino acid is below its
requirement, then protein synthesis will be reduced and the intake of all other essential amino
acids will be in relative excess; because amino acids cannot be stored, this excess must be
catabolized by the body and excreted as bicarbonate and ammonia via carbon dioxide and
urea, respectively. Increasing the intake of the test amino acid will result in greater protein
synthesis and the concomitant reduction in excess amino acid catabolism indicated by lower
carbon dioxide and urea excretion. At intakes above its requirement, the test amino acid will
no longer be the first limiting one and additional intakes will not result in greater protein
synthesis. At these intakes, the test amino acid itself is in relative excess and must be catab-
olized to carbon dioxide and urea. This general scheme has been used to develop almost all
techniques employed to determine amino acid requirements, including growth assays, serial
slaughter, nitrogen balance, plasma amino acids, plasma urea, direct amino acid oxidation,
amino acid balance, and indicator amino acid oxidation (fig. 1).
It is also important to note that the chosen biological response should be amenable to sta-
tistical modelling techniques so that an objective estimate for an amino acid requirement can
be determined, preferably with an estimate of population variance. Because the pattern of the
biological response is rarely predictable over deficient to excess intakes of the test amino
acid, several statistical models have been proposed and employed. The overall response is
Role of intestinal first-pass metabolism 129

Fig. 1. Graphical representation of various response curves to increasing intakes of the test amino acid.
The “breakpoint” requirement is usually determined using two-phase linear regression.

often argued, from a biological standpoint, to be a quadratic model; however, a two-phase


linear regression breakpoint model sometimes fits better. In practice, when both models fit
well, the breakpoint estimate is usually similar between techniques (Baker et al., 2002). Using
either model, the requirement can be determined within individuals and then averaged for a
requirement estimate with population variance. Or the model can be applied to a complete
data set of many animals over many different intakes and the error of the fit could be used to
predict the population variance. In any event, it is obvious that the more data are available for
statistical manipulation, the more versatile the modelling can be. This issue of statistical
manipulation is not a trivial one. Indeed, it has been recently demonstrated that re-analysis of
the classic nitrogen balance studies of Rose and Jones yielded very different conclusions
about the amino acid requirements in humans using the exact same data (Rand and Young,
1999; Di Buono et al., 2001a). These studies have clearly demonstrated that the importance
of the chosen statistical model is almost as important as the data.

2.1. Growth assays

Because the primary role of an amino acid is its incorporation into protein, the measurement
of protein synthesis itself, during varying intakes of the limiting amino acid, can be consid-
ered to be the most direct of approaches. As such, growth assays, and more specifically the
serial slaughter technique, have often been considered as the “gold standard” of techniques in
the determination of amino acid requirements in animals. As the test amino acid intake is
increased towards its requirement, then more protein is synthesized which leads to increased
lean tissue deposition. In young animals with minimal fat deposition, growth is almost
directly proportional to lean tissue deposition and the growth assay is appropriate. Obviously,
as animals approach maturity, more fat is deposited and this proportional relationship
between lean and body growth is not constant. To resolve this discrepancy, the serial slaugh-
ter technique employs body composition analysis to accurately determine lean tissue content.
Using a reference group of animals analysed at the starting body weight, lean tissue accretion
can be determined. However, the main drawback to this approach is the necessity of using
130 R. F. P. Bertolo et al.

different animals at different time points. Also, body tissue analysis for amino acids is not
very accurate given the heterogeneity of ground body samples and the problems associated
with protein-bound amino acid analysis. So although these techniques are simple and direct,
they are limited to fast-growing lean animals with low and constant fat deposition or must
involve large numbers of slaughtered growing animals of similar genetic background to minimize
the inter-animal variation. These techniques are not useful in adult, non-growing animals or
in animals with special conditions (i.e. gestation, lactation, egg production, disease, etc.).

2.2. Nitrogen balance

When protein synthesis is limited, excess amino acids are catabolized to their metabolic end-
products which for all amino acids include bicarbonate and ammonia. The measurement of
these excreted biological products provides an indirect and inverse measurement of changes
in protein synthesis and thus can be used to determine amino acid requirements. Ammonia
enters the nitrogen pool of the body and is excreted primarily as urea in mammals and as uric
acid in birds. Because such excreta are relatively easy to collect and analyse for total nitrogen,
the nitrogen balance method was one of the first to be developed for assessment of amino acid
requirements in humans and animals. The amino acid requirement can be determined from
either balance calculations or over a range of test amino acid intakes. The balance approach
regresses nitrogen balance on test amino acid intake and defines the requirement as the intake
level at which optimum balance is achieved (i.e. zero or positive balance in adults). A consid-
erable drawback with balance calculations is the need to make an assumption of unmeasurable
losses (sweat, skin, nails, hair, etc.). Indeed, some of the original landmark experiments by
Rose and Jones to determine human amino acid requirements did not include such an assump-
tion; when these data were later corrected for an estimate of these losses, amino acid
requirement estimates more than doubled (Young and Marchini, 1990; Rand and Young, 1999).
And because two large numbers are being subtracted (nitrogen intake and nitrogen output),
the difference is relatively very small and this assumption becomes extremely important. An
alternative approach not involving this correction is to measure the qualitative change in nitro-
gen balance over a range of test amino acid intakes. Nitrogen output is high when protein
synthesis is limited by a single amino acid because the other amino acids are in relative excess
and are catabolized. As the test amino acid is added incrementally to the diet, nitrogen output
decreases until the requirement is met and further increments will not stimulate protein
synthesis and nitrogen balance will remain constant (assuming isonitrogenous diets).
The many technical advantages and disadvantages of the nitrogen balance technique have
been extensively reviewed over the years (Waterlow, 1999; Tome and Bos, 2000) and will not
be discussed here. However, several important points need to be mentioned in the present
chapter. An important advantage of the nitrogen balance technique is that unlike growth
assays, this method can be successfully applied to adult as well as young species. The major
disadvantage of the technique is that because of the very large urea pool in most species,
its response to dietary manipulation is rather slow and thus adaptations of a week or greater
are generally required. If one is studying 6 or 7 test amino acid levels, this long adaptation
results in a lengthy experimental period which is not useful in special physiological situations
such as gestation, lactation or disease progression. In addition, this technique is also inap-
propriate in vulnerable populations where long-term feeding of deficient diets is not ethical.
The nitrogen balance technique has been extensively applied for all indispensable amino
acids in many species.
Role of intestinal first-pass metabolism 131

2.3. Direct amino acid oxidation technique

As opposed to the nitrogen balance technique, oxidation methods monitor the excretion of
carbon dioxide, the other obligatory end-product of amino acid catabolism. However, the
basic principle is similar in concept to other methods to determine amino acid requirements.
At deficient test amino acid intakes, the test amino acid is efficiently utilized for protein
synthesis and its oxidation is low and constant. At intakes above requirement, protein syn-
thesis is maximized and excess test amino acid is preferentially oxidized to its end-products.
Instead of measuring changes in nitrogen excretion as in the nitrogen balance technique,
amino acid oxidation methods measure changes in carbon dioxide excretion in breath.
The use of isotopically labelled amino acids allows for an extremely sensitive means of
measuring small changes in amino acid oxidation in response to changes in intake. Although
all carbons of amino acid skeletons are eventually oxidized, the most sensitive approach is to
monitor the expiration of the cleaved carboxyl group at the 1-carbon position. Thus, oxida-
tion of uniformly labelled amino acids results in distribution of the label among many
metabolites which makes interpretation difficult and somewhat less sensitive. Alternatively,
the oxidation of carboxyl-labelled amino acids is more direct and easier to interpret provided
the decarboxylation step is irreversible. However, because one is measuring carbon dioxide
in breath, the carboxyl group must also enter general bicarbonate pools which equilibrate
readily with carbon dioxide expiration at the lungs. For example, experiments with labelled
threonine and methionine have found that non-linear responses are typical with infusion of
these amino acids over varying intakes as a result of their more complex degradative pathways
(Chavez and Bayley, 1976; Zhao et al., 1986; Storch et al., 1988; Ballevre et al., 1990). Thus, as
summarized by Zello et al. (1995), there are general criteria for choosing appropriate carboxyl-
labelled amino acids for oxidation studies: (1) the amino acid must be indispensable; (2) it must
be primarily partitioned between oxidation to carbon dioxide and protein incorporation; and
(3) the labelled carboxyl group must be irreversibly oxidized and sufficiently equilibrated
with labelled carbon dioxide in breath. These restrictions adequately apply for some indis-
pensable amino acids such as phenylalanine (provided excess dietary tyrosine is fed), lysine
and the branched-chain amino acids; as expected, these amino acids have been used exten-
sively in direct oxidation studies. Although the carboxyl group of methionine is irreversibly
oxidized, methionine can equilibrate reversibly with homocysteine prior to the irreversible
oxidative pathway, thus complicating interpretation of oxidation data.
Brookes et al. (1972) were the first to use the oxidation of isotopically labelled amino acids
to determine amino acid requirements in rats. Since then, the direct oxidation technique has
been used to determine the requirements of several amino acids in growing rats (Kang-Lee
and Harper, 1977, 1978; Harper and Benjamin, 1984), adult rats (Simon et al., 1978), young
pigs (Kim et al., 1983a,b; Ball and Bayley, 1984, 1986; House et al., 1997a,b), infants
(Roberts et al., 2001a) and adult humans (Meguid et al., 1986a,b; Meredith et al., 1986; Zhao
et al., 1986; Zello et al., 1990). For animals, the direct oxidation technique has yielded simi-
lar or slightly lower amino acid requirements compared to “classical” techniques such as
nitrogen balance and growth assays; this finding validates the approach to a certain extent. In
contrast, direct oxidation studies in adult humans have yielded much higher (i.e. 2- to 3-fold)
requirements for all tested amino acids compared to those proposed by the FAO/WHO/UNU
(1985) based upon nitrogen balance studies. This latter discrepancy has been extensively
debated and reviewed (Young and Borgonha, 2000). However, one cannot ignore the prob-
lems demonstrated in the interpretation of the small amount of original balance data (Rand
and Young, 1999). Indeed, the human lysine requirement estimated from the re-analysed
132 R. F. P. Bertolo et al.

original nitrogen balance data (Rand and Young, 1999) actually agrees very well with the esti-
mates derived from various techniques using direct oxidation (Meredith et al., 1986), indicator
amino acid oxidation (Zello et al., 1993; Duncan et al., 1996), 24 h oxidation (el-Khoury et al.,
2000) and indicator amino acid balance (Kurpad et al., 2001). In addition, animal studies
employing nitrogen balance techniques, which are numerous and well controlled, result in
more reliable and agreeable results compared to those of more recent kinetic techniques. In
any event, the direct oxidation technique provides as biologically valid an approach as the
nitrogen balance technique and generally agrees with the growth assays performed to date in
animals.
A very important advantage of the oxidation method compared to the nitrogen balance
technique is the more rapid adaptation of the biological response to test amino acid intake
changes. Initial studies in direct amino acid oxidation fed particular test amino acid levels for
7–10 days prior to oxidation measurement, analogous to nitrogen balance studies. However,
more recently it has been demonstrated that prior adaptation to amino acid intake (Zello et al.,
1990; Motil et al., 1994) does not affect the breakpoint estimate of its requirement using the
direct oxidation approach. Therefore, only hours of adaptation to a deficient or excess level
of test amino acid seems to be necessary to measure changes in oxidation, and thus to deter-
mine requirement. To our knowledge, there have been no studies in animals designed to
address this adaptation issue using the direct oxidation technique. However, we have suc-
cessfully employed the direct oxidation technique to measure phenylalanine requirement in
parenterally fed piglets using only 16 h of adaptation prior to oxidation measurement (House
et al., 1997a). The adaptation issue has been carefully addressed in indicator amino acid
oxidation studies in animals and will be discussed below.
It is important to note that this issue of sufficient adaptation is the subject of considerable
debate and has been reviewed (Young and Marchini, 1990). With long-term adaptation to a
deficient diet, the subject will “accommodate” to this situation and possibly become more
efficient in its metabolism. The question is: does this accommodation come at the cost of
other unmeasured metabolic functions? If such costs were incurred, then it violates
Waterlow’s (1985) reasonable definition of adaptation: the process that permits the organism
to respond to a dietary change without adverse consequences. We also need to consider the
definition of amino acid requirement which has been proposed by Young and Borgonha
(2000) as the minimal intake level needed to maintain a specific nutritional criterion such as
growth, body composition, body amino acid balance, organ or system function, etc: the
choice of nutritional criterion then becomes the subject of debate. In spite of this ongoing
debate, the studies in humans have methodically shown that the direct oxidation method has
the distinct advantage over the nitrogen balance technique of very short adaptation periods
resulting in more time-efficient and cost-effective studies.
Although the direct oxidation technique has provided a more sensitive approach to deter-
mination of amino acid requirements, its disadvantages have limited its widespread use in
many species. Most importantly, as mentioned above, not all indispensable amino acids can
be easily used for direct oxidation measurements, limiting its general application. More
specifically, of the indispensable amino acids, only phenylalanine (with excess dietary tyro-
sine), methionine, lysine and the branched-chain amino acids undergo irreversible oxidation
of the carboxyl carbon so that amino acid oxidation can be calculated from expired breath
carbon dioxide. However, Young and colleagues have attempted to predict the requirements
of the other amino acids using previously determined tracer techniques, composition of body
proteins and assumed obligatory oxidative amino acid losses (Young and el-Khoury, 1995);
these predictions were subsequently validated (Raguso et al., 1999). An additional criticism
Role of intestinal first-pass metabolism 133

from a kinetic standpoint is that the feeding of deficient to excess amounts of test amino acid
changes that amino acid’s pool size dramatically, thereby diluting the tracer being measured.
This variable dilution of the tracer in the pool increases variability and reduces sensitivity of
oxidation measurements and hence requirement estimates.

2.4. Indicator amino acid oxidation technique

The indicator amino acid oxidation (IAAO) technique is an extrapolation of the earlier work
with the direct oxidation technique. Both require the accurate measurement of amino acid
oxidation by collecting isotopically labelled carbon dioxide in breath. As with the direct
oxidation method, the IAAO method is based on the hypothesis that the partitioning of amino
acid metabolism between incorporation into protein and catabolism via oxidation is deter-
mined by the most limiting amino acid in the diet. However, the difference between the
techniques is that instead of measuring oxidation of the test amino acid, the IAAO method
measures the oxidation of one of the other amino acids that is also responding to changes in
protein synthesis. When the test amino acid is deficient, protein synthesis is limited and other
amino acids are in excess. Indispensable amino acids in excess must be catabolized at a level
inversely reflecting the rate of protein synthesis which is dictated by the test amino acid
intake. By monitoring the oxidation of one of these “indicator” amino acids over a range of
test amino acid intakes, one can estimate the test amino acid requirement for protein synthesis.
As intake of the test amino acid increases towards requirement, protein synthesis increases
which utilizes more of the indicator amino acid resulting in a smaller excess and lower oxida-
tion. Once the test amino acid intake equals requirement, then greater intakes of this amino acid
will not lead to greater protein synthesis and therefore indicator amino acid oxidation remains
constant.
The choice of indicator amino acid depends on its metabolic characteristics. Phenylalanine
(see below), lysine (Ball and Bayley, 1984; Roberts et al., 2001b) and leucine (Kurpad et al.,
2001) have been used with success to determine various amino acid requirements. Methionine
(Brookes et al., 1972) has been employed unsuccessfully due to its complicated metabolic
pathways, as mentioned previously. When the requirement of a test amino acid has been
determined using more than one of the indicators, the requirement estimates were very similar
(Ball and Bayley, 1984; Zello et al., 1993; Kurpad et al., 2001). Phenylalanine has been used
most often as the indicator with successful determinations of the requirements for lysine (Kim
et al., 1983a; House et al., 1998a), histidine (Kim et al., 1983b), threonine (Kim et al., 1983a;
Bertolo et al., 1998), tryptophan (Cvitkovic et al., 2000), methionine and total sulphur amino
acids (Kim and Bayley, 1983; Shoveller et al., 2001), branched-chain amino acids (Elango
et al., 2002a), arginine (Ball et al., 1986), proline (Ball et al., 1986) and total protein (Ball and
Bayley, 1986) in piglets, tryptophan in trout (Were, 1989), lysine in chickens (Coleman et al.,
2002), lysine in growing pigs (Bertolo et al., 2001), and in humans, lysine (Zello et al., 1993;
Duncan et al., 1996; Kriengsinyos et al., 2002), threonine (Wilson et al., 2000), tryptophan
(Lazaris-Brunner et al., 1998), methionine (Di Buono et al., 2001b), total sulphur amino acids
(Di Buono et al., 2001a), branched-chain amino acids (Mager et al., 2001, 2002; Riazi et al.,
2001) and tyrosine (Bross et al., 2000; Roberts et al., 2001b).
The IAAO technique to determine amino acid requirements was first developed in the
neonatal piglet by Bayley and colleagues. Following observations that amino acid catabolism
depends on the balance of other amino acids (Brookes et al., 1972; Newport et al., 1976),
these researchers successfully demonstrated that the IAAO technique can be used to deter-
mine the amino acid requirements in piglets. In particular, Kim et al. (1983b) showed that the
134 R. F. P. Bertolo et al.

estimate for histidine requirement in piglets was similar using both the direct oxidation and
IAAO techniques with phenylalanine as the indicator. In addition, Ball and Bayley (1984)
found that either phenylalanine or lysine could be used as an indicator because the oxidation
of both responded to varying tryptophan intakes similarly. To further validate the theoretical
concept, Ball and Bayley (1984) also demonstrated that liver protein synthesis was inversely
correlated with phenylalanine oxidation. In addition, this research group demonstrated that
the amino acid requirements for piglets determined using the IAAO technique, namely histi-
dine (Kim et al., 1983b), sulphur amino acids (Kim and Bayley, 1983), lysine and threonine
(Kim et al., 1983a), tryptophan (Ball and Bayley, 1984), proline and arginine (Ball et al.,
1986) and total protein (Ball and Bayley, 1986), agreed very closely with those determined
by classical techniques (NRC, 1979, 1998).
From the initial studies of Bayley and colleagues, the IAAO method has been subsequently
refined and expanded from the original approach. All indispensable amino acids have been
tested using this method; this aspect is the main advantage of the IAAO method over the
direct oxidation technique. In addition, the aforementioned criticism of the direct oxidation
technique regarding amino acid pool size does not apply to the IAAO approach. Indeed,
because the indicator amino acid is fed at the same level over varying test amino acid intakes,
its pool size is not permitted to change and therefore the tracer is not variably diluted. This
unchanging pool size is probably the main reason why the IAAO approach tends to give
requirement estimates with less variability compared to direct oxidation estimates.
As with the direct oxidation technique, the adaptation period required to a particular test
amino acid intake has been shown to be minimal. Indeed, in recent human studies, the lysine
requirement was similar whether hours (Zello et al., 1993; Duncan et al., 1996; Kriengsinyos
et al., 2002), 7 days or 21 days of adaptation (Kurpad et al., 2002b) were employed. This finding
is profound in context with the aforementioned ongoing debate about adaptation versus
accommodation. Furthermore, this adaptation seems to be relatively insensitive to body size
or growth. We have also recently found that phenylalanine oxidation after 1.5 days of adap-
tation (the shortest adaptation tested) to a high or deficient lysine diet was not different up to
8 days of adaptation in both 25 kg growing pigs and 250 kg sows (Bertolo et al., 2001). This
distinct advantage of short adaptation over the classical nitrogen balance or growth assays
allows great versatility in the application of oxidation techniques, especially to vulnerable
populations (Brunton et al., 1998). Indeed, we have recently determined the branched-chain
amino acid requirement of children with the inherited genetic disorder, maple syrup urine
disease (unpublished), as well as the tyrosine requirements of parenterally fed infants (Roberts
et al., 2001a) and the phenylalanine (unpublished) and tyrosine (Bross et al., 2000) requirements
of children with phenylketonuria.
Because an estimate of population variance is critical to recommending amino acid
requirements, the more data in the breakpoint model, the better. Because of the short adapta-
tion time associated with the IAAO method, it is possible to measure amino acid requirements
in individuals. To accurately determine a breakpoint in a two-phase linear regression model,
at least six test amino acid intakes should be included (i.e. three oxidation measurements per
regression). In the nitrogen balance technique, this would require at least 6 weeks of experi-
mentation assuming 7 days of adaptation per diet. Because we have shown that only 1.5 days
of adaptation were necessary in the IAAO method (Bertolo et al., 2001), we recently deter-
mined the lysine requirement of individual growing pigs in 2 weeks by changing diets and
measuring indicator oxidation every other day (Moehn et al., 2001). Similarly, we have
demonstrated that individual amino acid requirements of chickens can be determined in less
than 3 weeks (Coleman et al., 2002). With the determination of enough individual amino acid
Role of intestinal first-pass metabolism 135

requirements, an accurate population variance can be calculated; for animal species, this tech-
nique allows such a calculation for the first time. Indeed, more accurate diet formulation and
animal performance can be achieved with knowledge of such an error within a genetic popu-
lation of animals. In addition, long-term genetic improvement can also be achieved if animals
with low amino acid requirements, or more efficient utilization of dietary amino acids, could
be selected and subsequently bred. The obvious potential for such a technique to be exploited
by various animal production groups is enormous.

2.5. 24 hour oxidation/indicator amino acid balance

The balance technique is based on the principle that in non-growing adults, protein synthesis
is balanced with protein breakdown and thus protein or nitrogen intake is balanced with nitro-
gen excretion. These two balances are connected through the free amino acid pool which is
relatively tightly regulated and represents a minute proportion of total body nitrogen. This
relationship between the two balances is often expressed by the steady-state flux equation
proposed by Waterlow et al. (1978): flux (Q) = synthesis (S) + oxidation/excretion (O) = break-
down (B) + input (I). This equation is rearranged so that protein balance (S − B) = input/output
balance (I − O). This latter equation is the basis of the nitrogen balance calculation technique
as well as the amino acid balance technique.
A recent adaptation of the oxidation techniques incorporates the balance concept described
by Waterlow et al. (1978). Young and colleagues have developed a new method involving a
24 h infusion of amino acid tracer and the measurement of labelled carbon dioxide output in
adult humans (el-Khoury et al., 1994a,b, 1995). These data are then used to calculate carbon
balance at different levels of test amino acid intake. The requirement is taken as the minimal
amino acid intake necessary to maintain balance. This amino acid balance is the difference
between the intake of the test amino acid and whole-body oxidation of that amino acid. This
approach was first employed to determine the leucine requirement in adult humans which
compared very well with their previous direct oxidation experiments (Young et al., 1989).
Subsequently, the technique was successfully used to verify direct oxidation determinations
of aromatic amino acid (Basile-Filho et al., 1997, 1998; Sanchez et al., 1995, 1996) and lysine
(el-Khoury et al., 1998, 2000) requirements. These experiments also demonstrated that when
the test amino acids were fed at the FAO/WHO/UNU (1985) requirement levels, subjects
were in significantly negative amino acid balance, indicating that the present accepted
requirements are too low (Young and Borgonha, 2000). This approach was advanced by the
same group by applying the IAAO technique. Using 24 h labelled leucine infusions, lysine
(Kurpad et al., 2001, 2002b) and threonine (Kurpad et al., 2002a) requirements were
determined by measuring leucine oxidation and balance over a range of test amino acid
intakes. These new techniques are particularly suited for non-growing adults and account for
amino acid metabolism in both the fasting and fed states. However, fasting-state amino acid
kinetics are not relevant to young suckling animals (Bertolo et al., 2000a). Furthermore, this
approach has only been used in humans and may not be widely applicable to fast-growing
meat-producing animals that are in continuous positive nitrogen balance due to high rates of
lean tissue deposition.
As with the nitrogen balance calculation technique, the most important criticism of the
amino acid balance techniques is the reliance on absolute calculations based on various
assumptions. With amino acid kinetic calculations, the most important assumption is that the
precursor amino acid enrichment in readily accessible body pools (i.e. plasma) is representa-
tive of the enrichment of the true intracellular precursors for protein synthesis (i.e. tRNA)
136 R. F. P. Bertolo et al.

and oxidation; this assumption is almost always untrue in constant infusion experiments
(Wolfe, 1992). Intracellular exchange of labelled amino acids is incomplete and variable
among tissues. Indeed, plasma free amino acid enrichments have been found to be 2 to 3 times
greater than those for corresponding amino acyl-tRNA enrichments for leucine, lysine and
phenylalanine (Caso et al., 2001). Because α-ketoisocaproate (KIC) is synthesized from
intracellular leucine only and is released easily into plasma, KIC enrichment has often been
proposed as a suitable representation of intracellular leucine enrichments. However, several
studies have shown that KIC enrichment is not in equilibrium with the entire intracellular
leucine pool (Cobelli et al., 1991; Chinkes et al., 1993) or leucine-tRNA in pigs (Baumann
et al., 1994) or rats (Watt et al., 1991). Alternatively, others have used enrichments of amino
acids that have been incorporated into apolipoprotein B-100 (Reeds et al., 1992; Cayol et al.,
1996; Stoll et al., 1999), fibrinogen (Bennet and Haymond, 1991; Stoll et al., 1999) or albumin
(Cayol et al., 1996). These very rapidly synthesized plasma proteins of hepatic origin can
reflect isotopic steady state within hours. These enrichments are much lower than KIC enrich-
ments but similar to tRNA measurements. Another issue is the possibility that precursors for
oxidation may not equilibrate with tRNA pools, either intracellularly or between tissues, so
that different enrichments may need to be measured to accurately calculate balance. The
amino acid balance techniques will need to address this precursor enrichments issue.
An advantage of the “relative” techniques comparing biological outcomes across dietary
intakes is the avoidance of absolute assumptions. Indeed, in direct oxidation and IAAO analy-
sis of amino acid requirements, the most reliable estimate of requirement with the lowest
error is when percent dose oxidized is used as the biological outcome (House et al., 1997a,b,
1998a; Bertolo et al., 1998; Lazaris-Brunner et al., 1998; Bross et al., 2000; Roberts et al.,
2001a,b). This outcome is in contrast to equivalent measurements of oxidation rate, which
employ flux calculations that also use assumptions about the precursor pool. Ultimately, the
“black box” approach of total labelled carbons in and total labelled carbons out provides the
most reliable requirement estimates. Despite these methodological comparisons, it is impor-
tant to reiterate that the amino acid balance studies to date have calculated amino acid
requirements that are very close to those determined by the more qualitative direct oxidation
and IAAO techniques. Therefore, the error associated with these kinetic assumptions and
calculations may not be as significant as some have proposed.

3. RECENT DEVELOPMENTS IN THE INDICATOR AMINO ACID


OXIDATION TECHNIQUE
Considering that all of the techniques used to determine amino acid requirements generally
agree in their final estimate, one must choose the most versatile method available for wider
application. Given the short adaptation time, applicability to all indispensable amino acids,
ease of biological sampling (i.e. breath) and applicability to most populations, the IAAO
technique has been successfully adapted for use in a variety of situations and purposes. An
important advance in the method was its adaptation for use in parenterally fed piglets as
models for parenterally fed infants (Wykes et al., 1993; Bertolo et al., 1998; House et al.,
1998a). The original work by Bayley in piglets employed one or two oral bolus feeds which
included the isotope dose; the total labelled carbon dioxide excretion was not kinetically
quantified, but rather relatively compared over different test amino acid intakes. With par-
enterally fed piglets, primed constant intravenous infusions of labelled indicator amino acid
were introduced to acquire added information through kinetic calculations. With these studies,
it was shown that the indicator (phenylalanine) flux rate was unchanging across test amino
Role of intestinal first-pass metabolism 137

acid intakes (Bertolo et al., 1998; House et al., 1998a). In addition, it was demonstrated that,
statistically, percent dose oxidized was the most reliable estimate in terms of variability
(House et al., 1997a,b, 1998a; Bertolo et al., 1998); this finding has been confirmed in human
studies (Lazaris-Brunner et al., 1998; Bross et al., 2000; Roberts et al., 2001a,b). Recently, we
have also directly compared amino acid requirements using either intravenous or oral/gastric
infusion of labelled phenylalanine as the indicator amino acid. The breakpoint estimates of
the lysine requirement in adult humans were the same whether the indicator was delivered
intravenously or orally (Kriengsinyos et al., 2002); similarly, the tryptophan requirements in
gastrically fed piglets were similar with intravenous or intragastric infusion of the indicator
(Cvitkovic et al., 2000). These subsequent methodological developments were critical in
adapting the IAAO technique for vulnerable populations.
In order to use the IAAO technique in infants and children, the dietary interventions must
be short and the sample collections must be non-invasive. We have recently validated the use
of oral isotope dosing and collection of urinary amino acids to measure enrichment as repre-
sentative of plasma enrichment (Bross et al., 1998). Study diets were fed hourly for 4 h prior
to dosing and subsequent half-hourly meals with isotope led to enrichment plateaux within 2 h.
This simple non-invasive protocol can be used to study many different vulnerable popula-
tions, provided that dietary ingestion and breath collection are feasible. Similarly, this oral
dosing protocol has been shown to be very effective in determining amino acid requirements
in large pigs where the implantation and maintenance of catheters can be problematic (unpub-
lished data). In addition, such a simplified protocol allows for a broader application of the
technique to other experimental models. Such models include neonatal and adult animals,
gestating or lactating animals, as well as compromised populations which include disease or
surgical interventions.
The IAAO technique can also be applied to research investigating other aspects of protein
metabolism where the goal is not simply to determine amino acid requirements. Because indi-
cator amino acid oxidation responds to protein synthesis, intracellular changes in test amino
acid availability are reflected in changes in indicator oxidation. Recently we have exploited
this principle by adapting the technique to determine true metabolic availability of lysine
from feedstuffs in growing pigs (Ball et al., 2001). We designed a low-lysine diet with all
other indispensable amino acids above requirements. With incremental additions of synthetic
lysine, which is assumed to be 100% available, phenylalanine oxidation declined linearly
until the requirement was met. Within this deficient range of lysine intakes, we used the linear
response equation to predict true availability of lysine from added feedstuffs (fig. 2). When
peas were added to the low (50% of requirement) lysine diet so that the total true available
lysine content was 90% of lysine requirement, the phenylalanine oxidation corresponded to
the availability predicted by ileal digestibility estimates (i.e. the true available amount accord-
ing to NRC, 1998). When peas were heated to render some lysine unavailable via Maillard
products, the phenylalanine oxidation increased and corresponded to an availability of 50%
of unheated peas. When synthetic lysine was added to the heated peas, phenylalanine oxida-
tion decreased below that determined with 90% of requirement, demonstrating that the
increase in oxidation due to heating was due solely to loss of available lysine. This technique
is a new, rapid approach to determining true metabolic availability of lysine that does not
require a series of tenuous assumptions about endogenous losses, which are problematic to
the true ileal digestibility method. This novel application of IAAO principles to improve esti-
mates of amino acid availability has demonstrated the technique’s versatility.
We have also used the basic qualitative principle of the IAAO technique to determine
whether amino acid formulations are inadequate. Recently, by adding amino acids suspected
138 R. F. P. Bertolo et al.

Fig. 2. Representation of novel technique to determine metabolic amino acid availability by measuring
indicator amino acid oxidation.

of being deficient to TPN solutions and monitoring lysine (the indicator) oxidation, we have
systematically demonstrated that the amino acid profile of some commercial TPN solutions
are inadequate (unpublished data). This approach can be further adapted to simulate the strategy
pioneered by Baker and colleagues to determine amino acid requirements by dietary amino
acid supplementation and deletion by employing the ideal ratios to lysine (Mavromichalis
et al., 1998). The important advantage of the IAAO approach is that multiple modifications
can be made to the diets of individual animals because short oxidation measurements are
used, as opposed to the more time-consuming growth assays. Because of the relatively simple
and direct biological response in the IAAO technique, many more extrapolations and adaptations
of the original technique will probably be developed in the future.

4. INTESTINAL FIRST-PASS METABOLISM


In addition to being responsible for the digestion and absorption of nutrients, the gut is also
a major metabolic organ in the body, responsible for the synthesis, conversion and degrada-
tion of amino acids. The gut has a very high metabolic activity and extracts a significant
proportion of the absorbed dietary and endogenous amino acids before transport to the portal
circulation and the rest of the body. Indeed, although the portal-drained viscera (PDV) (intes-
tines, pancreas, spleen, stomach) represents only ~5−7% of body mass, these tissues
disproportionately account for 20–35% of whole-body energy expenditure and protein syn-
thesis (Lobley et al., 1980; McNurlan and Garlick, 1980; Burrin et al., 1990). This significant
extraction of dietary amino acids by splanchnic tissues has been demonstrated by comparing
amino acid kinetics when isotopes are delivered intravenously or orally. The “splanchnic
disappearance” of labelled amino acids when given orally has led many researchers to spec-
ulate on the fate of this irreversible loss of label in kinetics experiments. From these types of
studies, several researchers have determined that in humans and animals, the splanchnic
tissues metabolize between 20% and 50% of dietary essential amino acids (leucine, lysine,
phenylalanine) on first-pass (Yu et al., 1990, 1995; Hoerr et al., 1991, 1993; Biolo et al., 1992;
Matthews et al., 1993; van Goudoever et al., 2000). In addition, with low-protein diets, this
extraction is as high as 70% of lysine intake (van Goudoever et al., 2000). Albeit the data are
very impressive, one problem with this approach is that it is difficult to separate the metabo-
lism of the liver from the PDV, although this can be overcome by measuring amino acid
Role of intestinal first-pass metabolism 139

enrichments and flow in the portal vein (van Goudoever et al., 2000). An additional problem
with the technique is that these types of studies vastly underestimate arterial extraction of
recirculating enteral isotope and hence overestimate first-pass extraction. Indeed, the elegant
study by van Goudoever et al. (2000) demonstrated that a 46% portal mass balance extraction
corresponded to a 22% isotopic extraction by the PDV, which when corrected for arterial
recirculation amounted to insignificant utilization of dietary lysine on first-pass.
The relative importance of nutrient processing by the small intestine (the predominant PDV
organ) versus the liver has only recently been elucidated. Indeed, we have demonstrated that
the small intestine is more important than the liver in modifying nitrogen utilization when
pigs are chronically fed by central vein, portal vein or stomach (Bertolo et al., 1999). In addi-
tion, other researchers have demonstrated that intestinal metabolism dominates splanchnic
metabolism of phenylalanine (Stoll et al., 1997) and lysine (van Goudoever et al., 2000) in
pigs and leucine in dogs (Yu et al., 1990). In addition to dietary extraction of amino acids, the gut
also transports an enormous amount of amino acids from the arterial circulation, especially
during the post-absorptive state. Isotopic data describing the incorporation of amino acids
from both arterial and dietary sources have demonstrated that both sources of precursor amino
acids are critical and that the partitioning between them is dependent on specific amino acid
and dietary protein level (MacRae et al., 1997; Stoll et al., 1999; van Goudoever et al., 2000).
Recent work by van Goudoever et al. (2000) has determined that with high-protein feeding,
almost all of the lysine utilized by the PDV was of arterial origin, whereas with low-protein
feeding, approximately half of the lysine utilized was from both arterial and dietary sources.
It is important to note that level of protein feeding did not affect total lysine use by the PDV,
demonstrating an enormous obligatory utilization of amino acids by the gut for normal func-
tion and growth.
Because of this high obligatory protein turnover in the gut, it follows that with gut chal-
lenges (i.e. tissue damage, increased growth, pathogen exposure, dietary anti-nutritional
factors, etc.) the amino acid requirements of the gut increase. Indeed, first-pass intestinal
extraction of amino acids is proportional to mucosal mass. Stoll et al. (1997, 1999) demon-
strated that phenylalanine splanchnic extraction was 50% higher in pigs raised outdoors,
where rooting and pathogen challenges are greater, compared to pigs raised in a relatively
clean indoor research facility; this higher extraction correlated with measured mucosal mass.
Infestation of pathogens is known to affect growth rate as well as gut function. The stimula-
tion of whole-body and gut immune systems must impart a protein synthetic cost to the
infected animal. Indeed, sepsis in rats has been shown to stimulate intestinal protein synthe-
sis (von Allmen et al., 1992; Higashiguchi et al., 1994b); in particular, sepsis stimulates the
synthesis of endogenous and secretory proteins, including certain gut peptides, in small intes-
tine mucosa (Higashiguchi et al., 1994a). More recently, Yu et al. (2000) have shown that
subclinical nematode infection in sheep increased total gastrointestinal tract leucine seques-
tration by 24% and gastrointestinal tract oxidative losses of leucine by 22–41%. In another
study in pigs, the infusion of endotoxin led to enhanced intestinal catabolism of amino
acids (Bruins et al., 2000). These intriguing studies suggest a possible mechanism for the
growth-stimulating effect of feed-grade antibiotics; increased protein synthesis during
infection by the already metabolically demanding intestinal tissues limits the availability of
amino acids for extraintestinal lean tissue growth. With the impending discontinuation of
prophylactic feeding of antibiotics in animal production, the prevention of subclinical infec-
tions or challenges needs to be a priority in the development of alternative strategies in the
near future.
140 R. F. P. Bertolo et al.

The primary fate of indispensable amino acids is presumably to protein synthesis; however,
recent intriguing data have demonstrated that catabolism dominates the first-pass utilization
of these amino acids by the gut (Stoll et al., 1998; Wu, 1998; van Goudoever et al., 2000).
This seemingly wasteful oxidation of indispensable amino acids amounts to a small but signif-
icant proportion of dietary intake (Yu et al., 1992, 2000; Cappelli et al., 1997; van Goudoever
et al., 2000), but a large proportion of whole-body amino acid oxidation (van Goudoever et al.,
2000). Indeed, we have recently demonstrated that phenylalanine oxidation is significantly
greater when labelled phenylalanine is delivered orally, as opposed to intravenously in the
IAAO technique (Cvitkovic et al., 2000; Kriengsinyos et al., 2002). This increased oxidation
during feeding of adequate diets demonstrates that first-pass catabolism of phenylalanine by
the gut is significant.
Whatever the fate of indispensable amino acids extracted by the gut, the evidence clearly
suggests that this tissue plays a significant role in modulating the profile of amino acids delivered
to the rest of the body (Stoll et al., 1998; Wu, 1998; Bertolo et al., 2000b). Much of this role
results in a net loss of amino acids to extraintestinal tissues. Presumably, when gut metabolic
activity is increased by growth or stress, so is the loss of dietary amino acids from whole-body
functions. This aspect of whole-body amino acid requirements has not been fully explored.
In addition to gut metabolic activity, the quality of dietary protein (Deutz et al., 1998; Gaudichon
et al., 1999; Mariotti et al., 1999) and type of dietary carbohydrate (van der Meulen et al.,
1997) also influence first-pass extraction of amino acids. So the actual availability of dietary
amino acids for muscular protein synthesis is highly dependent on the metabolic activity of
intestinal tissues. It is important to note that this concept is accommodated by our recent
adaptation of the IAAO technique mentioned above for determination of true metabolic avail-
ability of dietary amino acids from heat-treated feedstuffs.
Given the significant demand of the gut for dietary and arterial amino acids, it is obvious
that the maintenance and growth of this organ already constitutes a significant proportion of
whole-body amino acid requirements. Furthermore, it follows that in certain situations that
increase the metabolic activity of the gut, this proportion will increase at the expense of
whole-body growth. Indeed, this hypothesis is supported by the abovementioned studies in
subclinical nematode infection and the reduced growth rate commonly observed in gastro-
intestinal disease. If the availability of an indispensable amino acid is already limiting in an
animal’s diet, then an unobservable subclinical challenge to the gut could feasibly limit
animal growth further. Although this concept seems intuitive, it is very difficult to demonstrate
experimentally.

5. IMPACT OF INTESTINAL METABOLISM ON AMINO


ACID REQUIREMENTS
Many of the approaches recently employed to demonstrate changes in amino acid utilization
by the gut or PDV tissues could be adapted to quantify amino acid requirements of these
tissues relative to whole-body requirements. It has been well demonstrated that the propor-
tion of dietary amino acids extracted by the gut changes with dietary composition; but a
systematic approach to dietary ingredient requirements for the gut has yet to be published. By
employing graded intakes of indispensable amino acids, it is possible to determine the mini-
mum amount required by the gut for normal function and growth. To date, our body of
literature regarding amino acid requirements during intravenous or intragastric feeding in
piglets provides the only attempt, to the authors’ knowledge, to quantify requirements with
and without first-pass splanchnic metabolism.
Role of intestinal first-pass metabolism 141

5.1. The parenterally fed piglet as a model of splanchnic metabolism

The emerging evidence that the gut utilizes a significant proportion of dietary amino acids led
us to speculate that in situations of gut bypass or stress, the whole-body amino acid require-
ments must be different. We proposed that total parenteral nutrition (TPN), which bypasses
first-pass metabolism by the small intestine and liver and can cause gut atrophy (Johnson
et al., 1975; Goldstein et al., 1985; Alverdy, 1995; Bertolo et al., 1999; Burrin et al., 2000),
would lead to changes in amino acid requirements that can be measured. Indeed, we (Duffy
and Pencharz, 1986; Bertolo et al., 1999) and others (Sim et al., 1979; Lanza-Jacoby et al.,
1982; Jeevanandam et al., 1987) have shown that TPN feeding alters whole-body nitrogen
metabolism compared to oral feeding. The different nitrogen utilization as a consequence of
parenteral feeding may be due to reduced gastrointestinal metabolism associated with gut
atrophy and/or due to lack of hepatic first-pass metabolism. With the development of the TPN-
fed piglet model (Wykes et al., 1993), we subsequently demonstrated using three infusion
routes that intestinal atrophy has a greater impact on nitrogen metabolism than liver bypass
(Bertolo et al., 1999). Whatever the fate of indispensable amino acids in the gut, an atrophied
gut will utilize fewer amino acids and thereby affect whole-body requirements. Therefore, we
have proposed that the TPN-fed piglet is a “gut-deficient” model. So the amino acid require-
ments of TPN-fed piglets approximate the requirements for extraintestinal tissues. When
compared to a piglet gastrically fed identical diets, the amino acid requirements for the intact
gut could at least be estimated (table 1).
However, TPN feeding is only one of many relevant clinical scenarios that lead to com-
promised gut metabolic capacity. Gut dysfunction can also be caused by malnutrition,
diarrhoea, chemotherapy, gastrointestinal surgery and gastrointestinal diseases. Furthermore,
with relevance to the animal industry, gut stress is often associated with weaning, especially

Table 1
Effects of gut atrophy and bypass on whole-body amino acid requirements as determined in
parenterally and enterally fed piglets using the indicator amino acid oxidation technique

Oral requirement Parenteral requirement Gut bypass effecta


Amino acid (g/kg/d) (g/kg/d) (%)

Threonineb 0.42 0.19 55


Methioninec 0.25 0.18 28
Total sulphursc 0.42 0.29 31
Branched-chaind 2.64 1.53 42
Tryptophane 0.13 and 0.11 0.14 0
Lysinef 0.85 0.79 7
Phenylalanineg 0.50 0.45 10
Tyrosineh 0.30 0.35 0
a This effect includes atrophy from 7 d on TPN as well as bypass of first-pass gut metabolism.
b Bertolo et al. (1998).
c Shoveller et al. (2001). Methionine requirement was determined with excess dietary cysteine; total sulphurs refer

to the methionine requirement determined with no dietary cysteine.


d Elango et al. (2002a). Leucine, isoleucine and valine dietary ratio (NRC, 1998) was maintained across intakes.
e Cvitkovic et al. (2000). Tryptophan requirement using intravenous (0.13) or oral isotope (0.11).
f House et al. (1998a). Oral requirement for lysine was estimated from NRC (1998).
g House et al. (1997b). Parenteral phenylalanine requirement was determined using direct oxidation technique;

oral phenylalanine requirement was estimated from NRC (1998).


h House et al. (1997a). Oral tyrosine requirement was estimated from NRC (1998).
142 R. F. P. Bertolo et al.

with the increasingly popular early weaning practice in swine production. Any clinical situa-
tion which diminishes gut capacity would affect the metabolism of many amino acids; for
example, increased amino acid requirements discovered during TPN feeding should also be
applied to the treatment of any of the above conditions. Thus, our research has relevance to
many conditions in addition to the TPN-fed neonate. However, we chose to use the TPN-fed
piglet as a model for diminished small intestinal capacity because of demonstrated gut
atrophy, reproducibility and ease of development compared to gut disease models. With this
clinically relevant model, we could extrapolate our results to any situation involving gut
dysfunction.

5.2. Threonine

Perhaps the most impressive effect of reduced gut metabolism on whole-body amino acid
requirement was demonstrated for threonine. Using the TPN-fed versus gastrically fed piglet,
we have demonstrated that the threonine requirement is reduced by 55% when the gut is
bypassed and atrophied (Bertolo et al., 1998). These data are supported by Stoll et al. (1998),
who also observed that the PDV tissues extract 60% of dietary threonine measured by both
net portal balance and labelled threonine extraction. In addition, van Goudoever et al. (2000)
showed that when pigs were fed the high-protein diet, 84% of threonine was retained by the
gut, and on the low-protein diet, all of the threonine was retained. This enormous demand for
threonine by the gut is probably reflected by its role in mucin synthesis for maintenance of
the luminal mucus layer (Lamont, 1992). Intestinal mucins are continuously secreted by the
intestines and are critical in the defence of the mucosa from mechanical and pathogenic
insults. The core protein of mucins contains a disproportionate amount of threonine, proline
and cysteine (Specian and Oliver, 1991). With parenteral feeding and gut atrophy, mucin syn-
thesis is reduced and so is the gut’s requirement for threonine. Recently we have
demonstrated that feeding threonine-deficient diets to gastrically fed piglets reduces gut
growth and goblet cell numbers and alters the mucin profile of intestinal mucus; in addition,
parenteral threonine cannot completely restore normal gut function and histology compared
to enteral threonine (Ball et al., 1999; table 2).
This profound impact of gut metabolism on whole-body threonine requirement must also be
considered as a minimum effect. Because the parenteral threonine requirement (0.19 g/kg/d)
was so much lower than NRC (1998) recommendations (0.53 g/kg/d on a true ileal digestible
basis), we introduced the gastrically infused control group which received identical diets to
verify NRC estimates for piglets. The requirement for these control pigs (0.42 g/kg/d) was
lower than that recommended by NRC, but still substantially higher than the parenteral
requirement. This discrepancy with NRC values is not surprising given that the requirements
recommended by NRC (1998) include digestibility estimates for corn–soybean diets and adjust
these values for endogenous loss estimates, whereas our diet was completely elemental and
available. In subsequent IAAO experiments, we have demonstrated that these NRC estimates
are proportionately closer to our gastrically fed estimates for tryptophan (0.15 vs 0.15 g/kg/d
for NRC), methionine (0.25 vs 0.23 g/kg/d for NRC) and methionine plus cysteine (0.42 vs
0.48 g/kg/d for NRC). The discrepancy in threonine requirements between that recommended
by NRC and our oral estimates could be due to the increased sensitivity of the threonine
requirement to endogenous losses. Because the mucin protein core is resistant to digestion
and is almost completely recovered in ileal digesta (Mantle and Allen, 1981), a major com-
ponent of endogenous losses at the ileum is mucins which are rich in threonine, proline,
serine and cysteine (Specian and Oliver, 1991). Mucin secretion and hence losses are known
Role of intestinal first-pass metabolism 143

Table 2
Goblet cell parameters in piglets fed adequate oral threonine (IG-A), deficient oral threonine
(IG-D) or deficient oral threonine with adequate parenteral threonine (IV-A)a

PAS/AB 2.5b AB 2.5b AB 1.0b

Cellsc Staind Cells Stain Cells Stain

Duodenum IG-A 18.8a 39 17.6a 44a 16.3a 27


IG-D 15.2b 41 5.5c 6b 7.1b 12
IV-A 14.9b 39 12.6b 25a,b 13.5a,b 31
SDpooled 2.7 13 4.4 12 4.4 15
Ileum IG-A 20.9 79 30.1a 63a 24.6a 30
IG-D 22.6 68 13.9b 14b 13.3b 13
IV-A 27.5 82 19.9a,b 31a,b 20.7a,b 31
SDpooled 5.9 18 7.8 18 6.5 11
a Ball et al. (1999). Data are means for n = 7 piglets. For data with letter superscripts within a row, those not
sharing a letter are different (P < 0.05, LSD comparisons).
b PAS/AB 2.5: staining is combination of Alcian blue/periodic acid (5 min)-Schiff base (15 min) (PAS) reaction

allowing unsubstituted α-glycol-rich neutral mucins (pink) and acidic mucins (blue) to be differentiated; AB 2.5:
1% Alcian blue (AB, pH 2.5, 1 h) for the localization of carboxylated and/or sulphated acidic mucins; AB 1.0:
1% Alcian blue (AB, pH 1.0, 1 h) for the selective identification of sulphomucins.
c Goblet cells in the mucosa stained with PAS/AB 2.5, AB 2.5 or AB 1.0 were counted in 10 well-oriented

crypt-villus units ~25 μm in each animal.


d Semi-quantitative staining intensities based upon a scale ranging from 0 (unreactive) to 3 (intensely stained)

were multiplied by total number of goblet cells.

to be sensitive to dietary composition as well as presence of fibre and anti-nutritional factors


(More et al., 1987; Sharma and Schumacher, 1995). Indeed, in a very recent experiment, we
have demonstrated that dietary supplementation of wheat bran, a stimulant of mucin synthe-
sis, leads to increased ileal losses of threonine which may affect whole-body availability of
dietary threonine (Myrie et al., 2002). Therefore, this is probably the reason why the threonine
requirement of pigs fed a fibre-free elemental nutrition solution (i.e. our gastrically fed control
pigs) was lower than the estimated requirement of pigs fed a corn–soybean meal diet (i.e. NRC
recommendations). Indeed, as a percentage of the NRC recommendation, the threonine
requirement in parenterally fed pigs was 36% (instead of 45%), which translates to a gut
utilization of 64% of dietary threonine. Therefore, the requirement for threonine by the gut
versus the whole body depends on dietary composition.

5.3. Sulphur amino acids

Although methionine is an indispensable amino acid, cysteine is not because it can be syn-
thesized from methionine. However, increased metabolism of methionine to meet cysteine
needs could limit methionine availability for protein synthesis and growth. As a result, dietary
cysteine has a “sparing effect” on the amount of methionine required. In a series of experi-
ments, we have recently determined the methionine requirements of piglets fed orally and
intravenously, with and without dietary cysteine, using the IAAO technique (Shoveller et al.,
2001; table 1). With excess or without dietary cysteine, the methionine requirements in par-
enterally fed piglets were 72% or 69% of the respective requirements in orally fed piglets. In
other words, approximately 30% of dietary methionine is utilized by the gut whether dietary
cysteine is present or not. These data are supported by those of Stoll et al. (1998),
144 R. F. P. Bertolo et al.

who showed that approximately 30% of dietary methionine disappears in first-pass meta-
bolism by the PDV. Furthermore, the sparing effect of cysteine was similar in pigs fed either
intravenously or orally (i.e. excess cysteine reduced the respective methionine requirements
by 40% regardless of feeding route) (Shoveller et al., 2001). In addition to demands for
protein turnover, this relatively high demand by the gut for both sulphur amino acids can
also be attributed to other metabolic functions specific to methionine and cysteine. It is
possible that the high nucleic acid turnover of intestinal cells requires a significant amount
of methionine, an important methyl donor. Furthermore, cysteine, as a product of methionine
metabolism, is incorporated to a large extent into intestinal mucins and the tripeptide anti-
oxidant glutathione; both of these products are critical for the maintenance of the mucosal
tissue and protection against pathogens (Martensson et al., 1990). Indeed, the impact of a
pathogenic challenge on the methionine requirement may therefore be significant, but has yet
to be studied.

5.4. Branched-chain amino acids

The aforementioned data regarding substantial first-pass splanchnic metabolism of leucine


suggest that the splanchnic tissues extract a surprisingly large amount (i.e. 20−50%) of
branched-chain amino acids (BCAA). Recently, using a diet with a fixed ratio of BCAA
(1:1.8:1.2, isoleucine:leucine:valine), we have determined that the BCAA requirement in
intravenously fed piglets was 56% of that in intragastrically fed piglets (Elango et al., 2002a).
The apparent uptake of 44% of enterally fed BCAA by the splanchnic tissues is a significant
finding because it is generally accepted that the BCAA are predominantly metabolized by the
extrahepatic tissues due to the higher activity of branched-chain aminotransferase (BCAT),
the first enzyme in the catabolic pathway of the BCAA, in skeletal muscle compared to the
liver. In addition, the pattern of BCAA in the plasma of enterally fed piglets, when compared
with parenterally fed piglets, clearly demonstrates that the gut has a high demand for leucine
and a clear preference for leucine compared to isoleucine or valine. The observation, during
enteral feeding, that plasma valine and isoleucine concentrations increased while leucine
concentrations remained low indicates that leucine is being extracted by the gut and there-
fore may be limiting protein synthesis in the rest of the body. Valine and isoleucine do not
appear to be utilized by the gut to the same extent and are therefore being passed to the
systemic circulation, but because protein synthesis is limited by leucine, these two amino
acids, as well as most of the other indispensable amino acids, increase in concentration in
the plasma.
The difference in BCAA requirements between routes of feeding is supported by the data
in humans (Gelfand et al., 1988; Hoerr et al., 1991, 1993; Biolo et al., 1992; Matthews et al.,
1993) and dogs (Yu et al., 1990, 1995), regarding first-pass splanchnic extraction of leucine
measured by isotope infusions. In addition, Stoll et al. (1998) reported that the pig PDV
extracted 43% of leucine, 39% of valine and 31% of isoleucine. Altogether, the first-pass
extraction data compare well with the 44% lower BCAA requirement in parenterally fed
piglets observed in our recent study. In a subsequent study, we adapted the IAAO technique
to systematically determine which of the BCAA was most limiting (Elango et al., 2002b).
In both orally and intravenously fed piglets, diets moderately deficient in BCAA (75% of
respective requirement) were fed and indicator amino acid oxidation determined. Piglets
were then randomly assigned to receive one of three test diets containing either isoleucine,
leucine or valine to meet 100% of requirement, with the remaining two amino acids at 75%.
Role of intestinal first-pass metabolism 145

Fig. 3. The change in phenylalanine oxidation in parenterally (IV) or enterally (IG) fed pigs after
supplementation of individual amino acids (Elango et al., 2002b). Diets were formulated to be slightly
deficient in all branched-chain amino acids and indicator oxidation was performed before and after
isoleucine, leucine or valine supplementation. * indicates oxidation change was different than zero.

The difference in phenylalanine oxidation between unsupplemented and supplemented diets


was used as an indicator of BCAA adequacy (fig. 3). In orally fed piglets, the difference in per-
cent dose oxidized was not significant for any supplemented amino acid. However, in
parenterally fed piglets, isoleucine and valine supplementation decreased phenylalanine oxi-
dation; isoleucine had the greatest effect and was first limiting (i.e. oxidation decreased from
approximately 22% to 10%) and valine was second limiting (22% to 15%). Leucine, which
is the preferred amino acid by the gut according to our previous data, had no effect when sup-
plemented to gut-atrophied, parenterally fed pigs. The optimal ratio of BCAA for orally fed
pigs is adequately predicted by requirement estimates (NRC, 1998). However, because the
gut does not utilize the BCAA in this proportion, the ideal ratio of BCAA for maintenance of
the gut is not the same as that for the whole body and has not yet been determined.

5.5. Tryptophan and lysine

The tryptophan requirements of parenterally and orally fed piglets were not different when
identical diets were employed (Cvitkovic et al., 2000). This result suggests that the gut’s
requirement of tryptophan for protein synthesis and/or for oxidation does not significantly
impact whole-body requirements, possibly due to either efficient recycling by an atrophied
gut or to this amino acid’s low proportion in protein.
We have also determined the lysine requirement of parenterally fed piglets but did not
employ a gastrically fed control group (House et al., 1998a). Because this was the first amino
acid for which we determined the requirement during TPN feeding, we did not appreciate the
large impact of the gut. Because NRC estimates were proportionately similar to gastrically
fed control pigs for other amino acids, we therefore compared the parenteral lysine require-
ment of 0.79 g/kg/d to the NRC (1998) estimate of 0.85 g/kg/d, providing a 7% difference
due to gut metabolism. It is important to note that this comparison may not be valid given the
lack of empirical data for amino acid requirements in young piglets (NRC, 1998). The net
lysine utilization by the gut may still be significant as in previous studies regarding the
146 R. F. P. Bertolo et al.

substantial splanchnic extraction of lysine (Hoerr et al., 1993; van Goudoever et al., 2000);
however, it appears that in piglets, this utilization is not nearly as profound as that for threo-
nine, the sulphur amino acids or the branched-chain amino acids. The impact of parenteral
feeding (i.e. with gut atrophy and lack of first-pass metabolism) on whole-body lysine
requirements is probably accounted for by the reduced general protein turnover in the atro-
phied gut; in other words, there seems to be no disproportionate requirement for lysine by the
gut versus the whole body.

5.6. Phenylalanine and tyrosine

By employing the direct oxidation approach, we found only moderate differences in the
phenylalanine (House et al., 1997b) and tyrosine (House et al., 1997a) requirements of
parenterally fed piglets compared to the NRC (1998) estimates of oral requirements. The
parenteral tyrosine requirement was estimated at 0.35 g/kg/d; in addition, the parenteral
phenylalanine requirement (with excess dietary tyrosine) was only 10% lower than NRC
estimates (0.45 vs 0.50 g/kg/d). Combined, the phenylalanine plus tyrosine requirement
of parenterally fed piglets (0.80 g/kg/d) was equal to that estimated by NRC (0.80 g/kg/d).
Furthermore, we have recently been able to compare these results in piglets to the parenteral
tyrosine requirement in parenterally fed human infants (Roberts et al., 2001a); the tyrosine
requirement determined for parenterally fed infants was similar to the broad range recom-
mended for orally fed neonates (Snyderman, 1971). Again, this minor effect of gut
metabolism on whole-body requirements is in contrast to dual-isotope infusion studies (Biolo
et al., 1992; Matthews et al., 1993) where 29–58% of dietary phenylalanine was extracted by
splanchnic tissues in adult humans; however, arterial recirculation (van Goudoever et al.,
2000) was not estimated in these studies. In addition, we have shown that when the indicator
amino acid is delivered orally or intravenously, basal phenylalanine oxidation is significantly
increased when first-pass splanchnic metabolism is maintained in adults (Kriengsinyos et al.,
2002) or piglets (Cvitkovic et al., 2000).
However, an increase in phenylalanine oxidation with first-pass metabolism by the gut does
not necessarily translate to a substantial extraction of dietary phenylalanine, as demonstrated
with lysine by van Goudoever et al. (2000). Indeed, although phenylalanine oxidation
increased by 70% when infused orally versus intravenously in humans, the percent extraction
determined by flux rates was only increased by 30% (Kriengsinyos et al., 2002). Furthermore,
phenylalanine oxidation amounted to less than 15% of intake during either route of infusion.
In piglets, when the indicator phenylalanine was fed at the requirement, oxidation rates were
increased from 0.6% to 0.8% of dose, which translated to 0.9% and 1.5% of phenylalanine
intake or requirement (Cvitkovic et al., 2000). These latter data also provide a reason why
phenylalanine is a good choice for the indicator amino acid. Even if there are large differences
between diets or individuals in the level of phenylalanine oxidation, the total amount oxidized
is still rather insignificant relative to the quantity used for protein synthesis which is driving
the whole-body requirement for the amino acid.
As with lysine and tryptophan, it appears that dietary phenylalanine is primarily utilized for
non-specific protein synthesis, which does not appear to disproportionately impact whole-
body requirements. It has been suggested that the gut may hydroxylate a substantial amount
of phenylalanine to tyrosine (Stoll et al., 1998); however, the impact of this capacity on
whole-body hydroxylation or requirements has yet to be explored. Estimates of phenylalanine
hydroxylation to determine tyrosine requirements have been shown to be unreliable (House
et al., 1998b; Thorpe et al., 2000; Roberts et al., 2001a).
Role of intestinal first-pass metabolism 147

5.7. Arginine and proline

Although for most species studied arginine is considered dispensable, arginine has been
found to be an indispensable amino acid in some species such as cats (Morris and Rogers,
1978), chicks (Tamir and Ratner, 1963) and ferrets (Deshmukh and Shope, 1983). Furthermore,
arginine may be conditionally indispensable in young mammals including the dog (Visek,
1984), rat (Borman et al., 1946) and piglet (Mertz et al., 1952; Ball et al., 1986; Brunton et al.,
1999), which means arginine can be synthesized de novo, but not at sufficient rates to maintain
required functions (i.e. syntheses of protein, urea cycle intermediates, creatine, nitric oxide, etc.).
The neonatal small intestine has been suggested to be the major site of arginine synthesis (Wu
et al., 1994; Stoll et al., 1998) and the ontogeny of the necessary enzymes in enterocytes has
been well described (Wu, 1998).
We initially planned to use the IAAO technique to determine the arginine requirement of
piglets during parenteral feeding. An initial pilot experiment was conducted whereby par-
enterally fed piglets were fed arginine-free diets so that growth and nitrogen balance data
could be assessed. Both pigs experienced severe hyperammonemia after only ~16 h without
dietary arginine; one pig died and the other was comatose. Because plasma ammonia con-
centration was found to be a sensitive indicator of arginine deficiency, we used this biological
outcome to determine if the arginine synthesis rate of the piglet gut was sufficient to main-
tain the urea cycle and whether proline, the primary precursor of arginine synthesis in the gut,
must be available to maintain synthesis rates. The subsequent study successfully demon-
strated that parenterally fed piglets could not synthesize sufficient arginine to maintain the
urea cycle, let alone to maintain growth, whether or not proline was present in the diet. This
study also demonstrated that orally fed piglets could not synthesize arginine and proline at
rates sufficient to maintain plasma concentrations or to prevent hyperammonemia. However,
unlike the gut-atrophied parenterally fed piglets, the gut-intact orally fed piglets experienced
less severe hyperammonemia when proline was provided in the diet. These data suggested
that the conversion of proline to arginine occurs in the piglet, but only during oral feeding.
We therefore hypothesized that this conversion occurs almost exclusively in the gut and that
parenterally fed piglets could not use proline for arginine synthesis because of gut atrophy
and/or gut bypass during feeding. This experiment convincingly demonstrated the essentiality
of arginine and proline in continuously fed piglets; we predicted that with voluntary feeding,
animals would refuse feed if severely deficient, especially if elevated plasma ammonia levels
develop. Indeed, vomiting is a symptom of hyperammonemia in pigs, which functions to
lessen the ammonia load by expelling potentially toxic amino acids.
This clear demonstration of arginine indispensability in piglets was followed by a multi-
isotope, dual-route infusion study whereby labelled proline, ornithine and arginine were
infused intragastrically or into the portal vein to isolate the in vivo effects of small intestinal
first-pass metabolism. This experiment demonstrated that the conversion from proline to argi-
nine is completely dependent on the small intestine, confirming the conclusions from our
previous experiment assessing hyperammonemia. Thus, in situations where gut metabolism
is bypassed or compromised, such as during TPN feeding or gut disease, arginine synthesis
is diminished, increasing the overall requirement compared to normal oral feeding. In addi-
tion, this study demonstrated that proline synthesis from arginine is also dependent on gut
metabolism and its requirement also would be higher when gut metabolism is bypassed.
Although the cumulative evidence clearly indicates that arginine synthesis is dependent on
the neonatal small intestine, the impact of this first-pass metabolism on whole-body arginine
requirements can only be estimated.
148 R. F. P. Bertolo et al.

Most of these data support the hypothesis that the arginine requirement is much higher
during parenteral feeding, for both maintenance and growth, due to lower synthetic capacity
of an atrophied gut. Indeed, we speculate that for arginine, a separate maintenance require-
ment can be distinguishable from the growth requirement. Such a hypothesis has enormous
implications in neonatal populations in which small intestinal first-pass metabolism is
bypassed or compromised by gastrointestinal disease or stress. Indeed, these implications
have been demonstrated in parenterally fed infants (Heird et al., 1972) and adult rats with
small intestinal resection (Wakabayashi et al., 1995). This latter study is interesting consider-
ing that adult rats normally can synthesize adequate amounts of arginine in the kidney, but the
citrulline precursor originates in the small intestine (Morris, 1992). So although net intestinal
arginine synthesis declines during late suckling as renal synthesis increases (Wu, 1998), the
importance of intestinal metabolism in the inter-organ synthesis of arginine is still potentially
critical in adult species that normally do not require arginine.
Proline has been suggested to be an indispensable amino acid for the piglet (Ball et al.,
1986), but subsequent studies indicated that proline indispensability is dependent on avail-
ability of precursors such as glutamate (Murphy et al., 1996; Wu, 1998) and arginine (Brunton
et al., 1999). The extensive gut metabolism of glutamate and glutamine (Windmueller and
Spaeth, 1980; Stoll et al., 1999; Reeds et al., 2000) may limit arginine and proline synthesis
in certain conditions. Furthermore, the importance of proline for collagen synthesis probably
increases in situations of injury and stress. The obvious interdependence of arginine and pro-
line requirement on gut health as well as availability of precursors makes the quantification
of such requirements very complicated. However, there is enough evidence to date to suggest
that the impact of gut first-pass metabolism on whole-body requirements must be significant
and warrants future investigation.

6. FUTURE PERSPECTIVES
Albeit the parenterally fed piglet model has proved useful in estimating the impact of first-
pass gut metabolism on whole-body requirements, a more direct determination of gut
requirements for amino acids has yet to be developed. With a more direct technique,
researchers could then attempt to quantify the effects of gastrointestinal stress, injury, disease
or dysfunction on whole-body requirements. In particular, the recent intriguing work on
leucine extraction and nematode infection in sheep provides preliminary evidence for the
importance of this type of investigation. In addition, Klasing and Calvert (1999) have pro-
vided an important advance in this area by estimating that the percent of lysine intake
consumed by the chicken immune system increases from 1.2% to 6.7% with an injected
immune challenge. This “cost” of an immune challenge must be even more profound with a
gastrointestinal pathogenic challenge where the additional costs of gut secretion stimulation,
gut tissue repair and compromised dietary absorption of lysine must be considered.
Furthermore, this cost mostly relates to lysine requirements for non-specific protein synthetic
processes. The costs for threonine, the branched-chain amino acids, methionine and cysteine
would be much greater due to their more specialized roles in gut maintenance. In addition,
arginine and proline requirements would also be increased due to reduced synthesis as well
as increased utilization. The nutritional consequences of gastrointestinal disease and stress
require further investigation considering that its importance in the treatment of such condi-
tions cannot be underestimated.
Because emerging evidence is demonstrating a substantial impact of gut first-pass metab-
olism on amino acid requirements, the accommodation of this metabolism will undoubtedly
Role of intestinal first-pass metabolism 149

generate interest, especially among researchers in animal production and clinical treatment of
gut diseases. The role of the native microbial population and subclinical gut infections has an
important impact on whole-body requirements; indeed, the higher amino acid utilization effi-
ciency with growth-promoting feed-grade antibiotics may partially be explained by the
minimization of subclinical challenges. Considering that the small intestinal capacity to
digest and absorb protein and amino acids is substantially greater than possible dietary inputs
(Burrin et al., 1999), one is tempted to consider that much of this organ’s demand for amino
acids for maintenance may be an unnecessary burden. Considering that the protein compo-
nent, and especially synthetic amino acids, is the most expensive component of animals feeds,
the cost of maintaining the surplus capacity of the gut becomes significant. In addition, recent
evidence has demonstrated that a significant proportion of whole-body amino acid catabolism
occurs in the gut, presumably for energy. Alternative fuels may be sought to replace this per-
haps unnecessary, expensive source of energy. Another compelling solution would be to select
animals for lower intestinal metabolism without compromising gut absorptive capacity or
protective functions. These animals would have a substantially greater amino acid utilization
efficiency.
The IAAO technique has proven very useful and versatile in determining amino acid
requirements of animals. Its adaptation for vulnerable populations is especially advantageous
if amino acid requirements during pathogenic challenges or disease states are to be explored.
Furthermore, the development of the technique to determine metabolic availability of amino
acids is an important advance in the field of amino acid digestibility, which is of critical
importance when providing dietary amino acids to meet requirements. Because the IAAO
technique relies on relative differences, its dependence on questionable kinetics assumptions
is minimal. This is particularly important given the accumulating evidence that luminal and
arterial amino acids are channelled differently intracellularly; the amino acid pools for oxi-
dation and protein synthesis are probably separated to some extent so that respective
precursor enrichments may be very different, rendering kinetic equations irrelevant. Further
adaptation of the IAAO technique to answer many of these questions is eagerly sought.

REFERENCES
Alverdy, J.C., 1995. Amino acids to support gut function and morphology. In: Cynober, L.A. (Ed.), Amino
Acid Metabolism and Therapy in Health and Nutritional Disease. CRC Press, New York, pp. 435−440.
Baker, D.H., Batal, A.B., Parr, T.M., Augspurger, N.R., Parsons, C.M., 2002. Ideal ratio (relative to
lysine) of tryptophan, threonine, isoleucine, and valine for chicks during the second and third weeks
posthatch. Poultry Sci. 81, 485−494.
Ball, R.O., Atkinson, J.L., Bayley, H.S., 1986. Proline as an essential amino acid for the young pig.
Brit. J. Nutr. 55, 659−668.
Ball, R.O., Bayley, H.S., 1984. Tryptophan requirement of the 2.5-kg piglet determined by the oxidation
of an indicator amino acid. J. Nutr. 114, 1741−1746.
Ball, R.O., Bayley, H.S., 1986. Influence of dietary protein concentration on the oxidation of phenylala-
nine by the young pig. Brit. J. Nutr. 55, 651−658.
Ball, R.O., Bertolo, R.F.P., Pencharz, P.B., Moehn, S., 2001. A rapid method to determine “true metabolic
availability” of amino acids in feedstuffs for pigs. J. Anim. Sci. 79, Suppl. 1, 66.
Ball, R.O., Law, G., Bertolo, R.F.P., Pencharz, P.B., 1999. Adequate oral threonine is critical for mucin
production and mucosal growth by the neonatal piglet gut. In: Lobley, G.E., White, A., MacRae, J.C.
(Eds.), Protein Metabolism and Nutrition: Book of Abstracts of the VIIIth International Symposium.
Wageningen Press, Wageningen, The Netherlands, p. 31.
Ballevre, O., Cadenhead, A., Calder, A.G., Rees, W.D., Lobley, G.E., Fuller, M.F., Garlick, P.J., 1990.
Quantitative partition of threonine oxidation in pigs: effect of dietary threonine. Amer. J. Physiol. 259,
E483−E491.
150 R. F. P. Bertolo et al.

Basile-Filho, A., Beaumier, L., el-Khoury, A.E., Yu, Y.M., Kenneway, M., Gleason, R.E., Young, V.R.,
1998. Twenty-four-hour L-[1-13C]tyrosine and L-[3,3-2H2]phenylalanine oral tracer studies at generous,
intermediate, and low phenylalanine intakes to estimate aromatic amino acid requirements in adults.
Amer. J. Clin. Nutr. 67, 640−659.
Basile-Filho, A., el-Khoury, A.E., Beaumier, L., Wang, S.Y., Young, V.R., 1997. Continuous 24-h L-[1-13C]
phenylalanine and L-[3,3-2H2]tyrosine oral-tracer studies at an intermediate phenylalanine intake to
estimate requirements in adults. Amer. J. Clin. Nutr. 65, 473−488.
Baumann, P.Q., Stirewalt, W.S., O’Rourke, B.D., Howard, D., Nair, K.S., 1994. Precursor pools of pro-
tein synthesis: a stable isotope study in a swine model. Amer. J. Physiol. 267, E203−E209.
Bennet, W.M., Haymond, M.W., 1991. Plasma pool source for fibrinogen synthesis in postabsorptive
conscious dogs. Amer. J. Physiol. 260, E581−E587.
Bertolo, R.F.P., Brunton, J.A., Pencharz, P.B., Ball, R.O., 2000a. Steady state is not achieved for most
plasma amino acids during 12 hours of fasting in the neonatal piglet. Pediat. Res. 48, 701−707.
Bertolo, R.F.P., Chen, C.Z.L., Law, G., Pencharz, P.B., Ball, R.O., 1998. Threonine requirement of neonatal
piglets receiving total parenteral nutrition is considerably lower than that of piglets receiving an iden-
tical diet intragastrically. J. Nutr. 128, 1752−1759.
Bertolo, R.F.P., Chen, C.Z.L., Pencharz, P.B., Ball, R.O., 1999. Intestinal atrophy has a greater impact on
nitrogen metabolism than liver by-pass in piglets fed identical diets via gastric, central venous or
portal venous routes. J. Nutr. 129, 1045−1052.
Bertolo, R.F.P., Moehn, S., Pencharz, P.B., Ball, R.O., 2001. Adaptation of indicator amino acid oxidation
occurs within 2 days in 25 or 250 kg pigs fed varying levels of lysine or protein. FASEB J. 15, A266.
Bertolo, R.F.P., Pencharz, P.B., Ball, R.O., 2000b. Organ and plasma amino acid concentrations are
profoundly different in piglets fed identical diets via gastric, central venous or portal venous routes.
J. Nutr. 130, 1261−1266.
Biolo, G., Tessari, P., Inchiostro, S., Bruttomesso, D., Fongher, C., Sabadin, L., Fratton, M.G., Valerio, A.,
Tiengo, A., 1992. Leucine and phenylalanine kinetics during mixed meal ingestion: a multiple tracer
approach. Amer. J. Physiol. 262, E455−E463.
Borman, A., Wood, T.R., Black, H.C., Anderson, E.G., Osterling, M.J., Womack, M., Rose, W.C., 1946.
The role of arginine in growth with some observations on the effect of argininic acid. J. Biol. Chem.
166, 585−594.
Brookes, I.M., Owens, F.N., Garrigus, U.S., 1972. Influence of amino acid level in the diet upon amino
acid oxidation by the rat. J. Nutr. 102, 27−35.
Bross, R., Ball, R.O., Clarke, J.T., Pencharz, P.B., 2000. Tyrosine requirements in children with classical PKU
determined by indicator amino acid oxidation. Amer. J. Physiol. Endocrinol. Metab. 278, E195−E201.
Bross, R., Ball, R.O., Pencharz, P.B., 1998. Development of a minimally invasive protocol for the
determination of phenylalanine and lysine kinetics in humans during the fed state. J. Nutr. 128,
1913−1919.
Bruins, M.J., Soeters, P.B., Deutz, N.E., 2000. Endotoxemia affects organ protein metabolism differently
during prolonged feeding in pigs. J. Nutr. 130, 3003−3013.
Brunton, J.A., Ball, R.O., Pencharz, P.B., 1998. Determination of amino acid requirements by
indicator amino acid oxidation: applications in health and disease. Curr. Opin. Clin. Nutr. Metab. Care
1, 449−453.
Brunton, J.A., Bertolo, R.F.P., Pencharz, P.B., Ball, R.O., 1999. Proline ameliorates arginine deficiency
during enteral but not parenteral feeding in neonatal piglets. Amer. J. Physiol. 277, E223−E231.
Burrin, D.G., Ferrell, C.L., Britton, R.A., Bauer, M., 1990. Level of nutrition and visceral organ size and
metabolic activity in sheep. Brit. J. Nutr. 64, 439−448.
Burrin, D.G., Stoll, B., Chang, X., Yu, H., Reeds, P.J., 1999. Total parenteral nutrition does not compro-
mise digestion or absorption of dietary protein in neonatal pigs. FASEB J. 13, A1024.
Burrin, D.G., Stoll, B., Jiang, R., Chang, X., Hartmann, B., Holst, J.J., Greeley, G.H. Jr., Reeds, P.J., 2000.
Minimal enteral nutrient requirements for intestinal growth in neonatal piglets: how much is enough?
Amer. J. Clin. Nutr. 71, 1603−1610.
Cappelli, F.P., Seal, C.J., Parker, D.S., 1997. Glucose and [13C]leucine metabolism by the portal-drained
viscera of sheep fed on dried grass with acute intravenous and intraduodenal infusions of glucose.
Brit. J. Nutr. 78, 931−946.
Caso, G., Ford, G.C., Nair, K.S., Vosswinkel, J.A., Garlick, P.J., McNurlan, M.A., 2001. Increased con-
centration of tracee affects estimates of muscle protein synthesis. Amer. J. Physiol. Endocrinol.
Metab. 280, E937−E946.
Role of intestinal first-pass metabolism 151

Cayol, M., Boirie, Y., Prugnaud, J., Gachon, P., Beaufrere, B., Obled, C., 1996. Precursor pool for hepatic
protein synthesis in humans: effects of tracer route infusion and dietary proteins. Amer. J. Physiol.
270, E980−E987.
Chavez, E.R., Bayley, H.S., 1976. The oxidation in vivo of uniformly and carboxyl-14C-labelled methio-
nine by young pigs. Proc. Nutr. Soc. 35, 150A−151A.
Chinkes, D.L., Rosenblatt, J., Wolfe, R.R., 1993. Assessment of the mathematical issues involved in
measuring the fractional synthesis rate of protein using the flooding dose technique. Clin. Sci. (Lond.)
84, 177−183.
Cobelli, C., Saccomani, M.P., Tessari, P., Biolo, G., Luzi, L., Matthews, D.E., 1991. Compartmental
model of leucine kinetics in humans. Amer. J. Physiol. 261, E539−E550.
Coleman, R.A., Korver, D.R., Bertolo, R.F., Ball, R.O., 2002. Lysine requirements of pre-lay broiler
breeders as determined by the indicator amino acid oxidation method. Presented at Poultry Science
Association 91st Annual Meeting, Newark, Delaware.
Cvitkovic, S., Pencharz, P.B., Ball, R.O., 2000. Comparison of oral and intravenous isotopic tracers in
determining tryptophan requirements of piglets. FASEB J. 14, A745.
Deshmukh, D.R., Shope, T.C., 1983. Arginine requirement and ammonia toxicity in ferrets. J. Nutr. 113,
1664−1667.
Deutz, N.E., Bruins, M.J., Soeters, P.B., 1998. Infusion of soy and casein protein meals affects inter-
organ amino acid metabolism and urea kinetics differently in pigs. J. Nutr. 128, 2435−2445.
Di Buono, M., Wykes, L.J., Ball, R.O., Pencharz, P.B., 2001a. Total sulfur amino acid requirement in
young men as determined by indicator amino acid oxidation with L-[1-13C]phenylalanine. Amer. J. Clin.
Nutr. 74, 756−760.
Di Buono, M., Wykes, L.J., Ball, R.O., Pencharz, P.B., 2001b. Dietary cysteine reduces the methionine
requirement in men. Amer. J. Clin. Nutr. 74, 761−766.
Duffy, B., Pencharz, P.B., 1986. The effect of feeding route (IV or oral) on the protein metabolism of the
neonate. Amer. J. Clin. Nutr. 43, 108−111.
Duncan, A.M., Ball, R.O., Pencharz, P.B., 1996. Lysine requirement of adult males is not affected by
decreasing dietary protein. Amer. J. Clin. Nutr. 64, 718−725.
Elango, R., Pencharz, P.B., Ball, R.O., 2002a. The branched-chain amino acid requirement of parenter-
ally fed neonatal piglets is less than the enteral requirement. J. Nutr. 132, 3123−3129.
Elango, R., Pencharz, P.B., Ball, R.O., 2002b. The optimum ratio of branched-chain amino acids (BCAA)
differs between intragastric and intravenous routes of feeding in the neonatal piglet. FASEB J.
16, A746.
el-Khoury, A.E., Basile, A., Beaumier, L., Wang, S.Y., Al-Amiri, H.A., Selvaraj, A., Wong, S., Atkinson, A.,
Ajami, A.M., Young, V.R., 1998. Twenty-four-hour intravenous and oral tracer studies with
L-[1-13C]-2-aminoadipic acid and L-[1-13C]lysine as tracers at generous nitrogen and lysine intakes in
healthy adults. Amer. J. Clin. Nutr. 68, 827−839.
el-Khoury, A.E., Fukagawa, N.K., Sanchez, M., Tsay, R.H., Gleason, R.E., Chapman, T.E., Young, V.R.,
1994a. Validation of the tracer-balance concept with reference to leucine: 24-h intravenous tracer
studies with L-[1-13C]leucine and [15N-15N]urea. Amer. J. Clin. Nutr. 59, 1000−1011.
el-Khoury, A.E., Fukagawa, N.K., Sanchez, M., Tsay, R.H., Gleason, R.E., Chapman, T.E., Young, V.R.,
1994b. The 24-h pattern and rate of leucine oxidation, with particular reference to tracer estimates of
leucine requirements in healthy adults. Amer. J. Clin. Nutr. 59, 1012−1020.
el-Khoury, A.E., Pereira, P.C., Borgonha, S., Basile-Filho, A., Beaumier, L., Wang, S.Y., Metges, C.C.,
Ajami, A.M., Young, V.R., 2000. Twenty-four-hour oral tracer studies with L-[1-13C]lysine at a low
(15 mg kg−1.d−1) and intermediate (29 mg kg−1.d−1)) lysine intake in healthy adults. Amer. J. Clin. Nutr.
72, 122−130.
el-Khoury, A.E., Sanchez, M., Fukagawa, N.K., Gleason, R.E., Tsay, R.H., Young, V.R., 1995. The 24-h
kinetics of leucine oxidation in healthy adults receiving a generous leucine intake via three discrete
meals. Amer. J. Clin. Nutr. 62, 579−590.
FAO/WHO/UNU Expert Consultation, 1985. Energy and protein requirements. World Health
Organization, Geneva. Tech. Rep. Ser. No. 724.
Fuller, M.F., Garlick, P.J., 1994. Human amino acid requirements: can the controversy be resolved?
Annu. Rev. Nutr. 14, 217−241.
Gaudichon, C., Mahe, S., Benamouzig, R., Luengo, C., Fouillet, H., Dare, S., Van Oycke, M., Ferriere, F.,
Rautureau, J., Tome, D., 1999. Net postprandial utilization of [15N]-labeled milk protein nitrogen is
influenced by diet composition in humans. J. Nutr. 129, 890−895.
152 R. F. P. Bertolo et al.

Gelfand, R.A., Glickman, M.G., Castellino, P., Louard, R.J., DeFronzo, R.A., 1988. Measurement of
L-[1-14C]leucine kinetics in splanchnic and leg tissues in humans: effect of amino acid infusion.
Diabetes 37, 1365−1372.
Goldstein, R.M., Hebiguchi, T., Luk, G.D., Taqi, F., Guilarte, T.R., Franklin, F.A., Niemiec, P.W.,
Dudgeon, D.L., 1985. The effects of total parenteral nutrition on gastrointestinal growth and devel-
opment. J. Pediat. Surg. 20, 785−791.
Harper, A.E., Benjamin, E., 1984. Relationship between intake and rate of oxidation of leucine and
α-ketoisocaproate in vivo in the rat. J. Nutr. 114, 431-440.
Heird, W.C., Nicholson, J.F., Driscoll, J.M., Schullinger, J.N., Winters, R.W., 1972. Hyperammonemia
resulting from intravenous alimentation using a mixture of synthetic L-amino acids: a preliminary
report. J. Pediat. 81, 162−165.
Higashiguchi, T., Noguchi, Y., Noffsinger, A., Fischer, J.E., Hasselgren, P.O., 1994a. Sepsis increases
production of total secreted proteins, vasoactive intestinal peptide, and peptide YY in isolated rat ente-
rocytes. Amer. J. Surg. 168, 251−256.
Higashiguchi, T., Noguchi, Y., O’Brien, W., Wagner, K., Fischer, J.E., Hasselgren, P.O., 1994b. Effect of
sepsis on mucosal protein synthesis in different parts of the gastrointestinal tract in rats. Clin. Sci.
(Lond.) 87, 207−211.
Hoerr, R.A., Matthews, D.E., Bier, D.M., Young, V.R., 1991. Leucine kinetics from [2H3]- and
[13C]leucine infused simultaneously by gut and vein. Amer. J. Physiol. 260, E111−E117.
Hoerr, R.A., Matthews, D.E., Bier, D.M., Young, V.R., 1993. Effects of protein restriction and acute
refeeding on leucine and lysine kinetics in young men. Amer. J. Physiol. 264, E567−E575.
House, J.D., Pencharz, P.B., Ball, R.O., 1997a. Tyrosine kinetics and requirements during total parenteral
nutrition in the neonatal piglet: the effect of glycyl-L-tyrosine supplementation. Pediat. Res. 41, 575−583.
House, J.D., Pencharz, P.B., Ball, R.O., 1997b. Phenylalanine requirements determined by using
L-[1-14C]phenylalanine in neonatal piglets receiving total parenteral nutrition supplemented with tyro-
sine. Amer. J. Clin. Nutr. 65, 984−993.
House, J.D., Pencharz, P.B., Ball, R.O., 1998a. Lysine requirement of neonatal piglets receiving total par-
enteral nutrition as determined by oxidation of the indicator amino acid L-[1-14C]phenylalanine.
Amer. J. Clin. Nutr. 67, 67−73.
House, J.D., Thorpe, J.M., Wykes, L.J., Pencharz, P.B., Ball, R.O., 1998b. Evidence that phenylalanine
hydroxylation rates are overestimated in neonatal subjects receiving total parenteral nutrition with a
high phenylalanine content. Pediat. Res. 43, 461−466.
Jeevanandam, M., Lowry, S.F., Brennan, M.F., 1987. Effect of the route of nutrient administration on
whole-body protein kinetics in man. Metabolism 36, 968−973.
Johnson, L.R., Copeland, E.M., Dudrick, S.J., Lichtenberger, L.M., Castro, G.A., 1975. Structural and hor-
monal alterations in the gastrointestinal tract of parenterally fed rats. Gastroenterology 68, 1177−1183.
Kang-Lee, Y.A., Harper, A.E., 1977. Effect of histidine intake and hepatic histidase activity on the metab-
olism of histidine in vivo. J. Nutr. 107, 1427−1443.
Kang-Lee, Y.A., Harper, A.E., 1978. Threonine metabolism in vivo: effect of threonine intake and prior
induction of threonine dehydratase in rats. J. Nutr. 108, 163−175.
Kim, K.I., Bayley, H.S., 1983. Amino acid oxidation by young pigs receiving diets with varying levels of
sulphur amino acids. Brit. J. Nutr. 50, 383−390.
Kim, K.I., Elliott, J.I., Bayley, H.S., 1983a. Oxidation of an indicator amino acid by young pigs receiv-
ing diets with varying levels of lysine or threonine, and an assessment of amino acid requirements.
Brit. J. Nutr. 50, 391−399.
Kim, K.I., McMillan, I., Bayley, H.S., 1983b. Determination of amino acid requirements of young pigs
using an indicator amino acid. Brit. J. Nutr. 50, 369−382.
Klasing, K.C., Calvert, C.C., 1999. The care and feeding of an immune system: an analysis of lysine
needs. In: Lobley, G.E., White, A., MacRae, J.C. (Eds.), Protein Metabolism and Nutrition:
Proceedings of the VIIIth International Symposium. Wageningen Press, Wageningen, The
Netherlands, pp. 253−264.
Kriengsinyos, W., Wykes, L.J., Ball, R.O., Pencharz, P.B., 2002. Effect of oral and intravenous tracer proto-
cols on the estimate of lysine requirement in healthy adult males L-[1-13C]phenylalanine as an indicator
amino acid. J. Nutr. 132, 2251–2257.
Kurpad, A.V., Raj, T., el-Khoury, A., Beaumier, L., Kuriyan, R., Srivatsa, A., Borgonha, S., Selvaraj, A.,
Regan, M.M., Young, V.R., 2001. Lysine requirements of healthy adult Indian subjects, measured by
an indicator amino acid balance technique. Amer. J. Clin. Nutr. 73, 900−907.
Role of intestinal first-pass metabolism 153

Kurpad, A.V., Raj, T., Regan, M.M., Vasudevan, J., Caszo, B., Nazareth, D., Gnanou, J., Young, V.R.,
2002a. Threonine requirements of healthy Indian men, measured by a 24-h indicator amino acid
oxidation and balance technique. Amer. J. Clin. Nutr. 76, 789−797.
Kurpad, A.V., Regan, M.M., Raj, T., el-Khoury, A., Kuriyan, R., Vaz, M., Chandakudlu, D.,
Venkataswamy, V.G., Borgonha, S., Young, V.R., 2002b. Lysine requirements of healthy adult Indian
subjects receiving long-term feeding, measured with a 24-h indicator amino acid oxidation and
balance technique. Amer. J. Clin. Nutr. 76, 404−412.
Lamont, J., 1992. Mucus: the front line of intestinal mucosal defense. Ann. N.Y. Acad. Sci. 664, 190−201.
Lanza-Jacoby, S., Sitren, H.S., Stevenson, N.R., Rosato, F.E., 1982. Changes in circadian rhythmicity of
liver and serum parameters in rats fed a total parenteral nutrition solution by continuous and discon-
tinuous intravenous or intragastric infusion. J. Parent. Enter. Nutr. 6, 496−502.
Lazaris-Brunner, G., Rafii, M., Ball, R.O., Pencharz, P.B., 1998. Tryptophan requirement in young adult
women as determined by indicator amino acid oxidation with L-[13C]phenylalanine. Amer. J. Clin.
Nutr. 68, 303−310.
Lewis, A.J., 1992. Determination of the amino acid requirements of animals. In: Nissen, S. (Ed.), Modern
Methods in Protein Nutrition and Metabolism. Academic Press, San Diego, pp. 67−86.
Lobley, G.E., Milne, V., Lovie, J.M., Reeds, P.J., Pennie, K., 1980. whole-body and tissue protein
synthesis in cattle. Brit. J. Nutr. 43, 491−502.
MacRae, J.C., Bruce, L.A., Brown, D.S., Calder, A.G., 1997. Amino acid use by the gastrointestinal tract
of sheep given lucerne forage. Amer. J. Physiol. 273, G1200−G1207.
Mager, D.R., Ball, R.O., Pencharz, P.B., 2002. The effect of orthotopic liver transplantation (OLT) on
total branched-chain amino acid requirements in children as determined by indicator amino acid
oxidation (IAAO). FASEB J. 16, A259.
Mager, D.R., Pencharz, P.B., Ball, R.O., 2001. Branchedchain amino acid requirements in school aged
children determined by indicator amino acid oxidation (IAAO). FASEB J. 15, A266.
Mantle, M., Allen, A., 1981. Isolation and characterization of the native glycoprotein from pig small-
intestinal mucus. Biochem J. 195, 267−275.
Mariotti, F., Mahe, S., Benamouzig, R., Luengo, C., Dare, S., Gaudichon, C., Tome, D., 1999. Nutritional
value of [15N]-soy protein isolate assessed from ileal digestibility and postprandial protein utilization
in humans. J. Nutr. 129, 1992−1997.
Martensson, J., Jain, A., Meister, A., 1990. Glutathione is required for intestinal function. Proc. Natl.
Acad. Sci. USA 87, 1715−1719.
Matthews, D.E., Marano, M.A., Campbell, R.G., 1993. Splanchnic bed utilization of leucine and phe-
nylalanine in humans. Amer. J. Physiol. 264, E109−E118.
Mavromichalis, I., Webel, D.M., Emmert, J.L., Moser, R.L., Baker, D.H., 1998. Limiting order of amino
acids in a low-protein corn-soybean meal-whey-based diet for nursery pigs. J. Anim. Sci. 76, 2833−2837.
McNurlan, M.A., Garlick, P.J., 1980. Contribution of rat liver and gastrointestinal tract to whole-body
protein synthesis in the rat. Biochem. J. 186, 381−383.
Meguid, M.M., Matthews, D.E., Bier, D.M., Meredith, C.N., Soeldner, J.S., Young, V.R., 1986a. Leucine
kinetics at graded leucine intakes in young men. Amer. J. Clin. Nutr. 43, 770−780.
Meguid, M.M., Matthews, D.E., Bier, D.M., Meredith, C.N., Young, V.R., 1986b. Valine kinetics at
graded valine intakes in young men. Amer. J. Clin. Nutr. 43, 781−786.
Meredith, C.N., Wen, Z.M., Bier, D.M., Matthews, D.E., Young, V.R., 1986. Lysine kinetics at graded
lysine intakes in young men. Amer. J. Clin. Nutr. 43, 787−794.
Mertz, E.T., Beeson, W.M., Jackson, H.D., 1952. Classification of essential amino acids for the weanling
pigs. Arch. Biochem. Biophys. 38, 121−128.
Moehn, S., Bertolo, R.F.P., Ball, R.O., 2001. A method to measure the amino acid requirement of indi-
vidual pigs. J. Anim. Sci. 79, Suppl. 1, 66.
More, J., Fioramonti, J., Benazet, F., Bueno, L., 1987. Histochemical characterization of glycoproteins
present in jejunal and colonic goblet cells of pigs on different diets. Histochemistry 87, 189−194.
Morris, J.G., Rogers, Q.R., 1978. Arginine: an essential amino acid for the cat. J. Nutr. 108, 1944−1953.
Morris, S.M., 1992. Regulation of enzymes of urea and arginine synthesis. Annu. Rev. Nutr. 12, 81−101.
Motil, K.J., Opekun, A.R., Montandon, C.M., Berthold, H.K., Davis, T.A., Klein, P.D., Reeds, P.J., 1994.
Leucine oxidation changes rapidly after dietary protein intake is altered in adult women but lysine
flux is unchanged as is lysine incorporation into VLDL-apolipoprotein B-100. J. Nutr. 124, 41−51.
Murphy, J.M., Murch, S.J., Ball, R.O., 1996. Proline is synthesized from glutamate during intragastric
infusion but not during intravenous infusion in neonatal piglets. J. Nutr. 126, 878−886.
154 R. F. P. Bertolo et al.

Myrie, S.B., Bertolo, R.F.P., Sauer, W.C., Ball, R.O., 2002. Will anti-nutritive factors in foods stimulate
increased intestinal mucin secretion and increase threonine requirement? FASEB J. 16, A260.
NRC (National Research Council), 1979. Nutrient Requirements of Swine (8th Edition). National
Academy Press, Washington, DC.
NRC (National Research Council), 1998. Nutrient Requirements of Swine (10th Edition). National
Academy Press, Washington, DC.
Newport, M.J., Chavez, E.R., Horney, F.D., Bayley, H.S., 1976. Amino acid metabolism in the piglet:
influence of level of protein and of methionine in the diet on tissue uptake and in vivo oxidation.
Brit. J. Nutr. 36, 87−99.
Raguso, C.A., Pereira, P., Young, V.R., 1999. A tracer investigation of obligatory oxidative amino acid
losses in healthy, young adults. Amer. J. Clin. Nutr. 70, 474−483.
Rand, W.M., Young, V.R., 1999. Statistical analysis of nitrogen balance data with reference to the lysine
requirement in adults. J. Nutr. 129, 1920−1926.
Reeds, P.J., Burrin, D.G., Stoll, B., Jahoor, F., 2000. Intestinal glutamate metabolism. J. Nutr. 130,
978S−982S.
Reeds, P.J., Hachey, D.L., Patterson, B.W., Motil, K.J., Klein, P.D., 1992. VLDL apolipoprotein B-100,
a potential indicator of the isotopic labeling of the hepatic protein synthetic precursor pool in humans:
studies with multiple stable isotopically labeled amino acids. J. Nutr. 122, 457−466.
Riazi, R., Ball, R.O., Pencharz, P.B., 2001. Total branched-chain amino acid requirement in young
healthy adult men determined by indicator amino acid oxidation using L-[1-13C]phenylalanine.
FASEB J. 15, A266.
Roberts, S.A., Ball, R.O., Moore, A.M., Filler, R.M., Pencharz, P.B., 2001a. The effect of graded intake
of glycyl-L-tyrosine on phenylalanine and tyrosine metabolism in parenterally fed neonates with an
estimation of tyrosine requirement. Pediat. Res. 49, 111−119.
Roberts, S.A., Thorpe, J.M., Ball, R.O., Pencharz, P.B., 2001b. Tyrosine requirement of healthy
men receiving a fixed phenylalanine intake determined by using indicator amino acid oxidation.
Amer. J. Clin. Nutr. 73, 276−282.
Sanchez, M., el-Khoury, A.E., Castillo, L., Chapman, T.E., Filho, A.B., Beaumier, L., Young, V.R.,
1996. Twenty-four-hour intravenous and oral tracer studies with L-[1-13C]phenylalanine and
L-[3,3-2H2]tyrosine at a tyrosine-free, generous phenylalanine intake in adults. Amer. J. Clin. Nutr. 63,
532−545.
Sanchez, M., el-Khoury, A.E., Castillo, L., Chapman, T.E., Young, V.R., 1995. Phenylalanine and
tyrosine kinetics in young men throughout a continuous 24-h period, at a low phenylalanine intake.
Amer. J. Clin. Nutr. 61, 555−570.
Sharma, R., Schumacher, U., 1995. The influence of diets and gut microflora on lectin binding patterns
of intestinal mucins in rats. Lab. Invest. 73, 558−564.
Shoveller, A.K., Brunton, J.A., Bertolo, R.F.P., Pencharz, P.B., Ball, R.O., 2001. Intestinal metabolism of
methionine significantly affects requirement and proportion spared by cysteine. In: Lindberg, J.E.,
Ogle, B. (Eds.), Digestive Physiology of Pigs: Proceedings of the 8th Symposium. CABI Publishing,
New York, pp. 89−91.
Sim, A.J., Wolfe, B.M., Young, V.R., Clarke, D., Moore, F.D., 1979. Glucose promotes whole-body
protein synthesis from infused aminoacids in fasting man: isotopic demonstration. Lancet 1
(8107), 68−72.
Simon, O., Adam, K., Bergner, H., 1978. [Metabolism-oriented lysine requirement of mature rats based
on the catabolism rate for 14C- and 15N-labelled lysine]. Archiv. Tierernähr. 28, 609−617.
Snyderman, S.E., 1971. The protein and amino acid requirements of the premature infant. In:
Jonxis, J.H.P., Visser, H.K.A., Troelstra, J.A. (Eds.), Metabolic Processes in the Foetus and Newborn.
Stenfert Kroese, Leiden, The Netherlands, pp. 128−141.
Specian, R.D., Oliver, M.G., 1991. Functional biology of intestinal goblet cells. Amer. J. Physiol. 260,
C183−C193.
Stoll, B., Burrin, D.G., Henry, J., Jahoor, F., Reeds, P.J., 1997. Phenylalanine utilization by the gut and
liver measured with intravenous and intragastric tracers in pigs. Amer. J. Physiol. 273, G1208−G1217.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1999. Substrate oxidation by the portal
drained viscera of fed piglets. Amer. J. Physiol. 277, E168−E175.
Stoll, B., Henry, J., Reeds, P.J., Yu, H., Jahoor, F., Burrin, D.G., 1998. Catabolism dominates the first-
pass intestinal metabolism of dietary essential amino acids in milk protein-fed piglets. J. Nutr. 128,
606−614.
Role of intestinal first-pass metabolism 155

Storch, K.J., Wagner, D.A., Burke, J.F., Young, V.R., 1988. Quantitative study in vivo of methionine cycle
in humans using [methyl-2H3]- and [1-13C]methionine. Amer. J. Physiol. 255, E322−E331.
Tamir, H., Ratner, S., 1963. A study of ornithine, citrulline and arginine synthesis in growing chicks.
Arch. Biochem. Biophys. 102, 259−269.
Thorpe, J.M., Roberts, S.A., Ball, R.O., Pencharz, P.B., 2000. Effect of tyrosine intake on the rate of
phenylalanine hydroxylation in adult males. Metabolism 49, 444−449.
Tome, D., Bos, C., 2000. Dietary protein and nitrogen utilization. J. Nutr. 130, 1868S−1873S.
van der Meulen, J., Bakker, J.G., Smits, B., de Visser, H., 1997. Effects of source of starch on net portal
flux of glucose, lactate, volatile fatty acids and amino acids in the pig. Brit. J. Nutr. 78, 533−544.
van Goudoever, J.B., Stoll, B., Henry, J.F., Burrin, D.G., Reeds, P.J., 2000. Adaptive regulation of intes-
tinal lysine metabolism. Proc. Natl. Acad. Sci. USA 97, 11620−11625.
Visek, W.J., 1984. An update of concepts of essential amino acids. Annu. Rev. Nutr. 4, 137−155.
von Allmen, D., Hasselgren, P.O., Higashiguchi, T., Frederick, J., Zamir, O., Fischer, J.E., 1992.
Increased intestinal protein synthesis during sepsis and following the administration of tumour necrosis
factor α or interleukin-1α. Biochem. J. 286, 585−589.
Wakabayashi, Y., Yamada, E., Yoshida, T., Takahashi, N., 1995. Effect of intestinal resection and arginine-
free diet on rat physiology. Amer. J. Physiol. 269, G313−G318.
Waterlow, J.C., 1985. What do we mean by adaptation? In: Blaxter, K.L., Waterlow, J.C. (Eds.),
Nutritional Adaptation in Man. John Libbey, London, pp. 1−10.
Waterlow, J.C., 1999. The mysteries of nitrogen balance. Nutr. Res. Rev. 12, 25−54.
Waterlow, J.C., Garlick, P.J., Millward, D.J., 1978. Protein Turnover in Mammalian Tissues and in the
Whole Body. Elsevier/North-Holland Biomedical Press, Amsterdam.
Watt, P.W., Lindsay, Y., Scrimgeour, C.M., Chien, P.A., Gibson, J.N., Taylor, D.J., Rennie, M.J., 1991.
Isolation of aminoacyl-tRNA and its labeling with stable-isotope tracers: use in studies of human
tissue protein synthesis. Proc. Natl. Acad. Sci. USA 88, 5892−5896.
Were, K.E., 1989. Determination of the tryptophan requirement of the juvenile rainbow trout using the
indicator oxidation technique. M.Sc. Thesis, University of Guelph, Guelph, Canada.
Wilson, D.C., Rafii, M., Ball, R.O., Pencharz, P.B., 2000. Threonine requirement of young men
determined by indicator amino acid oxidation with use of L-[1-13C]phenylalanine. Amer. J. Clin. Nutr.
71, 757−764.
Windmueller, H.G., Spaeth, A.E., 1980. Respiratory fuels and nitrogen metabolism in vivo in small intes-
tine of fed rats: quantitative importance of glutamine, glutamate, and aspartate. J. Biol. Chem. 255,
107−112.
Wolfe, R.R., 1992. Radioactive and Stable Isotope Tracers in Biomedicine: Principles and Practice of
Kinetic Analysis. Wiley-Liss, Toronto, pp. 357−376.
Wu, G., 1998. Amino acid metabolism in the small intestine. Trends Comp. Biochem. Physiol. 4, 39−74.
Wu, G., Borbolla, A.G., Knabe, D.A., 1994. The uptake of glutamine and release of arginine, citrulline
and proline by the small intestine of developing pigs. J. Nutr. 124, 2437−2444.
Wykes, L.J., Ball, R.O., Pencharz, P.B., 1993. Development and validation of a total parenteral nutrition
model in the neonatal piglet. J. Nutr. 123, 1248−1259.
Young, V.R., Bier, D.M., Pellett, P.L., 1989. A theoretical basis for increasing current estimates of the
amino acid requirements in adult man, with experimental support. Amer. J. Clin. Nutr. 50, 80−92.
Young, V.R., Borgonha, S., 2000. Nitrogen and amino acid requirements: the Massachusetts Institute of
Technology amino acid requirement pattern. J. Nutr. 130, 1841S−1849S.
Young, V.R., el-Khoury, A.E., 1995. Can amino acid requirements for nutritional maintenance in adult
humans be approximated from the amino acid composition of body mixed proteins? Proc. Natl. Acad.
Sci. USA. 92, 300−304.
Young, V.R., Marchini, J.S., 1990. Mechanisms and nutritional significance of metabolic responses to
altered intakes of protein and amino acids, with reference to nutritional adaptation in humans.
Amer. J. Clin. Nutr. 51, 270−289.
Yu, F., Bruce, L.A., Calder, A.G., Milne, E., Coop, R.L., Jackson, F., Horgan, G.W., MacRae, J.C.,
2000. Subclinical infection with the nematode Trichostrongylus colubriformis increases gastroin-
testinal tract leucine metabolism and reduces availability of leucine for other tissues. J. Anim. Sci.
78, 380−390.
Yu, Y.M., Burke, J.F., Vogt, J.A., Chambers, L., Young, V.R., 1992. Splanchnic and whole body
L-[1-13C,15N]leucine kinetics in relation to enteral and parenteral amino acid supply. Amer. J. Physiol.
262, E687−E694.
156 R. F. P. Bertolo et al.

Yu, Y.M., Wagner, D.A., Tredget, E.E., Walaszewski, J.A., Burke, J.F., Young, V.R., 1990. Quantitative
role of splanchnic region in leucine metabolism: L-[1-13C,15N]leucine and substrate balance studies.
Amer. J. Physiol. 259, E36−E51.
Yu, Y.M., Young, V.R., Tompkins, R.G., Burke, J.F., 1995. Comparative evaluation of the quantitative
utilization of parenterally and enterally administered leucine and L-[1-13C,15N]leucine within the
whole-body and the splanchnic region. J. Parent. Enter. Nutr. 19, 209−215.
Zello, G.A., Pencharz, P.B., Ball, R.O., 1990. Phenylalanine flux, oxidation, and conversion to tyrosine
in humans studied with L-[1-13C]phenylalanine. Amer. J. Physiol. 259, E835−E843.
Zello, G.A., Pencharz, P.B., Ball, R.O., 1993. Dietary lysine requirement of young adult males deter-
mined by oxidation of L-[1-13C]phenylalanine. Amer. J. Physiol. 264, E677−E685.
Zello, G.A., Wykes, L.J., Ball, R.O., Pencharz, P.B., 1995. Recent advances in methods of assessing
dietary amino acid requirements for adult humans. J. Nutr. 125, 2907−2915.
Zhao, X.H., Wen, Z.M., Meredith, C.N., Matthews, D.E., Bier, D.M., Young, V.R., 1986. Threonine kinetics
at graded threonine intakes in young men. Amer. J. Clin. Nutr. 43, 795−802.
7 Splanchnic protein and amino acid
metabolism in growing animals1

D. G. Burrin and B. Stoll

USDA/ARS Children’s Nutrition Research Center, Department of Pediatrics,


Baylor College of Medicine, 1100 Bates Street, Houston, TX 77030, USA

The splanchnic tissues, namely liver and gut, play a major role in the regulation of whole-
body protein and amino acid metabolism. Given their anatomical design for assimilation of
food by the host, these tissues metabolize in first-pass a significant proportion of the dietary
amino acids via protein synthesis and oxidation and thereby limit the quantity and alter the
pattern of amino acids for systemic availability. This substantial “metabolic cost” incurred by
splanchnic tissues is in large part related to the numerous critical physiological functions they
perform for the mammalian host, such as digestion, ureagenesis, gluconeogensis, and acute-
phase protein synthesis. Splanchnic tissues also play a key regulatory role by transmitting
endocrine, immune, and neural signals in response to the diet and environment, which in turn
determine the rates of peripheral tissue protein metabolism and growth. Splanchnic tissue
protein metabolism is regulated by specific amino acids and hormones. Amino acids function
not only as substrates, but also as extracellular signals that influence cell functions, such as
protein turnover, proliferation, apoptosis, cell volume, and redox status. Many of the intra-
cellular signaling and biochemical pathways involved in amino acid metabolism have been
described in the liver, but less is known about the gut tissues. The beneficial health effects of
key immunonutrients are mediated by improved cell functions in different splanchnic tissues.

1. INTRODUCTION
Numerous studies with growing mammals, ruminant and nonruminant, have established
that the splanchnic tissues have a substantial impact on whole-body protein and amino

1 The authors thank Jane Schoppe for her assistance in the preparation of this manuscript. This work is a publication of
the USDA/ARS Children’s Nutrition Research Center, Department of Pediatrics, Baylor College of Medicine and Texas
Children’s Hospital, Houston, Texas. The work was supported in part by federal funds from the U.S. Department of
Agriculture Agricultural Research Service, Cooperative Agreement No. 58-6258-6001, and by the National Institutes of
Health R01 HD33920. The contents of this publication do not necessarily reflect the views or policies of the U.S.
Department of Agriculture, nor does mention of trade names, commercial products, or organizations imply endorsement
by the U.S. Government.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
157 © 2005 Elsevier Limited. All rights reserved.
158 D. G. Burrin and B. Stoll

acid metabolism. The liver and portal-drained visceral (PDV) tissues, composed mainly of
gastrointestinal tissues, contribute roughly 10% of body protein, yet they account for 30–35%
of whole-body protein turnover and energy expenditure (Burrin et al., 1989; Yen et al., 1989;
Nieto and Lobley, 1999; Stoll et al., 1999a). Relative to their mass, the splanchnic tissues
exert a disproportionate impact on whole-body metabolism due to their relatively high frac-
tional rates of protein synthesis and oxygen consumption. Studies with domestic animals have
demonstrated that the fractional protein synthesis rates in the liver and intestinal tissues are
several-fold higher than that in peripheral tissues, such as muscle (Lobley et al., 1980, 1992;
Attaix et al., 1986; Burrin et al., 1992; Davis et al., 1996).
The inherently high metabolic rate of the splanchnic tissues is a function of several factors
related to the multitude of biological functions performed by these tissues and the relatively
high rate of cell turnover, particularly in the intestine. The gastrointestinal tract, including the
stomach, small and large intestine, and pancreas, has a primary function to digest and absorb
nutrients from the diet. In these processes, these tissues consume substantial amounts of amino
acids for synthesis of structural and secretory proteins and for oxidative energy necessary to
actively transport nutrients and replenish a continual loss of epithelial cells and digestive
enzymes. However, in addition, these tissues function as a physical and immunological barrier
to environmental pathogens and noxious substances, representing one of the largest lymphoid
tissues in the body. Compromise of this critical gut barrier can not only lead to a suppression
of growth, but jeopardize survival of the organism. The GI tract collectively is also a major
endocrine organ secreting dozens of peptide hormones that provide key signals for the metab-
olism and growth of the organism as a whole. The gastrointestinal (GI) tract is also extensively
enervated with its own intrinsic (enteric nervous system) as well as an extrinsic neural network,
which allows it to function autonomously or in concert with the central nervous system.
Some important examples of how these GI functions affect the host include the regulation
of food intake (e.g. peptide YY, ghrelin, glucagon-like peptide 1) and substrate homeostasis
(e.g. insulin and glucagon). From a protein metabolic perspective, the GI tract functions to
assimilate dietary protein in a chemical form (e.g. amino acids) that can be readily used by
all somatic cells and simultaneously communicate the availability of nutrients (e.g. insulin)
to enhance their utilization by peripheral tissues (e.g. skeletal muscle). The metabolic func-
tion of the liver is closely linked to the GI tract by acting as a buffer and scavenger of the
dietary nutrients and byproducts absorbed into the portal venous circulation, such as amino
acids, ammonia, and bile acids. The liver also plays a major role in the metabolism of dietary
amino acids that serve as carbon precursors for gluconeogenesis and the transfer of nitrogen
released from skeletal muscle amino acid catabolism. The liver is a central organ involved in
the production of acute-phase proteins in response to inflammation and infection. The intent
of this review is to provide an overview of protein and amino acid metabolism in splanchnic
tissues and highlight some of the recent advancements in our understanding of how these
tissues function in growing mammals.

2. METABOLIC FATE OF AMINO ACIDS


2.1. Anatomical and morphological considerations

There have been significant advancements in our understanding of splanchnic amino acid
metabolism in the last thirty years. In the early studies, much was learned about amino acid
metabolism from in vivo measurements of splanchnic organ balance. These pioneering studies
were originally derived based on measurements of the net difference in the concentration of
Splanchnic protein and amino acid metabolism 159

amino acids in arterial input and venous drainage and of blood flow (Elwyn et al., 1968; Wolff
et al., 1972; Felig, 1975). More detailed kinetic information of unidirectional amino acid fluxes
was subsequently obtained by adapting this approach to the use of radioactive, isotopic amino
acid tracers (Heitmann and Bergman, 1978) and more recently with stable isotopes of amino
acids, in humans, pigs, dogs, mice, sheep, and cattle (Hoerr et al., 1991; Yu et al., 1995; Lobley
et al., 1996a; Stoll et al., 1998; Lapierre et al., 1999; Hallemeesch et al., 2001). Other important
experimental approaches include in situ organ perfusion, which has been extensively used for
liver and intestinal amino acid metabolism (Windmueller and Spaeth, 1980; Haussinger, 1990).
However, because there is considerable cellular heterogeneity in the liver and gut tissues, studies
with isolated cells have also provided critical evidence for the cellular basis of amino acid
metabolism (Haussinger, 1990, 1996; Wu, 1998; van Sluijters et al., 2000; Meijer, 2003).
The metabolic fate of amino acids in the splanchnic tissues differs not only between the
liver and gastrointestinal tissues, but also among cells within each tissue bed. Even within the
liver or gut, the metabolic fate of amino acids is significantly affected by how they are pre-
sented to the tissue. For the purposes of this review, the starting point in the metabolic fate of
amino acids will be in the intestinal epithelial cell after transport from the gut lumen.
However, it is important to recognize that within the gastrointestinal tissues (e.g. PDV),
amino acids are presented via both the lumen and arterial circulation. From a quantitative per-
spective, the rate of input of most amino acids from the arterial circulation is substantially
greater (3–5-fold) than that from the diet (fig. 1). However, the fractional PDV utilization (i.e.
uptake/input) of dietary amino acids (ranging from 95% to 20%) is generally much greater
than amino acids derived from the arterial blood, ranging from approximately 5% to 15%.
In the liver, the metabolic fate of amino acids may vary depending on whether the input is
from the portal vein or hepatic artery. Studies in piglets suggest that after feeding, portal
rather than arterial phenylalanine is preferentially used for the synthesis of constitutive
and secretory hepatic proteins (Stoll et al., 1999b). Within a tissue, the morphological local-
ization of a cell can dictate the metabolic fate of amino acids. For example, in the gut, the
extent to which epithelial cells derive their amino acids from the luminal or vascular input is
affected by their stage of differentiation and physical location along the crypt–villus axis.
Studies showed that crypt cells are more highly labeled with isotopic tracers derived from the
blood, whereas villus cells are more highly labeled with tracers given luminally (Alpers, 1972).

Fig. 1. Rates of amino acid input into the PDV tissues from the diet and arterial circulation in young piglets.
Dietary inputs were based on intake of sow’s milk replacer fed at 12 g protein/kg/day. Arterial inputs were
calculated from measurements of arterial amino acid concentration and portal blood flow rate; this assumed
total arterial and portal blood flow to be equal. Adapted from Stoll et al. (1998).
160 D. G. Burrin and B. Stoll

These results implied that crypt cells derive their nutrients predominantly from the arterial
circulation, whereas villus cells rely on nutrients absorbed luminally from the diet. Likewise,
in the liver, studies have also shown a cellular zonation in the acinus, in which periportal
hepatocytes scavenge excess portal ammonia by ureagenesis, whereas perivenous hepatocytes
sequester ammonia via glutamine synthetase (Haussinger, 1990).
Once taken up by the splanchnic tissues, amino acids have three possible metabolic fates:
(1) incorporation into protein; (2) conversion via transamination or deamination into other
amino acids, metabolic substrates, and biosynthetic intermediates; and (3) complete oxidation
to CO2. In the case of intestinal epithelial cells, a fourth metabolic fate is transport into
the portal blood stream. In the first two pathways, amino acids can be deposited and recycled
by the body for purposes of growth or other biological functions. This recycling process may
be affected by whether an amino acid is incorporated into a constitutive protein or secreted
protein. However, from a nutritional perspective, if essential amino acids are irreversibly
metabolized or completely oxidized to CO2, this represents a nutritional loss to the animal.
If we first consider the metabolic fate of dietary amino acids in the gut, there are some gen-
eral observations that can be made from estimates of the net portal balance expressed as a
proportion of intake (table 1). The net portal balance represents the quantity of dietary amino
acid absorption into the portal blood expressed as a percent of the intake. Many of the values

Table 1
Summary of portal amino acid balance estimates in young pigs fed liquid milk-replacer under
different feeding conditionsa

Gastric hourly Continuous Single oral Continuous


bolusb (6 h) duodenalc (6 h) bolusd (8 h) duodenale (24 h)

Lysine 54 51 49 64
Threonine 38 67 62 33
Leucine 60 55 74 66
Isoleucine 70 84 78 30
Valine 61 74 72 51
Methionine 48 82 70 75
Phenylalanine 61 81 61 49
Histidine – – 70 –
Arginine 138 108 149 155
Proline 62 43 88 43
Tyrosine 167 78 96 168
Cysteine – – 17 –
Alanine 205 110 190 112
Serine 58 69 84 51
Glycine 52 65 69 89
Glutamate 7 5 29 10
Glutamine –8 –10 –29 –18
Aspartate 4 7 24 6
a Values represent net portal balance expressed as percentage of dietary intake. Pigs ranged from 6 to 10 kg body

weight and diets were fed to supply the estimated daily NRC protein requirement.
b Pigs fed via intragastric hourly boluses for 6 h and average portal balance measured between 4 and 6 h

(Stoll et al., 1998).


c Pigs fed continuously via intraduodenal infusion for 6 h and average portal balance measured between

4 and 6 h (Burrin et al., 2003a).


d Pigs fed single bolus meal orally and cumulative portal balance measured for 8 h (Bos et al., 2003).
e Pigs fed continuously via intraduodenal infusion for 12 h and fasted for 12 h; cumulative portal balance

measured for total 24 h period (van der Schoor et al., 2002).


Splanchnic protein and amino acid metabolism 161

are less than 100%, which implies that a portion of the amino acid is utilized by the gut; this
is because the diet fed in these studies was based on milk protein and essentially 95–100%
digestible. It is also important to note that the mode of enteral feeding in these four studies
varied considerably, between bolus versus continuous and gastric versus duodenal adminis-
tration of the exact same diet. The first observation is that the net balances of glutamate,
glutamine, and aspartate are nearly zero. In other words, the net utilization of these amino
acids by the PDV is approximately equal to the dietary intake. In some cases, the net balance
of glutamine is negative even in the fed state due to the high rate of metabolism in gut tissues.
As will be discussed below, the high fractional metabolism of these amino acids is due to their
integral role as oxidative fuels. A second remarkable observation is that the net balance of
arginine, alanine, and in some cases tyrosine and proline is greater than 100%, which sug-
gests a net production of these amino acids by the gut. The last observation is that net balance
of many essential amino acids is significantly less than 100%, and in some cases less than
50% of the dietary intake. These results indicate that gut metabolism of the dietary amino
acids significantly alters both the amount and pattern of amino acids absorbed into the portal
circulation.
In contrast to the PDV, there are few reports in the literature describing the net amino acid
balance by the liver in nonruminants, particularly in the fed state (Elwyn et al., 1968; Barrett
et al., 1986; Rerat et al., 1992; de Blaauw et al., 1996). There are numerous reports of hepatic
amino acid balance in ruminants (Wolff et al., 1972; Lobley et al., 1996a; Lapierre et al.,
1999; Blouin et al., 2002). Studies in pigs and dogs fed enterally demonstrated that the net
hepatic uptake of glycine and alanine is substantially greater (150–250%) than the dietary
intake. The only amino acids that were significantly released were glutamate and aspartate;
the net balance of glutamine was essentially zero. Among the remaining amino acids, hepatic
uptake was approximately 50–60% of the dietary intake. Interestingly, the net uptake of the
branched-chain amino acids was lower (35–43% of intake) than the other essentials; this
latter observation translates into a greater net splanchnic output of BCAA compared to other
essential amino acids. In addition to the animals where gut and liver amino acid metabolism
have been measured separately, there also have been numerous reports of total splanchnic
amino acid uptake in adult humans using stable isotopes (Castillo et al., 1993a; Matthews
et al., 1993; Battezzati et al., 1995, 1999; Haisch et al., 2000). These studies demonstrated the
substantial first-pass splanchnic extraction (percent enteral input) of amino acids, including
glutamate (96%), glutamine (64%), alanine (69%), arginine (38%), leucine (21%), and
phenylalanine (29%). A recent report in humans describing the kinetics of ingested 15N-labeled
soy protein showed that the splanchnic bed extracted nearly 60% of the dietary N, of which
40% was channeled into protein anabolism and 20% was deaminated (Fouillet et al., 2003).
Similar studies examining leucine kinetics found that first-pass splanchnic uptake in preterm
infants and elderly men was approximately 2-fold higher than in young adult men (Beaufrere
et al., 1992; Boirie et al., 1997).

2.2. Protein synthesis

A major metabolic fate of amino acids taken up by the gut and liver is incorporation into cel-
lular protein. Numerous studies have measured the rates of protein synthesis in various tissues
of the gastrointestinal tract and the liver. Among the literature reports, various methods have
been used to measure tissue protein synthesis in vivo; however, the best established and vali-
dated approach has been the flooding dose method, first reported by Garlick and others (see
discussion in Chapter 18; Garlick et al., 1980, 1994). Review of these studies reveals some
162 D. G. Burrin and B. Stoll

Table 2
Summary of tissue fractional protein synthesis rates in various mammalian speciesa

Weaned ratb Weaned mousec Neonatal pigletd Preruminant lambe

Stomach 140 51 55 –
Reticulo-rumen – 67 – 30
Abomasum – 55 – 56
Duodenum – 84 – 86
Jejunum 150 78 124 93
Ileum – 79 60 84
Cecum – – – 45
Colon 58 43 – 38
Pancreas 440 – 143 –
Liver 125 106 85 115
Spleen 38 70 55 –
Kidney 44 38 45 –
Brain 12 12 20 –
Skeletal muscle 7 3 30 21
a Values expressed as %/day.
b Goldspink and Kelly (1984); Goldspink et al. (1984); Preedy et al. (1988); Burrin et al. (1991, 1992).
c Bark et al. (1998); Burrin et al. (1999).
d Burrin et al. (1995, 1997); Stoll et al. (2000a).
e Attaix and Arnal (1987); Attaix et al. (1992).

general characteristics of protein synthesis (table 2), the first of which is relative rates of pro-
tein synthesis within the splanchnic tissue bed. In growing animals, the fractional protein
synthesis rates (FSR) are generally highest in the pancreas and progressively decline in small
intestine, stomach, and large intestine. In a weanling, 28-day-old rat, FSRs (%/day) in the
pancreas, small intestine, and stomach are 440%, 150%, and 140%, respectively (Burrin et
al., 1991). In the preruminant, one-week-old lamb, the FSRs also vary considerably among
the regions of the gut, including the rumen (30%), abomasum (56%), small intestine (88%),
cecum (45%), and colon (38%) (Attaix and Arnal, 1987; Attaix et al., 1992). The pattern of
declining proximal to distal gradient in intestinal FSR is also observed in neonatal piglets
(Stoll et al., 2000a) and mice (Bark et al., 1998). In comparison to the gastrointestinal tissues,
the FSR in the liver is relatively high, being similar to that of the proximal small intestine, but
lower than the pancreas.
In comparison to the splanchnic tissues, mainly small intestine and liver, the protein syn-
thesis rates in the other visceral organs and peripheral tissues are substantially lower. It is
notable that the skeletal muscle generally has the lowest protein synthesis rate among all
tissues, yet comprises the largest proportion of whole-body protein mass. If we consider the issue
of age or stage of development, the fractional rates of whole-body protein metabolism are
highest during fetal life, and decline progressively with advancing postconceptual age, com-
mensurate with fractional growth rates (fig. 2). (Goldspink and Kelly, 1984; Goldspink et al.,
1984.) This is also evident in the splanchnic tissues, in which the fractional protein synthesis
rates in the small and large intestine (107% to 61%) and in the liver (134% to 48%) decline
during the lifespan in rats. Interestingly, in the intestine, after weaning, the decline in intes-
tinal FSR with age is largely due to a decreased synthesis in the muscularis and serosal layers,
whereas the mucosal FSR remains constant (Merry et al., 1992). However, in striking contrast
to the fractional protein synthesis rate, the protein mass increases approximately 750-fold
Splanchnic protein and amino acid metabolism 163

Fig. 2. Ontogeny of whole-body and splanchnic tissue protein synthesis rates in rats. Adapted from
Goldspink and Kelly (1984) and Goldspink et al. (1984).

(2.5 to 1893 mg) in the intestine and 600-fold (17 to 3252 mg) in the liver between 18 days
gestational age and 105 weeks postpartum age (Goldspink and Kelly, 1984; Goldspink et al.,
1984). Despite the overall age-related decline, studies in neonatal rats and mice indicate that
the FSRs of the stomach, small intestine, and pancreas increase significantly after weaning
(Burrin et al., 1991, 1999a). In domestic animals, there are few, if any, estimates of gastroin-
testinal FSRs before birth and beyond pubertal ages, yet the changes between birth and
weaning in pigs and sheep tend to parallel those in rodents (Seve et al., 1986; Attaix et al.,
1992; Davis et al., 1996). In milk-fed animals, the intestinal FSR declines during the neona-
tal period, but increases markedly (40–50%) after weaning. The sharp increase in gut protein
synthesis after weaning in pigs is likely due to the substantial change in the composition of
the diet, gut microflora, and resultant stimulation of mucosal cellularity and proliferation
(Attaix and Meslin, 1991; Jiang et al., 2000).

2.3. Protein degradation

The kinetics of protein degradation are technically difficult to quantify, especially in vivo, and
thus there is a limited understanding of how certain metabolic factors affect this in the liver
and particularly in gut tissues (see discussion in Chapter 4). With regard to the gut, some
reports based on indirect estimates suggest that the fractional rates of protein degradation are
quantitatively similar to rates of protein synthesis, because the net balance between these two
opposing phenomena, i.e. the fractional protein accretion rate, is relatively low, at least in gut
tissues (Burrin et al., 1999b; Stoll et al., 2000a). There is a considerably larger literature on
the factors that influence qualitative aspects of protein degradation, based on studies with
perfused livers and cultured hepatocytes and colon carcinoma cells (HT-29) (Mortimore et al.,
1989; Kadowaki and Kanazawa, 2003; Ogier-Denis and Codogno, 2003).
The major class of protein degradation in the liver is lysosomal autophagy via the cathepsin
proteases, but also includes the ubiquitin–proteosomal system. In the intestine, three proteolytic
systems have been identified – lysosomal–autophagic cathepsins, ubiquitin–proteosomal, and
calpains – yet the relative significance of these systems to overall protein degradation is
unknown (Baracos et al., 2000). Hepatic autophagy has been shown to be potently inhibited
164 D. G. Burrin and B. Stoll

by several anabolic factors, including amino acids and insulin, and activated by glucagon
(Mortimore and Poso, 1987; van Sluijters et al., 2000). The regulatory actions of amino acids
on hepatic autophagy are specific for certain amino acids, particularly leucine. Moreover, the
process of autophagy has been linked to cell volume regulation and apoptotic cell death
(Schliess and Haussinger, 2002). These extracellular and intracellular signaling pathways will
be discussed in more detail in subsequent sections. Studies with perfused rat liver and isolated
hepatocytes have shown that autophagy increases to a peak at 6 months of age and then
declines with age by a process that can be prevented with calorie restriction (Del Roso et al.,
2003). As mentioned, the quantitative significance of protein degradation in intestinal
mucosal growth is essentially unknown. However, recent reports have shown that degradation
is upregulated by nutrient deprivation and exercise (de Blaauw et al., 1996; Samuels et al.,
1996; Halseth et al., 1997), but suppressed by glucagon-like peptide 2 (Burrin et al., 2000a).
In the study of de Blaauw et al. (1996), it is of interest to note that the rate of proteolysis
derived from 3H-phenylalanine kinetics was markedly higher (5-fold) in the PDV than in the
liver of the fasted rat.

2.4. Endogenous protein secretion and amino acid recycling

Another aspect of protein metabolism in the gut involves the metabolic fate of endogenous
protein secreted into the gut lumen. Endogenous proteins include secretions arising from the
saliva, gastric mucosa, bile, pancreas, and the exfoliation of epithelial cells. The biochemical
characteristics of endogenous proteins include enzymes (amylases, pepsinogen, trypsinogen),
glycocalix constituents (mucins), growth factors (epidermal growth factor, insulin-like growth
factor), and antioxidants (glutathione). Another significant source of protein in terminal ileum
is of microbial origin. Some reports indicate that as much as 25–50% of the nitrogen appearing
in ileal output is of microbial origin (Stein and Nyachoti, 2003). Given the long list of
endogenous proteins, it is perhaps not surprising that, collectively, these proteins represent a
quantitatively significant amount of amino acid released into the gut lumen. Estimated ileal
endogenous protein losses range from approximately 10–40 g protein/kg dry matter intake
and represent up to 10–25% of the dietary protein intake and 5–10% of the whole-body protein
turnover (Fuller and Reeds, 1998; Reeds et al., 1999). Recent studies have also shown that
endogenous ileal protein loss increases with protein intake (Hodgkinson et al., 2000). A critical
nutritional and metabolic question with respect to endogenous proteins secreted into the gut
lumen centers around the extent to which they are recycled and absorbed by the host. It is
important to note that measurements of ileal protein losses represent a minimal estimate of
total upper gut endogenous protein secretion, because evidence suggests that most (70–80%)
of the proteins secreted are hydrolyzed and reabsorbed within the small intestine (Stein and
Nyachoti, 2003). A recent study in young pigs demonstrated that, over a 24 h period, 52% of
the dietary amino acid intake was absorbed into the portal circulation and one-third of this
was derived from recycled intestinal secretions (van der Schoor et al., 2002).
In is generally held that the amino acids that pass from the terminal ileum into the cecum
and large intestine are catabolized by microbial fermentation. The assumption that the
endogenous amino acids are fermented and lost in the large intestine is based on early reports
that colonic absorption of amino acid is limited and occurs only during early postnatal
development (Fuller and Reeds, 1998). The carbon from colonic microbial amino acid catab-
olism can be lost as CO2 or reabsorbed into the portal blood in the form of short-chain fatty
acids. The nitrogen released by microbial amino acid catabolism may be in the form of
ammonia, which can be absorbed and recycled into the body amino acid and urea pools.
Splanchnic protein and amino acid metabolism 165

Fig. 3. Model illustrating gut microbial amino acid synthesis. Model depicts microbial lysine synthesis
derived from luminal ammonia nitrogen and blood urea nitrogen.

Historically, the colonic microbial catabolism of essential amino acids has been considered a
nutritional loss, because by definition, if an amino acid is essential, then this usually means
that it cannot be synthesized by mammalian cells. This concept has recently been challenged by
a number of elegant studies in pigs and humans, which demonstrated the microbial synthesis
of some essential amino acids, particularly lysine, based on labeling with 15N-labeled ammo-
nia and urea (fig. 3) (Metges, 2000; Torrallardona et al., 2003a,b). These studies have revealed
two critical phenomena: (1) that microbial lysine synthesis occurs in the upper gastrointesti-
nal tract and thus can be absorbed in the small intestine; and (2) the synthesis of several
essential amino acids may be nutritionally significant. These studies raise several intriguing
questions, including, (1) What are the carbon and nitrogen precursors for microbial amino acid
synthesis?, (2) How does the gut microbial load affect amino acid synthesis?, and (3) How is
this phenomenon regulated by dietary nutrient intake?

2.5. Amino acids as oxidative fuels

In recent years, it has become increasingly apparent that the splanchnic bed derives a majority
of its oxidative energy from the catabolism of amino acids, rather than glucose or fatty acids.
The liver is clearly the major site of amino acid metabolism in mammals and has been his-
torically considered a major site of catabolism and oxidation. However, since the classic
studies of Windmueller and Spaeth (1974, 1975, 1976, 1978, 1980), it has become readily
apparent that the gut, particularly the intestine, is also a major site of catabolism of several
amino acids. An important distinction to be made, however, is that while amino acids are
catabolized in both the liver and gut tissues, the extent to which they are completely oxidized
to CO2 may vary. In a brilliant review of amino acid oxidation, Jungas et al. (1992) system-
atically characterized the metabolic fate of dietary amino acids in the gut, liver, kidney, and
muscle and derived two important conclusions. First, they estimated that a primary metabolic
fate of amino acid carbon in the liver is conversion to glucose. They make the argument that
166 D. G. Burrin and B. Stoll

sufficient ATP is generated from the partial oxidation of dietary amino acids to generate
roughly half of the liver’s energy needs, an amount that approximates the energy required for
the synthesis of glucose. Thus, while amino acids are consumed in oxidative metabolic path-
ways in the liver, the complete oxidation of amino acids would far exceed the liver’s energy
needs and capacity to handle the end-products. A second important observation from their
analysis is that the hepatic oxidation of amino acids to glucose makes nearly two-thirds of the
total energy content of dietary amino acids available to peripheral tissues (as glucose). As
a result, there is no need for peripheral tissues to synthesize a complex array of enzymes to
oxidize amino acids and synthesize urea.
As discussed below, the gut is a willing participant in this process by releasing some of the
carbon derived from nonessential amino acid catabolism into the portal vein as alanine and
lactate, both of which are key precursors for hepatic gluconeogenesis. Glutamate is a key
amino acid linking hepatic amino acid catabolism and gluconeogenesis (Brosnan, 2000,
2003). Thus, it is not surprising that the gut releases proline, arginine, and ornithine into the
portal vein, all amino acids whose metabolism converges at glutamate in the liver. Therefore,
the splanchnic tissues, by design, are situated anatomically and metabolically within the
mammalian organism to regulate the flow of dietary amino acids, in such manner as to meet
their oxidative energy needs first, while at the same time ensuring the delivery of the primal
oxidative fuel for peripheral tissues, namely glucose.
The seminal studies of Windmueller and Spaeth (1974, 1975, 1976, 1978, 1980) were the
first to show evidence of extensive metabolism of glutamine, glutamate, and aspartate in
in situ intestinal perfusions in fasted, anesthetized rats. Results from young piglets fed a high-
protein, milk-based formula indicated that more than 95% of the dietary glutamine,
glutamate, and aspartate is utilized by the gut (Stoll et al., 1998). The studies of Windmueller
and Spaeth focused attention on the role of glutamine as the major oxidative fuel in the gut.
However, it is important to note that both glutamate and aspartate are of perhaps equal impor-
tance as intestinal oxidative fuels. Recent studies in young pigs and humans confirm the
extensive intestinal oxidation of dietary 13C-labeled glutamate and glutamine (Battezati et al.,
1995; Stoll et al., 1999a; Haisch et al., 2000).
The metabolism of glutamine is accomplished first by the catalysis via phosphate-dependent
glutaminase and subsequently by glutamate dehydrogenase (GDH) enzymes, both of which
are present in the stomach, small intestine, and colon of the young pig (Madej et al., 1999).
Interestingly, the activity of GDH is increased approximately 3-fold in the small intestine
after weaning. The resulting ketoacid product of GDH is α-ketoglutarate, which is then
metabolized yielding CO2 via the tricarboxylic acid cycle. It is important to note that,
although there is extensive uptake and metabolism of these three amino acids, their carbon
skeletons are not completely oxidized to CO2 and they do not account for all of the CO2
released by the gut. The in situ studies with perfused rat intestine and those in vivo with
piglets and humans indicate that most of the glutamine (55–70%), glutamate (52–64%), and
aspartate (52%) are oxidized to CO2 (Windmueller and Spaeth, 1976, 1978; Stoll et al.,
1999a). The remaining carbon atoms from these three substrates, which are not oxidized to
CO2, are converted to lactate, alanine, proline, citrulline, ornithine, and arginine and then
released into the portal circulation (Windmueller and Spaeth, 1975; Stoll et al., 1999a). The
metabolic fate of nitrogen from these amino acids is not fully understood. However, evidence
suggests that a portion of the nitrogen derived from glutamine and glutamate metabolism is
transferred to ammonia and other amino acids, including citrulline, ornithine, proline, and
arginine; much of the nitrogen from these products is converted to urea in the liver.
Splanchnic protein and amino acid metabolism 167

The rate of glutamine oxidation in isolated enterocytes decreases by roughly 90% in the
first 3 weeks of life (Darcy-Vrillon et al., 1994; Wu et al., 1995). Also, after weaning, in intra-
epithelial lymphocytes, glutamine is mainly metabolized to glutamate and ammonia (92%),
with minimal oxidation (4%). The decline in glutamine oxidation with age is paralleled by
increased activities of glutamine synthetase, glutaminase, and glutamate dehydrogenase
(Hahn et al., 1988; Shenoy et al., 1996; Madej et al., 1999). Studies with isolated enterocytes
also have shown that, along with glutamine, glucose is an important intestinal oxidative fuel
(Darcy-Vrillon et al., 1994; Kight and Fleming, 1995; Wu et al., 1995). However, glutamine
effectively suppresses glucose oxidation in enterocytes, whereas glucose has little impact on
glutamine oxidation (Kight and Fleming, 1995; Wu et al., 1995). The oxidation of glutamine
and its suppression of glucose oxidation was also found to be nearly twice as high (60%) in
the proximal compared to the distal (31%) small intestine (Kight and Fleming, 1995). The
relationship is consistent with in vivo studies in piglets demonstrating that, although glucose
represents an important oxidative fuel (29%), the proportion of glucose oxidized completely
to CO2 is substantially less than that of either glutamate or glutamine. The implication is that
glutamine and glutamate are preferentially channeled toward mitochondrial oxidation, while
most of the glucose is utilized for other metabolic or biosynthetic purposes.
Recent studies based on isotopic PDV tracer kinetics have shown that dietary essential
amino acids are also oxidized within the gut. Studies in young pigs showed that intestinal
oxidation of dietary lysine accounted for about one-third of whole-body lysine oxidation
and was completely suppressed by feeding a low-protein diet (van Goudoever et al., 2000).
Interestingly, although arterial lysine was taken up by the PDV, none of this was oxidized,
suggesting a preferential oxidation of dietary lysine (van Goudoever et al., 2000). As with
lysine, there is significant leucine metabolism by the gut via both transamination to ketoiso-
caproic acid (KIC) and complete oxidation to CO2. Studies in young pigs and dogs have
demonstrated that approximately 5–10% of whole-body leucine flux is oxidized by the PDV
(Yu et al., 1995; van der Schoor et al., 2001). Although approximately 40% of the leucine
taken up by the gut was converted to KIC, nearly all of this is transaminated back to leucine;
thus, the net KIC release is negligible (Yu et al., 1995). Studies in young, grower pigs (15–20 kg)
suggest that approximately 40% of the whole-body phenylalanine oxidation occurred in the
PDV tissues (Bush et al., 2003a). The oxidation of phenylalanine implies that phenylalanine
hydroxylation occurs in the gut and is consistent with previous observations suggesting net
portal tyrosine production in excess of dietary intake. Given that hydroxylation rather than
complete oxidation to CO2 represents the point of irreversible loss of phenylalanine, further
studies are warranted to quantify the proportion of whole-body phenylalanine flux metabolized
to tyrosine by the gut.
The reports of essential amino acid oxidation by the gut have raised the question of whether
this is due to mucosal metabolism or microbial fermentation. The recent evidence of de novo
lysine synthesis in the proximal intestine in pigs implicates a metabolically significant microbial
flora, which could also catabolize dietary amino acids. To address this issue, emerging studies
are beginning to identify the localization of essential amino acid catabolic enzymes within the
different mucosal cell phenotypes, i.e. enterocytes and lymphoid cells. Two reports have char-
acterized BCAA and lysine catabolic enzymes in enterocytes isolated from piglets at 0, 3 and
7 days of age (Elango et al., 2003; Pink et al., 2003). Branched-chain amino acid transferase
activity (BCAT) was detected in enterocytes and liver of 7-day-old piglets at a level that was
18% and 9%, respectively, of muscle mitochondrial BCAT activity. Enterocyte BCAT and
branched-chain dehydrogenase (BCKD) activity also increased between 0 and 7 days-of age.
168 D. G. Burrin and B. Stoll

BCAT and BCKD activity also was present in small intestinal mucosal tissue in piglets
(Burrin et al., 2003a). In the case of lysine, the first catabolic enzyme in the pathway, lysine
ketoglutarate reductase, was found in mitochondria from freshly isolated enterocytes at an
activity level about 50–60% as high as that measured in liver, whereas the activity of saccha-
ropine dehydrogenase was low. These findings are important for two reasons, the first of
which is that they demonstrate the amino acid catabolic capacity of the major mucosal cell
type relative to liver. Secondly, the results suggest that the lysine and leucine oxidation
reported in vivo in piglets is mediated partially by mucosal metabolism; however, the relative
significance of microbial catabolism remains to be determined.

2.6. Amino acids as biosynthetic precursors

Besides their incorporation into the bulk protein pool and use as oxidative fuels, amino acids
are metabolized by splanchnic tissues into a variety of end-products, which serve a variety of
key functions for the cell, specifically, and the host in general (table 3). An especially impor-
tant phenomenon is the interconversion of arginine, ornithine, and proline and their respective
roles in the biosynthesis of polyamines and nitric oxide in the intestinal mucosa. Studies have
shown that neonatal small intestine is an important site of arginine and proline synthesis and
this interconversion depends on first-pass metabolism (Murphy et al., 1996; Wu, 1998; Stoll
et al., 1999a; Bertolo et al., 2003). The milk of humans, pigs, rats, and many other mammals
is relatively deficient in arginine (Davis et al., 1994). Moreover, in the suckling pig, the intes-
tinal synthesis of arginine provides only about half of the animal’s needs for growth. Thus,
supplementation of dietary arginine is considered to be essential for maximal growth in young
piglets. The intestinal synthesis of arginine declines, while arginase activity increases sub-
stantially during the late suckling period (Wu and Morris, 1998). Also during this time, the
synthesis of citrulline and ornithine increases with age, particularly after weaning, and proline
and arginine are major precursors for their synthesis (Wu, 1997). Thus, in adult rats and weaning
pigs, the intestinal conversion of glutamine, glutamate, and proline to citrulline provides a
critical precursor for arginine synthesis in the kidney (Windmueller and Spaeth, 1980; Dugan
et al., 1995). The dependence on intestinal first-pass metabolism for synthesis of either arginine
(in neonates) or its immediate renal precursor, citrulline (in adults), results in arginine defi-
ciency when intestinal metabolism is either bypassed during TPN (Brunton et al., 1999) or
removed surgically by resection (Wakabayashi et al., 1995). The dependence on the intestine
for citrulline synthesis has led to development of plasma citrulline as a marker for enterocyte
mass in conditions of disease (Crenn et al., 2003).
Studies in cultured intestinal cells have shown that ornithine derived from arginine metab-
olism is converted to polyamines (Blachier et al., 1995). Polyamines (putrescine, spermidine,

Table 3
Amino acid precursors and functional end-products produced in splanchnic tissues

Precursors • Arginine • Glutamine • Methionine • Threonine


• Proline • Glutamate • Cysteine • Serine
• Aspartate
Products • Nitric oxide • Nucleotides • Cysteine • Mucins
• Polyamines • Glutathione • Taurine
• Creatinine • Glucosamine • Glutathione
• Thioredoxin
Splanchnic protein and amino acid metabolism 169

spermine, cadaverine) are ubiquitous cationic amines involved in cell proliferation and
differentiation in many tissues, including the gastrointestinal tract. Ornithine decarboxylase
(ODC) and S-adenosylmethionine decarboxylase (SAMDC), converting ornithine to putrescine
and putrescine to spermine, respectively, are the rate-limiting enzymes in polyamine synthesis.
The synthesis of polyamines from arginine is negligible in enterocytes of newborn and suck-
ling animals (Blachier et al., 1991, 1992), but polyamines are present in mammalian milk
(Pollack et al., 1992; Buts et al., 1995). As piglets progress from suckling to weaning, major
end-products of intestinal enterocyte arginine metabolism are proline and ornithine. Thus,
when the ingestion of milk-borne polyamines by the neonate ceases after weaning, the induction
of intestinal polyamine synthesis from ornithine, arginine, and proline becomes physiologi-
cally significant for the maintenance of normal intestinal growth and function (Wu et al.,
2000a,b). Furthermore, the induction of intestinal polyamine synthesis is dependent on the
weaning-induced cortisol surge.
Another important end-product of arginine metabolism is nitric oxide. Nitric oxide is a
major physiological regulator in the body, particularly by enhancing vascular function, tissue
perfusion, and immune function. For this reason, arginine is a key component supplemented
to enteral formulas designed for trauma and surgically stressed patients (Huang et al., 2003;
McCowen and Bistrian, 2003) and may become limiting under conditions of increased NO
production (Hallemeesch et al., 2002). Nitric oxide is produced by conversion of arginine to
citrulline by the enzyme nitric oxide synthase (NOS), which is expressed in three isoforms,
all of which are found in gastrointestinal and liver tissue. In the whole body, the proportion
of arginine that is converted to NO is relatively low (1–10% arginine flux) (Castillo et al.,
1996), but is increased in response to stress and trauma (Argaman et al., 2003). However, the
first-pass splanchnic utilization of enteral arginine is about 40%, of which metabolism to NO
represents 16% of the whole-body nitrate production (Castillo et al., 1993a,b). The rates of
liver and PDV NO production have been measured using a similar isotopic approach based
on kinetic conversion of 15N-arginine to 15N-citrulline (Luiking and Deutz, 2003). These studies
showed that as much as 35% of the arginine utilization is converted to NO in the liver and gut
in endotoxemic pigs; NO synthesis also is increased markedly with supplemental arginine
(Bruins et al., 2002a,b).
The nonessential amino acids, aspartate, glutamine, and glycine, are key precursors for the
synthesis of nucleotides. This fact is quantitatively important in the intestinal mucosa given
the high rate of cell proliferation coupled with the fact that most of the nucleotides, at least
ribonucleotides, are synthesized de novo (Boza et al., 1996). The amide nitrogen from gluta-
mine serves as the nitrogen donor for the synthesis of both purines and pyrimidines, whereas
most of the carbon skeleton of nucleotides is derived from aspartate and glycine. The
biochemical mechanism whereby glutamine affects intestinal function also may be related to
its conversion to glucosamine, which reduces the cellular NADPH and suppresses nitric oxide
synthesis (Wu et al., 2001). In immune cells, glutamine and glutamate also provide an impor-
tant source of NADPH via conversion of malate to pyruvate; NADPH is critical in these cells
for production of superoxide and NO and for glutathione reductase activity (Newsholme
et al., 2003).
Amino acids also serve as precursors for synthesis of compounds involved in support of
innate immunity in the gut and antioxidant function in both liver and gut. Dietary threonine
and cysteine are considered to be important for mucin synthesis by goblet cells within the
stomach and intestinal mucosa. The secretory mucins play a key role in the innate immune
defense of the mucosa, and the core protein of the major intestinal mucins contains a large
amount of threonine and cysteine (van Klinken et al., 1997; Faure et al., 2002). Studies in pigs
170 D. G. Burrin and B. Stoll

indicated that as much as 60% of dietary threonine is utilized by the gut in first-pass (Stoll
et al., 1998). Consistent with these findings, other studies in piglets (Bertolo et al., 1998)
demonstrated that the threonine requirement of piglets maintained by parenteral nutrition was
nearly 60% lower than that of piglets receiving enteral feedings. A subsequent report found
that feeding threonine-deficient diets to piglets significantly reduces intestinal mass and
goblet cell numbers, and this suppression of intestinal growth cannot be fully restored by pro-
viding threonine parenterally (Ball et al., 1999). The mucosal synthesis and secretion of
mucins are likely to be quantitatively significant, and thus the needs for dietary threonine and
cysteine may be increased under conditions of gut hypersensitivity or inflammation. A recent
report in rats, using a novel approach to purify mucins, demonstrated that the synthesis rate
of mucin glycoproteins was relatively constant along the length of the intestine (range
112–138%/day), but substantially higher than the total mucosal protein synthesis rate, espe-
cially in the distal bowel: values for total protein were 77% and 44%/day in the ileum and
colon, respectively (Faure et al., 2002).
Besides incorporation into mucin and other mucosal proteins, the extent to which threonine
is further metabolized within the gut tissues is poorly understood. There is some debate as to
which metabolic pathway is most important in the catabolism of threonine in mammals
(House et al., 2001). The two predominant pathways of threonine catabolism in the pig are
believed to be catalyzed by either threonine aldolase/dehydrogenase or threonine dehydratase
(Ballevre et al., 1990; Le Floc’h et al., 1996, 1997). Studies in pigs suggest that conversion
to glycine via threonine aldolase/dehydrogenase is the predominant pathway of irreversible
threonine catabolism. Moreover, threonine dehydrogenase activity was localized in both
the liver and pancreas, but not other gut tissues, implicating the PDV as a possible site of
threonine catabolism.
In addition to mucins, methionine and cysteine serve as biosynthetic precursors for numer-
ous functional end-products, including glutathione, polyamines, and taurine. In numerous
cells within the body, methionine is metabolized via transmethylation to homocysteine and in
the process produces S-adenosylmethionine, which donates an aminopropyl moiety in the for-
mation of the polyamines, spermidine and spermine (fig. 4) (Finkelstein, 2000). Homocysteine
is converted to cysteine via transsulfuration. Cysteine is one of three constituent amino acids
of glutathione (GSH), along with glutamate and glycine. Moreover, cysteine is metabolized
to form taurine. Glutathione is a major cellular antioxidant in cells found in the intestinal
mucosa and liver (Lash et al., 1986; Martensson et al., 1991). However, cysteine and taurine
can also function as cellular antioxidants (Santangelo, 2002; Zafarullah et al., 2003). The
quantitative significance of these functional end-products to splanchnic methionine, cysteine,
and glutamate utilization is unknown. Early studies in humans suggested that splanchnic tis-
sues are an important site of transsulfuration (Stegink and Den Besten, 1972). Studies in
piglets indicate that first-pass utilization of dietary methionine ranged from 30% to 40%
(Rerat et al., 1992; Stoll et al., 1998; Bos et al., 2003). A recent study in neonatal piglets
showed that the methionine requirement (g/kg/day) was 0.42 and 0.29 in enteral and parenter-
ally fed piglets, respectively, which implies that first-pass splanchnic metabolism accounts for
30% of the dietary methionine requirement (Shoveller et al., 2003).
The original studies by Finkelstein (2000) demonstrated that gastrointestinal tissues
possess the enzymes necessary to metabolize methionine to cysteine, albeit at significantly
lower activities than the liver. However, there are few reports describing the kinetics of
methionine metabolism in the gut, either in vivo or in vitro with isolated enterocytes. Recent
studies based on enzyme assay and in vivo isotopic tracers in ruminants imply that methionine
transmethylation occurs in the ruminant gut and that the activities are comparable to the liver
Splanchnic protein and amino acid metabolism 171

Fig. 4. Schematic overview of sulfur amino acid metabolism. Abbreviations: SAM, S-adenosylmethionine;
SAH, S-adenosylhomocysteine.

(Lobley et al., 1996b, 2003; Lambert et al., 2002). In the case of dietary cysteine, studies in
pigs indicate that rate of appearance into the portal blood is very limited (less than 20%
dietary, intake), suggesting extensive intestinal utilization of cysteine in first-pass (Rerat et al.,
1992; Stoll et al., 1998; Bos et al., 2003). The first step in cysteine catabolism is conversion
to cysteinesulfinate via the enzyme cysteine dioxygenase. Cysteinesulfinate is then converted
to taurine via cysteinesulfinate decarboxylase or pyruvate via aspartate aminotransferase.
Rodent studies with 1-14C-labeled cysteine demonstrated significantly higher oxidation when
given via the intragastric (70%) than intraperitoneal (41%) route, suggesting that nearly half
of the whole-body cysteine oxidation occurs in splanchnic tissues (Stipanuk and Rotter,
1984). The increased oxidation of intragastric versus systemic cysteine was largely attributed
to increased oxidation to pyruvate rather than to taurine. Subsequent work demonstrated that
intestinal enterocytes extensively metabolize cysteine via cysteine dioxygenase to cysteine-
sulfanate (Coloso and Stipanuk, 1989). In vivo rodent studies with intravenous infusion of
isotopically labeled 15N-cysteine indicate that an important metabolic fate of cysteine in the
gut is incorporation into glutathione (GSH) (Malmezat et al., 2000a). With respect to gluta-
mate, studies in piglets demonstrated that enteral glutamate is preferentially incorporated into
mucosal GSH (Reeds et al., 1997).
172 D. G. Burrin and B. Stoll

3. FACTORS REGULATING PROTEIN AND AMINO ACID METABOLISM


3.1. Nutrition

Oral feeding is probably the most potent stimulus of splanchnic protein and amino acid
metabolism in growing animals (McNurlan et al., 1979; Burrin et al., 1991, 1992, 1995).
Prolonged fasting leads to markedly reduced protein mass in gut and liver tissues via sup-
pressed protein synthesis and increased protein degradation, especially in the small intestine
(Samuels et al., 1996). Recent studies have shown that the stimulatory effect of nutrient intake
on gut protein metabolism is dependent on enteral administration of nutrients (Dudley et al.,
1998; Burrin et al., 2000b, Stoll et al., 2000a). These studies showed that the fractional protein
synthesis rate and the net balance between intestinal protein loss and accretion is directly and
positively determined by the level of enteral intake. An enteral nutrient intake of 20% of total
was found to maintain intestinal protein balance. Similarly, a recent study in preterm infants
reported that minimal enteral feeding of approximately 10–15% of total increased splanchnic
leucine uptake (Saenz de Pipaon et al., 2003). In contrast to the gut, however, the rate of
hepatic protein synthesis decreased with the level of enteral nutrition in neonatal piglets,
which is consistent with an enlarged liver mass associated with TPN (fig. 5).
The composition of dietary nutrients can also substantially affect splanchnic protein and
amino acid metabolism. An important consideration in the neonate is the stimulation of liver
and gut function in response to the onset of suckling colostrum at birth (Burrin et al., 1995).
Studies in neonatal piglets showed that both liver and gut protein synthesis is rapidly upreg-
ulated with the first feeding of colostrum. The relatively high concentrations of colostral
growth factors are thought to provide key signals for intestinal and liver development.
However, studies comparing colostrum to macronutrient matched formula and those with
enteral supplementation of recombinant growth factors (IGF-I) suggest that milk-borne
growth factors have limited trophic effects on the intestine (Burrin et al., 1999a, 2001). The
impact of these milk-borne growth factors is probably more important for stimulation of the
gut immune system. Thus, it appears that the macronutrient intake is the major component of
the diet affecting splanchnic tissue protein metabolism in developing mammals.
In humans, first-pass splanchnic leucine uptake is nearly 2-fold lower in subjects fed a
protein-free diet compared to controls fed complete protein-containing diet (Cayol et al., 1997).

Fig. 5. Role of enteral and parenteral nutrition on intestinal and liver protein synthesis in neonatal piglets.
Adapted from Stoll et al. (2000a) and unpublished results.
Splanchnic protein and amino acid metabolism 173

Some studies have shown that restriction of dietary protein has limited effects on intestinal
protein synthesis (Seve et al., 1986), whereas others have reported a decrease in protein syn-
thesis in both liver and gut (McNurlan and Garlick, 1981; Wykes et al., 1996). Studies in
neonatal pigs demonstrated that protein malnutrition significantly reduces whole-body
growth and amino acid absorption, but does not affect gut tissue growth (Ebner et al., 1994;
van Goudoever et al., 2000). Consistent with this, studies in pigs fed low-protein diets showed
that PDV metabolized 75% of enteral 13C-lysine intake, compared to 45% in high-protein-fed
pigs (van Goudoever et al., 2000). Protein malnutrition also suppressed the gut oxidation of
lysine, leucine, and glutamate, whereas glucose oxidation increased (van Goudoever et al.,
2000; van der Schoor et al., 2001). These piglet studies suggest that when dietary protein
intake is reduced, the gut amino acid utilization for protein synthesis is maintained, but amino
acid oxidation is suppressed.
Studies focusing on the impact of dietary macronutrients found that enteral amino acids,
but not carbohydrate or lipid, stimulated intestinal protein synthesis, whereas each of these
stimulated gut protein accretion, suggesting that carbohydrate and lipid suppressed gut
proteolysis (Stoll et al., 2000b). In contrast, in other studies, enteral amino acids rapidly decreased
intestinal protein synthesis and proteolysis, whereas parenteral or luminal glucose infusion
increased intestinal protein synthesis (Weber et al., 1989; Adegoke et al., 1999, 2003). Amino
acids also potently inhibit hepatic autophagic and proteasomal proteolysis (Mortimore and
Poso, 1987; Hamel et al., 2003). Feeding a high-fat versus high-carbohydrate diet to young
pigs stimulated protein synthesis in the intestine, but not the liver (Ponter et al., 1994).
Numerous studies have examined the effect of dietary supplementation with amino acids on
intestinal and liver protein metabolism. In young pigs, amino acid deprivation suppresses
whereas supplementing amino acids parenterally stimulates hepatic protein and albumin syn-
thesis, but not intestinal protein synthesis (Davis et al., 2002a; Hellstern et al., 2002). Similarly,
in humans, parenteral amino acid infusion stimulated splanchnic tissue protein synthesis and
suppressed protein degradation (Nygren and Nair, 2003).
Among the specific amino acids, glutamine is most extensively studied. Supplementing
glutamine has been shown to stimulate protein synthesis and reduce proteolysis in the gut in
some cases (Coeffier et al., 2003), but not in others (Garcia-Arumi et al., 1995; Marchini et al.,
1999; Bouteloup-Demange et al., 2000). Glutamine also has been shown to stimulate protein
synthesis and suppress protein degradation in cultured enterocytes and hepatocytes
(Higashiguchi et al., 1993; Le Bacquer et al., 2001, 2003; Haussinger et al., 2001). Studies
have shown positive effects of glutamine-supplemented TPN in preventing atrophy, stimulat-
ing protein anabolism, and maintaining intestinal permeability (Tamada et al., 1992; Inoue
et al., 1993; Platell et al., 1993; Haque et al., 1996; Naka et al., 1997; Khan et al., 1999).
In contrast, other studies have shown no effect on gut growth or protein metabolism in ani-
mals receiving glutamine supplementation parenterally (Spaeth et al., 1993; Burrin et al.,
1994; Marchini et al., 1999; Humbert et al., 2001). Furthermore, parenteral infusion of lipid
has been shown to stimulate intestinal protein synthesis more than either glucose or glutamine
(Stein et al., 1994). Thus, the intestinal effects of glutamine-supplemented TPN are evident
under conditions of compromised gut function, such as sepsis, inflammation, and small-bowel
resection, while in healthy animals glutamine has limited effect.
Dietary leucine stimulates liver protein synthesis (Kimball and Jefferson, 2001; Lynch et al.,
2002) and suppresses proteolysis (Mortimore and Poso, 1987). Dietary tryptophan increased
(Ponter et al., 1994) whereas arginine supplementation reduced (Bruins et al., 2002b) hepatic
protein synthesis in pigs. Other dietary components that stimulate mucosal proliferation and
cell turnover also tend to increase protein synthesis. Feeding fiber and lectins has been shown
174 D. G. Burrin and B. Stoll

to stimulate intestinal protein synthesis in some cases (Southon et al., 1985; Palmer et al.,
1987), but not in others (Nyachoti et al., 2000). Studies show that neonatal pigs fed elemen-
tal diets have higher rates of protein synthesis than those fed a polymeric, cow’s milk formula
(Stoll et al., 2000c).

3.2. Anabolic and catabolic hormones

3.2.1. Insulin

Insulin is generally considered to be one of the most anabolic hormones in the body, affect-
ing multiple metabolic pathways, including those involving protein and amino acid
metabolism. However, with respect to protein and amino acid metabolism, the effects of
insulin on splanchnic tissues (i.e. liver and gut) are somewhat different from other major
tissues, such as muscle. Early studies in vivo and with perfused livers and isolated hepatocytes
from normal and streptozotocin-induced diabetic rats indicated that insulin increased protein
synthesis and suppressed proteolysis (McNurlan and Garlick, 1981; Mortimore and
Poso, 1987; Kimball and Jefferson, 1994). The effects of insulin on hepatic protein synthesis
are most pronounced for albumin. In the intestine, however, diabetes has no effect on protein
synthesis despite the induction of significant hyperphagia. More recently, several in vivo
studies in pigs, mice, and humans suggest that insulin has either no effect or suppresses
hepatic protein synthesis (Mosoni et al., 1993; Nair et al., 1995; Bark et al., 1998; Meek
et al., 1998; Ahlman et al., 2001, Boirie et al., 2001; Davis et al., 2001, 2002a; Nygren and
Nair, 2003). Similarly, studies in pigs reported no effect of insulin on intestinal protein syn-
thesis (Davis et al., 2001, 2002a), whereas insulin administration to adult, type I diabetic
subjects only modestly (~15%) increased intestinal protein synthesis (Charlton et al., 2000).
Thus, with respect to protein synthesis, the splanchnic tissues appear to be relatively insulin-
insensitive. This finding is consistent with the idea that protein synthesis in the splanchnic
tissues is more tightly regulated by nutrient availability, especially amino acids, than by
hormones.
The early observations that insulin suppresses proteolysis in perfused liver and hepatocytes
have been supported by recent in vivo studies in humans (Nair et al., 1995; Nygren and Nair,
2003). Moreover, the studies have established several cellular signaling pathways that appear
to mediate the insulin-induced suppression of hepatic proteolysis (Kadowaki and Kanazawa,
2003; Schliess and Haussinger, 2003). In isolated hepatocytes, insulin activates its membrane
receptor, which triggers activation of downstream factors including PI3-kinase, protein kinase B,
and p70S6-kinase (fig. 6) (Krause et al., 2002). The key amino acid-sensing and insulin-
inducible signaling component involved in cell protein metabolism is mTOR, yet it is unclear
whether mTOR mediates the insulin-induced suppression of proteolysis in hepatocytes. With
respect to protein metabolism, the explanation for the observed insulin-insensitivity of intes-
tinal tissue is largely unexplored, yet intestinal epithelial cells express insulin receptors and
express all of the signaling pathways known to be responsive to insulin. Cell volume or hydra-
tion state is another key cellular signaling process found to be involved in insulin-mediated
suppression of hepatic proteolysis. Studies in hepatocytes indicate that increased cell volume
is critical for many of the insulin-mediated effects. Insulin produces cell swelling by induc-
tion of sodium and potassium accumulation. The signaling mechanisms involved with
insulin-induced cell swelling are not fully established, although cell swelling has been shown
to activate three major mitogen-activated protein (MAP)-kinase pathways (ERK, JNK,
p38MAPK) (Haussinger and Schliess, 1999).
Splanchnic protein and amino acid metabolism 175

Fig. 6. Intracellular signaling pathways for selected hormones and nutrients that alter tissue protein metabo-
lism. Abbreviations: cAMP, cyclic AMP; PKA, protein kinase A; AMPK, adenosine monophosphate-activated
protein kinase; ATP, adenosine triphosphate; GLP-2, glucagon-like peptide 2: GCN2, general control non-
depressing kinase-2 amino acid-regulated eukaryotic initiation factor kinase; mTOR, mammalian target of
rapamycin; eIF2B, eukaryotic initiation factor (eIF) 2B; PKB, protein kinase B; IRS, insulin receptor substrate;
GSK3, glycogen synthase kinase-3; 4E-BP1, eukaryotic initiation factor 4E binding protein-1; p70S6K, 70 kDa
ribosomal protein S6 protein kinase; PI3K, phosphatidylinositol 3-kinase; IGF-I, insulin-like growth factor I.

3.2.2. Growth hormone–insulin-like growth factor I axis

Another major anabolic influence on protein metabolism in developing mammals is the soma-
totropic axis, which involves pituitary growth hormone (GH) secretion and local expression
and secretion of insulin-like growth factors (IGF), mainly IGF-I. The somatomedin theory
originally proposed was based on the idea that GH-induced secretion of IGF-I in the liver is
a key endocrine signal for somatic growth and metabolism during postnatal development
(Butler et al., 2002). However, recent use of tissue-specific gene deletion of hepatic IGF-I has
challenged this idea and suggestes that the endocrine role of circulating IGF-I is not essential
for normal postnatal growth. Both GH and IGF-I appear to have trophic effects in splanchnic
tissues that appear to be mediated by stimulation of protein and amino acid metabolism; how-
ever, there is considerably less information than for skeletal muscle. The receptors for both
GH and the type I IGF receptor are expressed in liver and intestine tissues. Systemic admin-
istration of either GH or IGF-I has been shown to stimulate liver and intestinal growth in
growing animals, yet the response is dependent on stage of development, being less respon-
sive during early postnatal life (Etherton and Bauman, 1998; Wester et al., 1998). The
diminished effect of GH in neonatal animals is due to lower expression of the GH receptor.
Early studies with rats in vivo and perfused livers indicated that hepatic protein synthesis,
particularly albumin, and amino acid transport were suppressed by hypophysectomy and
restored with growth hormone treatment (Jefferson et al., 1975; Feldhoff et al., 1977). These
findings have been confirmed by more recent studies in pigs, rats, and humans showing
GH-mediated increases in liver protein synthesis (Pell and Bates, 1992; Wester et al., 1998; Barle
et al., 1999, 2001; O’Leary et al., 2003; Bush et al., 2003b). In pigs, the stimulation of liver pro-
tein synthesis occurred in both fasted and fed animals and was associated with increased
ribosome number rather than translational mechanisms (Bush et al., 2003a). Evidence of GH
176 D. G. Burrin and B. Stoll

action in the gut is limited, with two studies reporting a stimulation of protein synthesis in
intestinal (Lo and Ney, 1996) and PDV tissues (Bush et al., 2003b).
An important consideration with respect to the mechanism of GH action is whether the
effects are mediated directly via the GH receptor or indirectly via increased local expression
of IGF-I. Interpretation of this issue has been complicated by studies in which IGF-I infusion
was found to stimulate hepatic (Douglas et al., 1991; Koea et al., 1992) and intestinal (Lo and
Ney, 1996; Tashiro et al., 1999) protein synthesis in some cases, but not in others (Pell and
Bates, 1992; Ling et al., 1995; Lo and Ney, 1996; Bark et al., 1998; Davis et al., 2002b).
However, enteral administration of IGF-I in neonatal piglets and mice has no effect on liver or
intestinal protein synthesis (Burrin et al., 1999a, 2001). The observation of consistent
responses to GH, but not IGF-I infusion alone, suggests a direct effect of GH on hepatic
protein synthesis. However, there is virtually no information describing the intracellular
signaling pathways, which link GH receptor signals (i.e. Janus family kinases [JAK] and signal
transducers and activators of transcription [STAT]) to the cellular protein synthesis machinery.
To the extent that the anabolic effects of GH are driven by IGF-I production, it is of interest
that studies in vivo and in cultured hepatocytes indicate that the hepatic response to GH is
reduced by limiting dietary protein intake or amino acid availability (Harp et al., 1991;
Brameld, 1997; Brameld et al., 1999). Recent studies in hepatocytes suggest that the avail-
ability of amino acids and glucose directly suppresses IGF-I and GH receptor expression,
respectively (Brameld et al., 1999; Stubbs et al., 2002). Moreover, some specific amino acids
(methionine, lysine, leucine, tryptophan) appear to be essential for the GH-mediated induc-
tion of IGF-I expression in hepatocytes. A further contributing factor to the positive
interaction between dietary protein and GH-induced IGF-I expression is insulin. Studies indi-
cate that insulin stimulates hepatic IGF-I expression, thus increased insulin secretion in
response to higher dietary protein may contribute to the stimulation of IGF-I expression
(Boni-Schnetzler et al., 1991; Brameld, 1997). Other hormones, namely thyroxine and gluco-
corticoids, have been shown to increase hepatic GH receptor and IGF-I expression when given
in combination with GH, in some cases (Brameld, 1997) but not others (Beauloye et al., 1999).
The suppression of hepatic amino acid catabolism and urea synthesis also contributes to the
protein anabolic effect of GH. A number of studies in rodents, pigs, and humans have shown
that GH reduces the in vivo synthesis of urea, activity of urea cycle enzymes, and catabolism
of essential amino acids (Dahms et al., 1989; Blemings et al., 1996; Grofte et al., 1997; Gahl
et al., 1998; Bush et al., 2002). GH has also been reported to increase the expression of
hepatic glutamine synthetase (Nolan et al., 1990). Whether these effects are mediated directly
via GH receptor signaling or indirectly via IGF-I expression is unknown and warrants further
study. In addition, it is unknown how the intracellular signaling pathways associated with GH
and IGF-I are linked to cytosolic and mitochondrial amino acid catabolic and urea cycle
enzymes. Thus, the general metabolic effect of GH is to reduce amino acid catabolism and
increase scavenging of ammonia in the liver, thereby channeling amino acids into protein
synthesis in both the liver and their release to peripheral tissues, such as muscle.

3.2.3. Glucocorticoids and glucagon

Glucocorticoids play a critical role in the induction of differentiation and development of


many organ systems in mammals. However, glucocorticoids appear to have tissue-specific
actions on protein metabolism, being catabolic in skeletal muscle and intestine and anabolic
in the liver. Pharmacological doses of glucocorticoids (e.g. dexamethasone) usually increase
liver protein mass, thus its classification as a catabolic hormone is not strictly correct.
Splanchnic protein and amino acid metabolism 177

Early studies in rodents in vivo and in cultured hepatocytes showed that corticosterone or
dexamethasone treatment induced a significant stimulation of constitutive and secretory hepatic
protein synthesis (Odedra et al., 1983; Hutson et al., 1987; Southorn et al., 1990).
Glucocorticoids also stimulate proteolysis in hepatocytes isolated from weanling but not fetal or
neonatal rats, which may in part explain the generally higher state of liver protein anabolism in
early development (Hopgood et al., 1981; Blommaart et al., 1993). However, under conditions
where plasma glucocorticoids are physiologically elevated, as in starvation and stress, hepatic
protein mass declines and gluconeogenic enzyme activity is increased. Consistent with this,
dexamethasone has been shown to increase the hepatic activity of amino acid catabolic enzymes
involved in transsulfuration and branched-chain amino acid metabolism (Huang and Chuang,
1999; Ratnam et al., 2002). These findings in the liver contrast with the suppression of protein
synthesis by glucocorticoids in intestine (Burrin et al., 1999c; Boza et al., 2001) and muscle
(Southorn et al., 1990). Dexamethasone results in a catabolic suppression of intestinal growth, as
well as a stimulation of the transport and metabolism of amino acids, especially glutamine
(Souba et al., 1985; Salleh et al., 1988; Iannoli et al., 1998). The dexamethasone induction of
glutamine metabolism is associated with an increase in glutaminase activity and gene expression.
Glucagon is generally considered as a catabolic hormone associated with starvation, but
also is significantly increased in the circulation during high-protein feeding. The available
literature describing the effects of glucagon on splanchnic tissue protein metabolism is con-
fined largely to the liver. Studies in vivo and in isolated hepatocytes showed that glucagon
stimulates hepatic amino acid catabolism, urea cycle enzyme activities, and urea synthesis
(Morris, 2002). Most of the effects of glucagon are considered to be mediated by induction
of cellular cAMP production. These studies have reported a 2-fold stimulation of methionine
uptake and a 5-fold stimulation of transsulfuration via cystathionine β-synthase in glucagon-
treated hepatocytes (Jacobs et al., 2001). Glucagon treatment also increased threonine uptake
and oxidation via activation of threonine dehydratase activity in hepatocytes (House et al.,
2001). Oxidation of arginine, but not ornithine, in hepatocytes is increased by glucagon treat-
ment (O’Sullivan et al., 2000). Studies with perfused livers have shown an increased
incorporation of glutamine nitrogen into urea in association with increased glutaminase activ-
ity after glucagon treatment (Brosnan et al., 1995; Nissim et al., 1999). With respect to protein
turnover, studies with hepatocytes and perfused livers indicate that glucagon stimulates both
albumin synthesis (Kimball et al., 1995) and autophagic proteolysis (Mortimore and Poso,
1987). There is little known about the effect of glucagon, per se, on the intestinal protein
metabolism. However, there are numerous studies that have examined the effects of glucagon-
like peptides (GLP-1 and GLP-2) on the intestine and pancreas (Drucker, 2002; Burrin et al.,
2003b); yet few of these have examined aspects of protein and amino acid metabolism. GLP-1
and GLP-2 have anabolic effects on the pancreatic beta cells and small intestinal mucosa,
respectively. Recent studies in TPN-fed neonatal piglets have demonstrated that chronic GLP-2
treatment prevents mucosal atrophy by suppressing apoptosis and proteolysis (Burrin et al.,
2000a), whereas acute GLP-2 infusion upregulates intestinal amino acid uptake and protein
synthesis (Guan et al., 2003).

3.3. Infection, inflammation, and commensal microflora

3.3.1. Pathogenic infection

In the past twenty years, it has become increasingly evident that the relationship between the
host and the ecology of resident microbes plays an integral role in the normal development,
metabolism, and survival of mammalian species (Hooper et al., 2002). The infestation with
178 D. G. Burrin and B. Stoll

pathogenic microbial and viral organisms, or exposure to the toxins they produce, has a potent
protein anabolic effect on splanchnic tissues, whereas protein catabolism is increased in
skeletal muscle in response to infection and inflammation (MacRae, 1993; Grimble, 2001).
Proinflammatory cytokines, (i.e. tumor necrosis factor, interleukin-1, and interleukin-6) are
the key signals that transmit the presence of pathogenic insult to the organism resulting in
activation of the immune system and acute-phase response and suppression of feeding behavior
(Johnson, 1997, 2002). Splanchnic tissues play a central role in the proinflammatory response
since the liver, spleen, and intestine are the major lymphoid tissues and primary sites of acute-
phase protein synthesis in the body. The acute-phase response functions to mobilize
endogenous amino acids from muscle, which are used for support of acute-phase protein syn-
thesis and cell-mediated immune response. Numerous studies have demonstrated that
treatment with proinflammatory stimuli, including live bacteria, enteric parasites, endotoxin,
and specific cytokines, significantly increases acute-phase protein synthesis in visceral tis-
sues, especially the liver, gut, and spleen (Jepson et al., 1986; Vary and Kimball, 1992;
von Allmen et al., 1992; Higashiguchi et al., 1994; Breuille et al., 1998; Wang et al., 1998;
Breuille et al., 1999; Mack et al., 1999). Conversely, these factors increase catabolism and net
loss of muscle protein mass. In growing animals, the amino acids required to maintain the
proinflammatory, acute-phase protein synthesis in splanchnic tissues impart a metabolic cost,
which results in suppression of muscle protein synthesis and increased catabolism and loss of
skeletal muscle mass. In addition to suppressed growth, the loss of lean body mass coupled
with disruption of organ function contributes to the increased morbidity and mortality associated
with infection.

3.3.2. Commensal microflora

The phenomenon described above illustrates an acute mechanism whereby the pathogenic
microflora activate the immune system, suppress the growth rate, induce fever and diarrhea,
and in some cases cause death in domestic animals. In the long term, however, over the life-
span of all mammals beginning at birth, animals are naturally colonized with commensal
microbes and these mostly bacterial species coexist with the host organism. This symbiosis
serves to activate the maturation of the immune system and enables nonruminant animals to
utilize dietary carbohydrate fermented to short-chain fatty acids from otherwise indigestible
plant polysaccharides and recycle body nitrogen when dietary protein availability is limited
(Fuller and Reeds, 1998; Hooper et al., 2002). Despite the general mutually beneficial rela-
tionship between commensal microbes and the host, there remains to a lesser degree some
activation of the immune system. The manifestations of commensal microbes have been
demonstrated most clearly in animals reared under germ-free compared to conventional envi-
ronments. These studies indicate that commensal microbes modestly reduced food intake,
increased metabolic rate, and increased mass and metabolic activity of gastrointestinal tissues
(Gaskins et al., 2002). The increase in gastrointestinal mass is associated with increased
numbers and activity of lymphoid cells and increased proliferation of epithelial crypt cells.
As with more severe, pathogenic infection, the presence of commensal bacteria results in
suppression of growth, albeit of lesser magnitude.
In the past fifty years, antimicrobial compounds have been fed to domestic animals in order
to suppress the activity of the gut microflora and enhance growth. Despite the widespread use
and success of antimicrobials, however, their exact mechanism of action remains poorly defined.
It has been shown that by suppressing microbial activity, antimicrobials reduce the luminal
concentration and associated toxic insult of ammonia, and thereby diminish the thickness and
Splanchnic protein and amino acid metabolism 179

mass of the intestinal mucosa and associated lymphoid tissue (Visek, 1978). Studies in pigs
and chickens have shown that feeding antimicrobial compounds significantly reduces liver
and small intestinal mass, cell proliferation, protein synthesis, and intestinal ammonia absorp-
tion (Muramatsu et al., 1983; Yen et al., 1987; Yen and Pond, 1990; Krinke and Jamroz,
1996). Additional evidence indicates that much of the luminal ammonia originates from
bacterial hydrolysis of urea and deamination of dietary amino acids. Thus, it appears that
part of the protein anabolic response of antimicrobials is associated with three phenomena:
(1) reduced microbial degradation of dietary essential amino acids, (2) increased intestinal
absorption of dietary amino acids, and (3) reduced utilization of dietary amino acids by
splanchnic tissues for maintenance of immune-associated lymphoid cells and acute-phase
protein synthesis.

3.3.3. Immunonutrients

During the proinflammatory, acute-phase response, the metabolic basis for increased splanch-
nic amino acid utilization has been linked to several cellular immune functions, including
synthesis of specific acute-phase proteins, glutathione synthesis and antioxidant function,
lymphocyte and mucosal crypt cell proliferation, and nitric oxide production. Researchers
have determined that a particular group of nutrients, termed “immunonutrients”, play a key
supportive role in these immune-related processes, serving to either act as biosynthetic sub-
strates or alter cellular function (Grimble, 2001; Huang et al., 2003). Among those considered
as immunonutrients are the sulfur amino acids, glutamine, arginine, nucleotides, and omega-
3 fatty acids. There are numerous acute-phase proteins synthesized by the liver, whose plasma
concentration increases significantly during infection and trauma, including C-reactive pro-
tein, serum amyloid A, fibrinogen, and haptoglobin. Cysteine is considered to be one of the
limiting amino acids for acute-phase protein synthesis, based on the estimated balance of
amino acids released from muscle proteolysis (Reeds et al., 1994). Cysteine is also a con-
stituent amino acid of glutathione and thioredoxin, both major cellular antioxidants (see
section below). Studies in infected and stressed rats indicate that cysteine utilization and GSH
synthesis and concentration in splanchnic tissues increased markedly (Malmezat et al., 1998,
2000a; Mercier et al., 2002). In protein-malnourished children and pigs, GSH concentrations
and synthesis rates in response to infection and stress are compromised, but can be restored
with cysteine supplementation (Jahoor et al., 1995; Reid et al., 2000; Badaloo et al., 2002).
The increased demand for cysteine during infection also markedly stimulated methionine
utilization via the transsulfuration pathway (Malmezat et al., 2000b). Recent studies provide
evidence that the proinflammatory cytokine, tumor necrosis factor-α, directly activates
cystathionine β-synthase activity, the enzyme catalyzing transsulfuration (Zou and Banerjee,
2003). In addition, methionine and cysteine are precursors for taurine, which also functions
as an important antioxidant in phagocytic cells (e.g. neutrophils) via stable neutralization of
intracellular hypochlorous acid (Santangelo, 2002). Hepatic taurine synthesis also increased
3-fold in infected rats (Malmezat et al., 2000a).
Glutamine and arginine also are considered immunonutrients because of their ability to
stimulate lymphoid and epithelial cell function and serve as precursors for nitric oxide, glu-
tathione, and nucleotide synthesis. Glutamine increased proliferation of lymphocytes and
intestinal crypt cells, increased macrophage phagocytic activity, reduced mucosal inflamma-
tory cytokine production, improved intestinal epithelial tight junction function, and increased
survival in bacterial-infected mice (Wilmore and Shabert, 1998; Huang et al., 2003).
Glutamine can be readily deaminated to glutamate in gut, liver, and lymphoid cells and can
180 D. G. Burrin and B. Stoll

also serve as a precursor for glutathione. Glutamine and glutamate nitrogen moieties are used
for nucleotide synthesis and nucleotides have been shown to enhance lymphoid cell function.
Sepsis has been shown to significantly increase uptake of glutamine by the liver, but not the
intestine (Karinch et al., 2001). Arginine deficiency induced by overexpression of intestinal
arginase I compromised the development of B-cell-related gut-associated lymphoid tissue
(de Jonge et al., 2002). Studies in endotoxemic pigs indicate that the liver and gut nitric oxide
production is increased in parallel with arginine utilization, whereas arginine is released from
muscle (Bruins et al., 2002a).

4. AMINO ACIDS AS EXTRACELLULAR SIGNALS


4.1. Glutamine

There is a clear recognition that extracellular amino acid availability has profound effects on
many aspects of cell function, including the control of cell signaling, gene expression, cell
volume, cell proliferation, apoptosis, and protein turnover. The precise cellular mechanisms
by which amino acids are able to elicit control over such diverse processes have become the
focus of intense investigation recently (fig. 6) (McDaniel et al., 2002; Averous et al., 2003;
Jefferson and Kimball, 2003; Kadowaki and Kanazawa, 2003; Meijer, 2003). Many of the
new developments in this area stem from the discovery of nutrient responsive genes and their
function in Drosophila and yeast systems. The effects of glutamine have been of particular
interest, since it has been shown to have pluripotent actions, including stimulation of cell pro-
liferation, protein synthesis, differentiation, ornithine decarboxylase (ODC) and immediate
early gene (c-jun) expression, and polyamine synthesis (Higashiguchi et al., 1993; Kandil
et al., 1995; Wang et al., 1996). In a recent review, Rhoads (1999) postulated that glutamine
stimulates cell proliferation by a signaling mechanism that involves the activation of two
related, but distinct, classes of mitogen-activated protein kinases (MAPK). Glutamine also
suppresses apoptosis and protein degradation in intestinal and liver tissue, which suggests that
it may be an important survival factor (Mortimore and Poso, 1987; Papaconstantinou et al.,
1998; Coeffier et al., 2003). Glutamine has been shown to suppress proteolysis in hepatocytes
in association with increased cell hydration and swelling, a process that involves the MAPK
pathways, extracellular-regulated kinases (ERK), and p38MAPK (Haussinger et al., 2001).
Recent studies also indicate that the antiapoptotic effect of glutamine involves an inhibitory
interaction between glutaminyl-tRNA synthetase and apoptosis signal-regulating kinase 1
(Ko et al., 2001).

4.2. Leucine

Leucine is another amino acid that has been shown to have important cell regulatory effects
in various tissues, including the liver and pancreatic beta-cells, yet there is limited informa-
tion of its effects in intestinal cells. Studies in perfused livers and hepatocytes demonstrated
that leucine was capable of suppressing proteolysis and stimulating protein synthesis
(Mortimore and Poso et al., 1987; Anthony et al., 2001a). In vivo studies also have shown that
feeding a leucine-deficient diet or leucine alone affects the expression of ribosomal protein
mRNA and activation of intracellular signaling molecules involved in protein synthesis
(Anthony et al., 2001b). In vivo studies in diabetic rats indicate that the effects of leucine
appear to be mediated, in part, via an insulin-independent pathway. The specific action
of leucine may be mediated by an interaction with a cell-membrane binding protein, since
Splanchnic protein and amino acid metabolism 181

cell-impermeant and nonmetabolizable analogs of leucine also suppress hepatic proteolysis and
stimulate protein synthesis (van Sluijters et al., 2000; Lynch, 2001; Lynch et al., 2002). Others
have suggested that leucine metabolism is important for cellular activation, based on reports that
BCAA ketoacids activate downstream signaling elements involved in protein synthesis.
The intracellular signaling pathway that mediates the effects of leucine on protein turnover
has been studied intensively in recent years. Amino acid deprivation, including leucine, acti-
vates the general control nonderepressing kinase 2 (GCN2), resulting in phosphorylation of
eukaryotic initiation factor-2 (eIF2), which leads to suppression of global protein synthesis
(Jefferson and Kimball, 2003). Specific deprivation of BCAA in cultured lymphocytes acti-
vates the branched-chain α-ketoacid dehydrogenase (BCKD) kinase, which translates into
reduced BCKD activity and BCAA catabolism (Doering and Danner, 2000). Other players in
the amino acid signaling pathway include those involved in assembly of initiation factors
(eIF4) and phosphorylation of ribosomal S6. A key intermediate in this pathway is the mam-
malian target of rapamycin (mTOR), which is a serine–threonine kinase involved in
coordinating nutrient availability with cell growth and proliferation. mTOR represents a point
of convergence in the pathways that mediate the stimulation of protein synthesis by amino
acids and insulin. Studies in hepatocytes show that leucine, glutamine, and insulin activate
p70 ribosomal S6 kinase, which is downstream of mTOR (Krause et al., 2002). However,
insulin also activates signals upstream from mTOR, namely phosphatidylinositol 3-kinase
(PI3K) and protein kinase B (PKB), whereas leucine and glutamine do not affect these
intermediates.
The factors that transduce the signal between leucine (or other effector amino acids) and
mTOR is the focus of intense interest. One such mechanism may involve adenosine
monophosphate-activated protein kinase (AMPK), which senses cellular AMP levels and is
activated under conditions of nutrient deprivation and ATP depletion. Activation of AMPK
leads to suppression of protein synthesis and mTOR phosphorylation and stimulation of
autophagic proteolysis (Kimura et al., 2003; Meijer, 2003). The role of AMPK is intriguing
in light of evidence that mTOR seems to function as a cellular ATP-sensing molecule (Dennis
et al., 2001). Another potential upstream activator of mTOR may be phosphatidic acid, which
is produced by the action of phospholipase D (Chen and Fang, 2002). This model of mTOR
activation by phosphatidic acid could form a link to other cellular signals such as the Rho
family of G proteins, protein kinase C, and intracellular calcium. Intracellular aminoacyl-tRNA
synthetases also have been implicated in the regulation of mTOR.

4.3. Sulfur amino acids

Sulfur amino acids (SAA), especially methionine and cysteine, play a key role in antioxidant
status and cellular function (Sen, 1998; Aw, 1999; Deplancke and Gaskins, 2002).
Glutathione and thioredoxin are the most important cellular antioxidants in mammals and
have a critical function in reacting with reactive oxygen species and maintaining cellular
redox status. Reduced glutathione (GSH) is an ubiquitous tripeptide (Glu-Cys-Gly) present
throughout the body at relatively high intracellular concentrations, especially in the small
intestine. Cellular GSH homeostasis is maintained through de novo synthesis from precursor
SAA methionine and cysteine, regeneration from its oxidized form glutathione disulfide
(GSSG), and uptake of extracellular intact GSH. Thioredoxin is another major cellular anti-
oxidant; it is a larger protein (12 kDa) than GSH, but also contains key cysteine residues in the
catalytic active site. Mediating oxidant stress and maintaining normal redox status is espe-
cially important in intestinal and liver cells. Studies with intestinal epithelial cells indicate
182 D. G. Burrin and B. Stoll

that increased oxidant stress and redox imbalance suppress cell proliferation and induce apop-
tosis and that this is closely correlated with a higher oxidized glutathione state, as measured
by the ratio of GSH : GSSG (Noda et al., 2001; Jonas et al., 2002; Pias and Aw, 2002). Culture
studies also show that cells grown in cysteine-deficient media have suppressed GSH concen-
trations and cell proliferation rates, both of which are stimulated with increased cysteine
supplementation (Miller et al., 2002; Noda et al., 2002). Other studies with human colonic
epithelial cells (Caco-2) indicate that, as differentiation proceeds, cell GSH concentration and
proliferation rate decrease, whereas apoptosis rate increases (Nkabyo et al., 2002).
Collectively, these studies suggest that cysteine availability and local GSH concentration have
a direct influence on epithelial cell proliferation and survival and are inversely proportional
to cellular differentiation state.
Given evidence that cysteine availability is important for maintenance of epithelial cell
GSH level and cell redox status, the question has been raised as to whether methionine can,
via transsulfuration, affect the cysteine availability. Evidence in support of this idea is the
finding that, in HepG2 cells, oxidant stress increased transsulfuration measured by cystathio-
nine synthesis and 35S-methionine incorporation into glutathione (Mosharov et al., 2000).
Moreover, studies in HepG2 cells also show that cystathionine synthase activity is coordi-
nately regulated with proliferation via a redox-sensitive mechanism (MacLean et al., 2002).
These results imply that cells exposed to oxidant stress may meet the increased cysteine
requirement for GSH synthesis via activation of methionine transsulfuration. A broader impli-
cation of these results is that methionine availability and its conversion to cysteine and GSH
via transsulfuration may be important for maintenance of normal intestinal and hepatic cell
proliferation and survival.

5. FUTURE PERSPECTIVES
The splanchnic tissues represent a quantitatively and functionally significant component of
whole-body protein and amino acid metabolism of growing animals. The gastrointestinal and
liver tissues consume a substantial fraction of the dietary amino acid intake for the purposes of
oxidative metabolism, protein synthesis, and gluconeogenesis. Relatively few nonessential
amino acids (glutamine, glutamate, aspartate) appear to be major oxidative fuels in the intes-
tine and this metabolism occurs in both epithelial and lymphoid cells. Although ammonia
appears to be the main fate of nitrogen derived from glutamine and glutamate catabolism,
incorporation of these nitrogen moieties into nucleotides and glucosamines also may be func-
tionally significant. However, some essential amino acids are also oxidized in the intestine, yet
the biochemical and cellular bases for this metabolism are poorly understood. Preliminary
studies suggest that essential amino acid catabolism occurs in epithelial cells, but further stud-
ies are warranted to examine the impact of the gut microflora on this process, especially in the
small intestine. Whether this phenomenon is a central mechanism explaining the growth-
promoting action for antimicrobials has yet to be proven. However, there is a compelling need
to address this question given the increasing concern about the impact of antimicrobials on the
environment and human health. Among the factors that regulate splanchnic tissue protein and
amino acid metabolism in healthy animals, nutrition plays a major role and it is apparent that
these tissues may be more responsive to extracellular nutrient availability than endocrine sig-
nals. However, under conditions of infection or stress, proinflammatory cytokines are key
activators of hepatic acute-phase protein synthesis, but stimulate protein catabolism in muscle.
This phenomenon serves to shuttle amino acids from muscle into hepatic protein synthesis, yet
it is not clear how the same cytokines differentially affect protein turnover in these two tissues.
Splanchnic protein and amino acid metabolism 183

The answers to many of these questions will require a greater understanding of the intra-
cellular signaling pathways that link both extracellular nutrient availability and endocrine
signals to cellular protein and amino acid metabolism. Despite the explosion of knowledge of
cellular signaling mechanisms in the past twenty years, there have been few efforts to explore
their relevance to protein and amino acid metabolism in splanchnic tissues, especially the
intestine. In the past, animal science has legitimately focused considerable attention on the
growth and metabolism of skeletal muscle and adipose tissue, given its relevance to lean
tissue growth. Yet, the splanchnic tissues act as key metabolic regulators of lean tissue growth,
serving to alter the peripheral availability of dietary nutrients and transmit nutrient-dependent
endocrine signals that determine peripheral tissue metabolism. Previous studies of physiology
and whole animal metabolism combined with clinical trials have shown that some nutrients
appear to be unique in their ability to ameliorate the metabolic effects of diseases. Moreover,
these nutrients seem to affect splanchnic tissue function, especially the immune system. Thus,
further study with these immunonutrients is necessary to establish their mechanism of action
and whether they can be used to improve growth and the health of domestic animals. An
expanding range of analytical tools, from the molecular to the whole-organ level, is now
becoming available to explore these issues in growing animals. The use of genomic and pro-
teomic analysis of individual cells and tissues coupled with genetic manipulation of animals
will aid in determining the essentiality and function of specific regulators of protein and
amino acid metabolism. In addition, the novel application of techniques such as germ-free
environments, mass isotopomer analysis, laser capture microdissection, and transorgan bal-
ance approaches should provide useful information on the localization of amino acid
metabolism in specific tissues and cells and how this is altered by the host environment.

REFERENCES
Adegoke, O.A., McBurney, M.I., Samuels, S.E., Baracos, V.E., 1999. Luminal amino acids acutely
decrease intestinal mucosal protein synthesis and protease mRNA in piglets. J. Nutr. 129, 1871–1878.
Adegoke, O.A., McBurney, M.I., Samuels, S.E., Baracos, V.E., 2003. Modulation of intestinal protein
synthesis and protease mRNA by luminal and systemic nutrients. Amer. J. Physiol. 284, G1017–G1026.
Ahlman, B., Charlton, M., Fu, A., Berg, C., O’Brien, P., Nair, K.S., 2001. Insulin’s effect on synthesis
rates of liver proteins: a swine model comparing various precursors of protein synthesis. Diabetes 50,
947–954.
Alpers, D.H., 1972. Protein synthesis in intestinal mucosa: the effect of route of administration of pre-
cursor amino acids. J. Clin. Invest. 51, 167–173.
Anthony, T.G., Anthony, J.C., Yoshizawa, F., Kimball, S.R., Jefferson, L.S., 2001a. Oral administration
of leucine stimulates ribosomal protein mRNA translation but not global rates of protein synthesis in
the liver of rats. J. Nutr. 131, 1171–1176.
Anthony, T.G., Reiter, A.K., Anthony, J.C., Kimball, S.R., Jefferson, L.S., 2001b. Deficiency of dietary
EAA preferentially inhibits mRNA translation of ribosomal proteins in liver of meal-fed rats. Amer.
J. Physiol. 281, E430–E439.
Argaman, Z., Young, V.R., Noviski, N., Castillo-Rosas, L., Lu, X.M., Zurakowski, D., Cooper, M.,
Davison, C., Tharakan, J.F., Ajami, A., Castillo, L., 2003. Arginine and nitric oxide metabolism in
critically ill septic pediatric patients. Crit. Care Med. 31, 591–597.
Attaix, D., Arnal, M., 1987. Protein synthesis and growth in the gastrointestinal tract of the young
preruminant lamb. Brit. J. Nutr. 58, 159–169.
Attaix, D., Meslin, J.C., 1991. Changes in small intestinal mucosa morphology and cell renewal in
suckling, prolonged-suckling, and weaned lambs. Amer. J. Physiol. 261, R811–R818.
Attaix, D., Manghebati, A., Grizard, J., Arnal, M., 1986. Assessment of in vivo protein synthesis in lamb
tissues with [3H]valine flooding doses. Biochim. Biophys. Acta 882, 389–397.
Attaix, D., Aurousseau, E., Rosolowska-Huszcz, D., Bayle, G., Arnal, M., 1992. In vivo longitudinal vari-
ations in protein synthesis in developing ovine intestines. Amer. J. Physiol. 263, R1318–R1323.
184 D. G. Burrin and B. Stoll

Averous, J., Bruhat, A., Mordier, S., Fafournoux, P., 2003. Recent advances in the understanding of
amino acid regulation of gene expression. J. Nutr. 133, 2040S–2045S.
Aw, T.Y., 1999. Molecular and cellular responses to oxidative stress and changes in oxidation-reduction
imbalance in the intestine. Amer. J. Clin. Nutr. 70, 557–565.
Badaloo, A., Reid, M., Forrester, T., Heird, W.C., Jahoor, F., 2002. Cysteine supplementation improves
the erythrocyte glutathione synthesis rate in children with severe edematous malnutrition. Amer. J. Clin.
Nutr. 76, 646–652.
Ball, R.O., Law, G., Bertolo, R.F.P., Pencharz, P.B., 1999. Adequate oral threonine is critical for mucin
production and mucosal growth by the neonatal piglet gut. VIIIth International Symposium on Protein
Metabolism and Nutrition, Aberdeen, UK, p. 31. (Abstr.)
Ballevre, O., Cadenhead, A., Calder, A.G., Rees, W.D., Lobley, G.E., Fuller, M.F., Garlick, P.J., 1990.
Quantitative partition of threonine oxidation in pigs: effect of dietary threonine. Amer. J. Physiol. 259,
E483–E491.
Baracos, V.E., Samuels, S.E., Adegoke, O.A., 2000. Anabolic and catabolic mediators of intestinal
protein turnover: a new experimental approach. Curr. Opin. Clin. Nutr. Metab. Care 3, 183–189.
Bark, T.H., McNurlan, M.A., Lang, C.H., Garlick, P.J., 1998. Increased protein synthesis after acute
IGF-I or insulin infusion is localized to muscle in mice. Amer. J. Physiol. 275, E118–E123.
Barle, H., Essen, P., Nyberg, B., Olivecrona, H., Tally, M., McNurlan, M.A., Wernerman, J., Garlick, P.J.,
1999. Depression of liver protein synthesis during surgery is prevented by growth hormone. Amer.
J. Physiol. 276, E620–Ε627.
Barle, H., Rahlen, L., Essen, P., McNurlan, M.A., Garlick, P.J., Holgersson, J., Wernerman, J., 2001.
Stimulation of human albumin synthesis and gene expression by growth hormone treatment. Clin.
Nutr. 20, 59–67.
Barrett, E.J., Gusberg, R., Ferrannini, E., Tepler, J., Felig, P., Jacob, R., Smith, D., DeFronzo, R.A., 1986.
Amino acid and glucose metabolism in the postabsorptive state and following amino acid ingestion
in the dog. Metabolism 35, 709–717.
Battezzati, A., Brillon, D.J., Matthews, D.E., 1995. Oxidation of glutamic acid by the splanchnic bed in
humans. Amer. J. Physiol. 269, E269–E276.
Battezzati, A., Haisch, M., Brillon, D.J., Matthews, D.E., 1999. Splanchnic utilization of enteral alanine
in humans. Metabolism 48, 915–921.
Beaufrere, B., Fournier, V., Salle, B., Putet, G., 1992. Leucine kinetics in fed low-birth-weight infants:
importance of splanchnic tissues. Amer. J. Physiol. 263, E214–E220.
Beauloye, V., Ketelslegers, J.M., Moreau, B., Thissen, J.P., 1999. Dexamethasone inhibits both growth
hormone (GH)-induction of insulin-like growth factor-I (IGF-I) mRNA and GH receptor (GHR)
mRNA levels in rat primary cultured hepatocytes. Growth Horm. IGF Res. 9, 205–211.
Bertolo, R.F., Chen, C.Z., Law, G., Pencharz, P.B., Ball, R.O., 1998. Threonine requirement of neonatal
piglets receiving total parenteral nutrition is considerably lower than that of piglets receiving an iden-
tical diet intragastrically. J. Nutr. 128, 1752–1759.
Bertolo, R.F., Brunton, J.A., Pencharz, P.B., Ball, R.O., 2003. Arginine, ornithine, and proline intercon-
version is dependent on small intestinal metabolism in neonatal pigs. Amer. J. Physiol. Endocrinol.
Metab. 284, E915–E922.
Blachier, F., Darcy-Vrillon, B., Sener, A., Duee, P.H., Malaisse, W.J., 1991. Arginine metabolism in rat
enterocytes. Biochim. Biophys. Acta. 1092, 304–310.
Blachier, F., M’Rabet-Touil, H., Posho, L., Morel, M.T., Bernard, F., Darcy-Vrillon, B., Duee, P.H.,
1992. Polyamine metabolism in enterocytes isolated from newborn pigs. Biochim. Biophys. Acta
1175, 21–26.
Blachier, F., Selamnia, M., Robert, V., M’Rabet-Touil, H., Duee, P.H., 1995. Metabolism of L-arginine
through polyamine and nitric oxide synthase pathways in proliferative or differentiated human colon
carcinoma cells. Biochim. Biophys. Acta 1268, 255.
Blemings, K.P., Gahl, M.J., Crenshaw, T.D., Benevenga, N.J., 1996. Recombinant bovine somatotropin
decreases hepatic amino acid catabolism in female rats. J. Nutr. 126, 1657–1661.
Blommaart, P.J., Zonneveld, D., Meijer, A.J., Lamers, W.H., 1993. Effects of intracellular amino acid
concentrations, cyclic AMP, and dexamethasone on lysosomal proteolysis in primary cultures of peri-
natal rat hepatocytes. J. Biol. Chem. 268, 1610–1617.
Blouin, J.P., Bernier, J.F., Reynolds, C.K., Lobley, G.E., Dubreuil, P., Lapierre, H., 2002. Effect of supply
of metabolizable protein on splanchnic fluxes of nutrients and hormones in lactating dairy cows.
J. Dairy Sci. 85, 2618–2630.
Splanchnic protein and amino acid metabolism 185

Boirie, Y., Gachon, P., Beaufrere, B., 1997. Splanchnic and whole-body leucine kinetics in young and
elderly men. Amer. J. Clin. Nutr. 65, 489–495.
Boirie, Y., Short, K.R., Ahlman, B., Charlton, M., Nair, K.S., 2001. Tissue-specific regulation of mito-
chondrial and cytoplasmic protein synthesis rates by insulin. Diabetes 50, 2652–2658.
Boni-Schnetzler, M., Schmid, C., Meier, P.J., Froesch, E.R., 1991. Insulin regulates insulin-like growth
factor I mRNA in rat hepatocytes. Amer. J. Physiol. 260, E846–E851.
Bos, C., Stoll, B., Fouillet, H., Gaudichon, C., Guan, X., Grusak, M.A., Reeds, P.J., Tome, D., Burrin, D.G.,
2003. Intestinal lysine metabolism is driven by the enteral availability of dietary lysine in piglets fed
a bolus meal. Amer. J. Physiol. Endocrinol. Metab. 285, E1246–E1257.
Bouteloup-Demange, C., Claeyssens, S., Maillot, C., Lavoinne, A., Lerebours E., Dechelotte P., 2000.
Effects of enteral glutamine on gut mucosal protein synthesis in healthy humans receiving glucocor-
ticoids. Amer. J. Physiol. 278, G677–G681.
Boza, J.J., Jahoor, F., Reeds, P.J., 1996. Ribonucleic acid nucleotides in maternal and fetal tissues derive
almost exclusively from synthesis de novo in pregnant mice. J. Nutr. 126, 1749–1758.
Boza, J.J., Turini, M., Moennoz, D., Montigon, F., Vuichoud, J., Gueissaz, N., Gremaud, G., Pouteau, E.,
Piguet-Welsch, C., Finot, P.A., Ballevre, O., 2001. Effect of glutamine supplementation of the diet on
tissue protein synthesis rate of glucocorticoid-treated rats. Nutrition 17, 35–40.
Brameld, J.M., 1997. Molecular mechanisms involved in the nutritional and hormonal regulation of
growth in pigs. Proc. Nutr. Soc. 56, 607–619.
Brameld, J.M., Gilmour, R.S., Buttery, P.J., 1999. Glucose and amino acids interact with hormones to
control expression of insulin-like growth factor-I and growth hormone receptor mRNA in cultured pig
hepatocytes. J. Nutr. 129, 1298–1306.
Breuille, D., Arnal, M., Rambourdin, F., Bayle, G., Levieux, D., Obled, C., 1998. Sustained modifications of
protein metabolism in various tissues in a rat model of long-lasting sepsis. Clin. Sci. 94, 413–423.
Breuille, D., Voisin, L., Contrepois, M., Arnal, M., Rose, F., Obled, C., 1999. A sustained rat model for
studying the long-lasting catabolic state of sepsis. Infect. Immun. 67, 1079–1085.
Brosnan, J.T., 2000. Glutamate, at the interface between amino acid and carbohydrate metabolism.
J. Nutr. 130, 988S–990S.
Brosnan, J.T., 2003. Interorgan amino acid transport and its regulation. J. Nutr. 133, 2068S–2072S.
Brosnan, J.T., Ewart, H.S., Squires, S.A., 1995. Hormonal control of hepatic glutaminase. Adv. Enzyme
Regul. 35, 131–146.
Bruins, M.J., Lamers, W.H., Meijer, A.J., Soeters, P.B., Deutz, N.E., 2002a. In vivo measurement of nitric
oxide production in porcine gut, liver and muscle during hyperdynamic endotoxaemia.
Brit. J. Pharmacol. 137, 1225–1236.
Bruins, M.J., Soeters, P.B., Lamers, W.H., Deutz, N.E., 2002b. L-arginine supplementation in pigs
decreases liver protein turnover and increases hindquarter protein turnover both during and after
endotoxemia. Am. J. Clin. Nutr. 75, 1031–1044.
Brunton, J.A., Bertolo, R.F.P., Pencharz, P.B., Ball, R.O., 1999. L-Arginine supplementation in pigs
decreases liver protein turnover and increases hindquarter protein turnover both during and after
endotoxemia. Amer. J. Physiol. 277, E223–E231.
Burrin, D.G., Ferrell, C.L., Eisemann, J., Britton, R.A., Nienaber, J.A., 1989. Effect of level of nutrition
on splanchnic blood flow and oxygen consumption in sheep. Brit. J. Nutr. 62, 23–34.
Burrin, D.G., Davis, T.A., Fiorotto, M.L., Reeds, P.J., 1991. Stage of development and fasting affect
protein synthetic activity in the gastrointestinal tissues of suckling pigs. J. Nutr. 121, 1099–1108.
Burrin, D.G., Shulman, R.J., Reeds, P.J., Davis, T.A., Gravitt, K.R., 1992. Porcine colostrum and milk
stimulate visceral organ and skeletal muscle synthesis in neonatal piglets. J. Nutr. 122, 1205–1213.
Burrin, D.G., Shulman, R.J., Langston, C., Storm, M.C., 1994. Supplemental alanylglutamine, organ
growth and nitrogen metabolism in neonatal pigs fed by total parenteral nutrition. J. Parent. Enter.
Nutr. 18, 313–319.
Burrin, D.G., Davis, T.A., Ebner, S., Schoknecht, P.A., Fiorotto, M.L., Reeds, P.J., McAvoy, S., 1995.
Nutrient-independent and nutrient-dependent factors stimulate protein synthesis in colostrum-fed
newborn pigs. Pediat. Res. 37, 593–599.
Burrin, D.G., Davis, T.A., Ebner, S., Schoknecht, P.A., Fiorotto, M.L., Reeds, P.J., 1997. Colostrum enhances
the nutritional stimulation of vital organ protein synthesis in neonatal pigs. J. Nutr. 127, 1284–1289.
Burrin, D.G., Fiorotto, M.L., Hadsell, D.L, 1999a. Transgenic hypersection of des(1–3) human insulin-
like growth factor I in mouse milk has limited effects on the gastrointestinal tract in suckling pups.
J. Nutr. 129, 51–56.
186 D. G. Burrin and B. Stoll

Burrin, D.G., Stoll, B., Chang, X., Yu, H., Reeds, P.J., 1999b. Total parenteral nutrition does not com-
promise digestion or absorption of dietary protein in neonatal pigs. FASEB J. 13, A1024.
Burrin, D.G., Wester, T.J., Davis, T.A., Fiorotto, M.L., Chang, X., 1999c. Dexamethasone inhibits small
intestinal growth via increased protein catabolism in neonatal pigs. Amer. J. Physiol. 276, E269–E277.
Burrin, D.G., Stoll, B., Jiang, R., Petersen, Y., Elnif, J., Buddington, R.K., Schmidt, M., Holst, J.J.,
Hartmann, B., Sangild, P.T., 2000a. GLP-2 stimulates intestinal growth in premature TPN-fed pigs by
suppressing proteolysis and apoptosis. Amer. J. Physiol. 279, G1249–G1256.
Burrin, D.G., Stoll, B., Jiang, R., Chang, X., Hartmann, B., Holst, J.J., Greeley, G.H. Jr., Reeds, P.J.,
2000b. Minimal enteral nutrient requirements for intestinal growth in neonatal piglets: how much is
enough? Amer. J. Clin. Nutr. 71, 1603–1610.
Burrin, D.G., Stoll, B., Fan, M.Z., Dudley, M.A., Donovan, S.M., Reeds, P.J., 2001. Oral IGF-I alters the
posttranslational processing but not the activity of lactase-phlorizin hydrolase in formula-fed neonatal
pigs. J. Nutr. 131, 2235–2241.
Burrin, D.G., Stoll, B., Chang, X., van Goudoever, J.B., Fujii, H., Hutson, S., Reeds, P.J., 2003a.
Parenteral nutrition results in impaired lactose digestion and hexose absorption when enteral feeding
is initiated in infant pigs. Amer. J. Clin. Nutr. 78, 461–470.
Burrin, D.G., Stoll, B., Guan, X., 2003b. Glucagon-like peptide 2 function in domestic animals. Domest.
Anim. Endocrinol. 24, 103–122.
Bush, J.A., Wu, G., Suryawan, A., Nguyen, H.V., Davis, T.A., 2002. Somatotropin-induced amino acid
conservation in pigs involves differential regulation of liver and gut urea cycle enzyme activity.
J. Nutr. 132, 59–67.
Bush, J.A., Burrin, D.G., Suryawan, A., O’Connor, P.M., Nguyen, H.V., Reeds, P.J., Steele, N.C.,
van Goudoever, J.B., Davis, T.A., 2003a. Somatotropin-induced protein anabolism in hindquarters
and portal-drained viscera of growing pigs. Amer. J. Physiol. 284, E302–E312.
Bush, J.A., Kimball, S.R., O’Connor, P.M., Suryawan, A., Orellana, R.A., Nguyen, H.V., Jefferson, L.S.,
Davis, T.A., 2003b. Translational control of protein synthesis in muscle and liver of growth hormone-
treated pigs. Endocrinology 144, 1273–1283.
Butler, A.A., Yakar, S., LeRoith, D., 2002. Insulin-like growth factor-I: compartmentalization within the
somatotropic axis? News Physiol. Sci. 17, 82–85.
Buts, J.P., De Keyser, N., De Raedemaeker, L., Collette, E., Sokal, E.M., 1995. Polyamine profiles in human
milk, infant artificial formulas, and semi-elemental diets. J. Pediat. Gastroenterol. Nutr. 21, 44–49.
Castillo, L., Chapman, T.E., Yu, Y.M., Ajami, A., Burke, J.F., Young, V.R., 1993a. Dietary arginine uptake
by the splanchnic region in adult humans. Amer. J. Physiol. 265, E532–E539.
Castillo, L., de Rojas, T.C., Chapman, T.E., Vogt, J., Burke, J.F., Tannenbaum, S.R., Young, V.R., 1993b.
Splanchnic metabolism of dietary arginine in relation to nitric oxide synthesis in normal adult man.
Proc. Natl. Acad. Sci. USA 90, 193–197.
Castillo, L., Beaumier, L., Ajami, A.M., Young, V.R., 1996. Whole body nitric oxide synthesis in healthy men
determined from 15N arginine-to-15N citrulline labeling. Proc. Natl. Acad. Sci. USA 93, 11460–11465.
Cayol, M., Boirie, Y., Rambourdin, F., Prugnaud, J., Gachon, P., Beaufrere, B., Obled, C., 1997. Influence of pro-
tein intake on whole body and splanchnic leucine kinetics in humans. Amer. J. Physiol. 272, E584–E591.
Charlton, M., Ahlman, B., Nair, K.S., 2000. The effect of insulin on human small intestinal mucosal
protein synthesis. Gastroenterology 118, 299–306.
Chen, J., Fang, Y., 2002. A novel pathway regulating the mammalian target of rapamycin (mTOR)
signaling. Biochem. Pharmacol. 64, 1071.
Coeffier, M., Claeyssens, S., Hecketsweiler, B., Lavoinne, A., Ducrotte, P., Dechelotte, P., 2003. Enteral
glutamine stimulates protein synthesis and decreases ubiquitin mRNA level in human gut mucosa.
Amer. J. Physiol. 285, G266–G273.
Coloso, R.M., Stipanuk, M.H., 1989. Metabolism of cyst(e)ine in rat enterocytes. J. Nutr. 119, 1914–24.
Crenn, P., Vahedi, K., Lavergne-Slove, A., Cynober, L., Matuchansky, C., Messing, B., 2003. Plasma cit-
rulline: a marker of enterocyte mass in villous atrophy-associated small bowel disease.
Gastroenterology 124, 1210–1219.
Dahms, W.T., Owens, R.P., Kalhan, S.C., Kerr, D.S., Danish, R.K., 1989. Urea synthesis, nitrogen bal-
ance, and glucose turnover in growth-hormone-deficient children before and after growth hormone
administration. Metabolism 38, 197–203.
Darcy-Vrillon, B., Posho, L., Morel, M.-T., Bernard, F., Blachier, F., Meslin, J.-C., Duee, P.-H., 1994.
Glucose, galactose, and glutamine metabolism in pig isolated enterocytes during development. Pediat.
Res. 36, 175–181.
Splanchnic protein and amino acid metabolism 187

Davis, T.A., Nguyen, H.V., Garcia-Bravo, R., Fiorotto, M.L., Jackson, E.M., Lewis, D.S., Lee, D.R.,
Reeds, P.J., 1994. Amino acid composition of human milk is not unique. J. Nutr. 124, 1126–1132.
Davis, T.A., Burrin, D.G., Fiorotto, M.L., Nguyen, H.V., 1996. Protein synthesis in skeletal muscle
and jejunum is more responsive to feeding in 7- than in 26-day-old pigs. Amer. J. Physiol. 270,
E802–E809.
Davis, T.A., Fiorotto, M.L., Beckett, P.R., Burrin, D.G., Reeds, P.J., Wray-Cahen, D., Nguyen, H.V.,
2001. Differential effects of insulin on peripheral and visceral tissue protein synthesis in neonatal
pigs. Amer. J. Physiol. 280, E770–E779.
Davis, T.A., Fiorotto, M.L., Burrin, D.G., Reeds, P.J., Nguyen, H.V., Beckett, P.R., Vann, R.C.,
O’Connor, P.M., 2002a. Stimulation of protein synthesis by both insulin and amino acids is unique to
skeletal muscle in neonatal pigs. Amer. J. Physiol. 282, E880–E890.
Davis, T.A., Fiorotto, M.L., Burrin, D.G., Vann, R.C., Reeds, P.J., Nguyen, H.V., Beckett, P.R., Bush, J.A.,
2002b. Acute IGF-I infusion stimulates protein synthesis in skeletal muscle and other tissues of
neonatal pigs. Amer. J. Physiol. 283, E638–E647.
de Blaauw, I., Deutz, N.E., von Meyenfeldt, M.F., 1996. In vivo amino acid metabolism of gut and liver
during short and prolonged starvation. Amer. J. Physiol. 270, G298–G306.
de Jonge, W.J., Kwikkers, K.L., te Velde, A.A., van Deventer, S.J., Nolte, M.A., Mebius, R.E., Ruijter, J.M.,
Lamers, M.C., Lamers, W.H., 2002. Arginine deficiency affects early B cell maturation and lymphoid
organ development in transgenic mice. J. Clin. Invest. 110, 1539–1548.
Del Roso, A., Vittorini, S., Cavallini, G., Donati, A., Gori, Z., Masini, M., Pollera, M., Bergamini, E.,
2003. Ageing-related changes in the in vivo function of rat liver macroautophagy and proteolysis.
Exp. Gerontol. 38, 519–527.
Dennis, P.B., Jaeschke, A., Saitoh, M., Fowler, B., Kozma, S.C., Thomas, G., 2001. Mammalian TOR:
a homeostatic ATP sensor. Science 294, 1102–1105.
Deplancke, B., Gaskins, H.R., 2002. Redox control of the transsulfuration and glutathione biosynthesis
pathways. Curr. Opin. Clin. Nutr. Metab. Care 5, 85–92.
Doering, C.B., Danner, D.J., 2000. Amino acid deprivation induces translation of branched-chain
α-ketoacid dehydrogenase kinase. Amer. J. Physiol. 279, C1587–C1594.
Douglas, R.G., Gluckman, P.D., Ball, K., Breier, B., Shaw, J.H., 1991. The effects of infusion of insulin-
like growth factor (IGF) I, IGF-II, and insulin on glucose and protein metabolism in fasted lambs.
J. Clin. Invest. 88, 614–622.
Drucker, D.J., 2002. Biological actions and therapeutic potential of the glucagon-like peptides.
Gastroenterology 122, 531–544.
Dudley, M.A., Burrin, D.G., Wykes, L.J., Toffolo, G., Cobelli, C., Nichols, B.L., Rosenberger, J., Jahoor, F.,
Reeds, P.J., 1998. Protein kinetics determined in vivo with a multiple-tracer, single-sample protocol:
application to lactase synthesis. Amer. J. Physiol. 274, G591–G598.
Dugan, M.E.R., Knabe, D.A., Wu, G., 1995. The induction of citrulline synthesis from glutamine in
enterocytes of weaned pigs is not due primarily to age or change in diet. J. Nutr. 125, 2388–2393.
Ebner, S., Schoknecht, P., Reeds, P.J., Burrin, D.G., 1994. Growth and metabolism of gastrointestinal
and skeletal muscle tissues in protein malnourished neonatal pigs. Amer. J. Physiol. 266,
R1736–R1743.
Elango, R., Pink, D., Pencharz, P.B., Ball, R.O., 2003. Branched chain amino acid catabolism in the
neonatal piglet gut. In: Ball, R.O. (Ed.), Proceedings of the 9th International Symposium on Digestive
Physiology in Pigs. University of Alberta, Department of Agriculture, Food, and Nutritional Science,
Edmonton, Alberta, Vol. 2, p. 207–209.
Elwyn, D.H., Parikh, H.C., Shoemaker, W.C., 1968. Amino acid movements between gut, liver, and
periphery in unanesthetized dogs. Amer. J. Physiol. 215, 1260–1275.
Etherton, T.D., Bauman, D.E., 1998. Biology of somatotropin in growth and lactation of domestic animals.
Physiol. Rev. 78, 745–761.
Faure, M., Moennoz, D., Montigon, F., Fay, L.B., Breuille, D., Finot, P.A., Ballevre, O., Boza, J., 2002.
Development of a rapid and convenient method to purify mucins and determine their in vivo synthe-
sis rate in rats. Anal. Biochem. 307, 244–251.
Feldhoff, R.C., Taylor, J.M., Jefferson, L.S., 1977. Synthesis and secretion of rat albumin in vivo, in per-
fused liver, and in isolated hepatocytes. J. Biol. Chem. 252, 3611–3616.
Felig, P., 1975. Amino acid metabolism in man. Annu. Rev. Biochem. 44, 933.
Finkelstein, J.D., 2000. Pathways and regulation of homocysteine metabolism in mammals. Semin.
Thromb. Hemost. 26, 219–225.
188 D. G. Burrin and B. Stoll

Fouillet, H., Gaudichon, C., Bos, C. Mariotti, F., Tome, D., 2003. Contribution of plasma proteins to
splanchnic and total anabolic utilization of dietary nitrogen in humans. Amer. J. Physiol. Endocrinol
Metab. 285, E88–E97.
Fuller, M.F., Reeds, P.J., 1998. Nitrogen cycling in the gut. Annu. Rev. Nutr. 18, 385–411.
Gahl, M.J., Benevenga, N.J., Crenshaw, T.D., 1998. Rates of lysine catabolism are inversely related to
rates of protein synthesis when measured concurrently in adult female rats induced to grow at differ-
ent rates. J. Nutr. 128, 1503–1511.
Garcia-Arumi, E., Schwartz, S., Lopez-Hellin, J., Arbos, M.A., Andreu, A.L., Farriol, M., 1995. Addition
of glutamine does not improve protein synthesis and jejunal mucosa morphology in non-hypercatabolic
stress. Physiol. Res. 44, 233–239.
Garlick, P.J., McNurlan, M.A., Preedy, V.R., 1980. A rapid and convenient technique for measuring the
rate of protein synthesis in tissues by injection of 3H-phenylalanine. Biochem. J. 192, 719–723.
Garlick, P.J., McNurlan, M.A., Essen, P., Wernerman, J., 1994. Measurement of tissue protein synthesis
rates in vivo: a critical analysis of contrasting methods. Amer. J. Physiol. 266, E287–E297.
Gaskins, H.R., Collier, C.T., Anderson, D.B., 2002. Antibiotics as growth promotants: mode of action.
Anim. Biotechnol. 13, 29–42.
Goldspink, D.F., Kelly, F.J., 1984. Protein turnover and growth in the whole body, liver and kidney of the
rat from the foetus to senility. Biochem. J. 217, 507–516.
Goldspink, D.F., Lewis, S.E.M., Kelly, F.J., 1984. Protein synthesis during the developmental growth of
the small and large intestine of the rat. Biochem. J. 217, 527–534.
Grimble, R.F., 2001. Nutritional modulation of immune function. Proc. Nutr. Soc. 60, 389.
Grofte, T., Wolthers, T., Jensen, S.A., Moller, N., Jorgensen, J.O., Tygstrup, N., Orskov, H., Vilstrup, H.,
1997. Effects of growth hormone and insulin-like growth factor-I singly and in combination on in vivo
capacity of urea synthesis, gene expression of urea cycle enzymes, and organ nitrogen contents in rats.
Hepatology 25, 964–969.
Guan, X., Stoll, B., Lu, X., Tappenden, K.A., Holst, J.J., Hartmann, B., Burrin, D.G., 2003. GLP-2-mediated
up-regulation of intestinal blood flow and glucose uptake is NO-dependent in TPN-fed piglets.
Gastroenterology 125, 136–147.
Hahn, P., Taller, M., Chan, H., 1988. Pyruvate carboxylase, phosphate-dependent glutaminase and gluta-
mate dehydrogenase in the developing rat small intestinal mucosa. Biol. Neonate 53, 362–366.
Haisch, M., Fukagawa, N.K., Matthews, D.E., 2000. Oxidation of glutamine by the splanchinic bed in
humans. Amer. J. Physiol. Endocrinol. Metab. 278, E593.
Hallemeesch, M.M., ten Have, G.A., Deutz, N.E., 2001. Metabolic flux measurements across portal
drained viscera, liver, kidney and hindquarter in mice. Lab. Anim. 35, 101–110.
Hallemeesch, M.M., Lamers, W.H., Deutz, N.E., 2002. Reduced arginine availability and nitric oxide
production. Clin. Nutr. 21, 273–279.
Halseth, A.E., Flakoll, P.J., Reed, E.K., Messina, A.B., Krishna, M.G., Lacy, D.B., Williams, P.E.,
Wasserman, D.H., 1997. Effect of physical activity and fasting on gut and liver proteolysis in the dog.
Amer. J. Physiol. 273, E1073–E1082.
Hamel, F.G., Upward, J.L., Siford, G.L., Duckworth, W.C., 2003. Inhibition of proteasome activity by
selected amino acids. Metabolism 52, 810–814.
Haque, S.M., Chen, K., Usui, N., Iiboshi, Y., Okuyama, H., Masunari, A., Cui, L., Nezu, R., Takagi, Y.,
Okada, A., 1996. Alanyl-glutamine dipeptide-supplemented parenteral nutrition improves intestinal
metabolism and prevents increased permeability in rats. Ann. Surg. 223, 334–341.
Harp, J.B., Goldstein, S., Phillips, L.S., 1991. Nutrition and somatomedin. XXIII. Molecular regulation
of IGF-I by amino acid availability in cultured hepatocytes. Diabetes 40, 95–101.
Haussinger, D., 1990. Nitrogen metabolism in liver: structural and functional organization and physio-
logical relevance. Biochem. J. 267, 281–290.
Haussinger, D., 1996. The role of cellular hydration in the regulation of cell function. Biochem. J. 313,
697–710.
Haussinger, D., Schliess, F., 1999. Osmotic induction of signaling cascades: role in regulation of cell
function. Biochem. Biophys. Res. Commun. 255, 551–555.
Haussinger, D., Graf, D., Weiergraber, O.H., 2001. Glutamine and cell signaling in liver. J. Nutr. 131, 2509S.
Heitmann, R.N., Bergman, E.N., 1978. Glutamine metabolism, interorgan transport and glucogenicity in
the sheep. Amer. J. Physiol. 234, E197–E203.
Hellstern, G., Kaempf-Rotzoll, D.D., Linderkamp, O.O., Langhans, K.D., Rating, D., 2002. Parenteral
amino acids increase albumin and skeletal muscle protein fractional synthetic rates in premature
newborn minipigs. J. Pediat. Gastroenterol. Nutr. 35, 270–274.
Splanchnic protein and amino acid metabolism 189

Higashiguchi, T., Hasselgren, P.O., Wagner, K., Fischer, J.E., 1993. Effect of glutamine on protein syn-
thesis in isolated intestinal epithelial cells. J. Parent. Enter. Nutr. 17, 307–314.
Higashiguchi, T., Noguchi, Y., O’Brien, W., Wagner, K., Fischer, J.E., Hasselgren, P.O., 1994. Effect of sepsis
on mucosal protein synthesis in different parts of the gastrointestinal tract in rats. Clin. Sci. 87, 207.
Hodgkinson, S.M., Moughan, P.J., Reynolds, G.W., James, K.A., 2000. The effect of dietary peptide
concentration on endogenous ileal amino acid loss in the growing pig. Brit. J. Nutr. 83, 421–430.
Hoerr, R.A., Matthews, D.E., Bier, D.M., Young, V.R., 1991. Leucine kinetics from 2H3- and 13C-leucine
infused simultaneously by gut and vein. Amer. J. Physiol. 260, E111–E117.
Hooper, L.V., Midtvedt, T., Gordon. J.I., 2002. How host-microbial interactions shape the nutrient envi-
ronment of the mammalian intestine. Annu. Rev. Nutr. 22, 283–307.
Hopgood, M.F., Clark, M.G., Ballard, F.J., 1981. Stimulation by glucocorticoids of protein degradation
in hepatocyte monolayers. Biochem. J. 196, 33–40.
House, J.D., Hall, B.N., Brosnan, J.T., 2001. Threonine metabolism in isolated rat hepatocytes. Amer.
J. Physiol. 281, E1300–E1307.
Huang, Y.S., Chuang, D.T., 1999. Down-regulation of rat mitochondrial branched-chain 2-oxoacid
dehydrogenase kinase gene expression by glucocorticoids. Biochem. J. 339, 503–510.
Huang, Y., Shao, X.M., Neu, J., 2003. Immunonutrients and neonates. Eur. J. Pediat. 162, 122–128.
Humbert, B., Le Bacquer, O., Nguyen, P., Dumon, H., Darmaun, D., 2001. Protein restriction and
dexamethasone as a model of protein hypercatabolism in dogs: effect of glutamine on leucine
turnover. Metabolism 50, 293–298.
Hutson, S.M., Stinson-Fisher, C., Shiman, R., Jefferson, L.S., 1987. Regulation of albumin synthesis by
hormones and amino acids in primary cultures of rat hepatocytes. Amer. J. Physiol. 252, E291–E298.
Iannoli, P., Miller, J.H., Ryan, C.K., Sax, H.C., 1998. Glucocorticoids upregulate intestinal nutrient trans-
port in a time-dependent and substrate-specific fashion. J. Gastrointest. Surg. 2, 449–457.
Inoue, Y., Grant, J.P., Snyder, P.J., 1993. Effect of glutamine-supplemented total parenteral nutrition on
recovery of the small intestine after starvation atrophy. J. Parent. Enter. Nutr. 17, 165–170.
Jacobs, R.L., Stead, L.M., Brosnan, M.E., Brosnan, J.T., 2001. Hyperglucagonemia in rats results in
decreased plasma homocysteine and increased flux through the transsulfuration pathway in liver.
J. Biol. Chem. 276, 43740–43747.
Jahoor, F., Wykes, L.J., Reeds, P.J., Henry, J.F., del Rosario, M.P., Frazer, M.E., 1995. Protein-deficient
pigs cannot maintain reduced glutathione homeostasis when subjected to the stress of inflammation.
J. Nutr. 125, 1462–1472.
Jefferson, L.S., Kimball, S.R., 2003. Amino acids as regulators of gene expression at the level of mRNA
translation. J. Nutr. 133, 2046S–2051S.
Jefferson, L.S., Schworer, C.M., Tolman, E.L., 1975. Growth hormone stimulation of amino acid trans-
port and utilization by the perfused rat liver. J. Biol. Chem. 250, 197–204.
Jepson, M.M., Pell, J.M., Bates, P.C., Millward, D.J., 1986. The effects of endotoxaemia on protein
metabolism in skeletal muscle and liver of fed and fasted rats. Biochem. J. 235, 329–336.
Jiang, R., Chang, X., Stoll, B., Ellis, K.J., Shypailo, R.J., Weaver, E., Campbell, J., Burrin, D.G., 2000.
Dietary plasma protein is used more efficiently than extruded soy protein for lean tissue growth in
early-weaned pigs. J. Nutr. 130, 2016–2019.
Johnson, R.W., 1997. Inhibition of growth by pro-inflammatory cytokines: an integrated view. J. Anim.
Sci. 75, 1244–1255.
Johnson, R.W., 2002. The concept of sickness behavior: a brief chronological account of four key
discoveries. Vet. Immunol. Immunopathol. 87, 443–450.
Jonas, C.R., Ziegler, T.R., Gu, L.H., Jones, D.P., 2002. Extracellular thiol/disulfide redox state affects pro-
liferation rate in a human colon carcinoma (Caco2) cell line. Free Radical Biol. Med. 33, 1499–1506.
Jungas, R.L., Halperin, M.L., Brosnan, J.T., 1992. Quantitative analysis of amino acid oxidation and
related gluconeogenesis in humans. Physiol. Rev. 72, 419.
Kadowaki, M., Kanazawa, T., 2003. Amino acids as regulators of proteolysis. J. Nutr. 133, 2052S–2056S.
Kandil, H.M., Argenzio, R.A., Chen, W., Berschneider, H.M., Stiles, A.D., Westwick, J.K., Rippe, R.A.,
Brenner, D.A., Rhoads, J.M., 1995. L-Glutamine and L-asparagine stimulate ODC activity and prolif-
eration in a porcine jejunal enterocyte line. Amer. J. Physiol. 269, G591–G599.
Karinch, A.M., Pan, M., Lin, C.M., Strange, R., Souba, W.W., 2001. Glutamine metabolism in sepsis and
infection. J. Nutr. 131, 2535S–2538S.
Khan, J., Iiboshi, Y., Cui, L., Wasa, M., Sando, K., Takagi, Y., Okada, A., 1999. Alanyl-glutamine-
supplemented parenteral nutrition increases luminal mucus gel and decreases permeability in the rat
small intestine. J. Parent. Enter. Nutr. 23, 24–31.
190 D. G. Burrin and B. Stoll

Kight, C.E., Fleming, S.E., 1995. Oxidation of glucose carbon entering the TCA cycle is reduced by glu-
tamine in small intestine epithelial cells. Amer. J. Physiol. 268, G879–G888.
Kimball, S.R., Jefferson, L.S., 1994, Mechanisms of translational control in liver and skeletal muscle.
Biochimie 76, 729–736.
Kimball, S.R., Jefferson, L.S., 2001. Regulation of protein synthesis by branched-chain amino acids.
Curr. Opin. Clin. Nutr. Metab. Care 4, 39–43.
Kimball, S.R., Horetsky, R.L., Jefferson, L.S., 1995. Hormonal regulation of albumin gene expression in
primary cultures of rat hepatocytes. Amer. J. Physiol. 268, E6–E14.
Kimura, N., Tokunaga, C., Dalal, S., Richardson, C., Yoshino, K., Hara, K., Kemp, B.E., Witters, L.A.,
Mimura, O., Yonezawa, K., 2003. A possible linkage between AMP-activated protein kinase (AMPK)
and mammalian target of rapamycin (mTOR) signalling pathway. Genes Cells 8, 65–79.
Ko, Y.G., Kim, E.Y., Kim, T., Park, H., Park, H.S., Choi, E.J., Kim, S., 2001. Glutamine-dependent anti-
apoptotic interaction of human glutaminyl-tRNA synthetase with apoptosis signal-regulating kinase 1.
J. Biol. Chem. 276, 6030–6036.
Koea, J.B., Douglas, R.G., Breier, B.H., Shaw, J.H., Gluckman, P.D., 1992. Synergistic effect of insulin-
like growth factor-I administration on the protein-sparing effects of total parenteral nutrition in fasted
lambs. Endocrinology 131, 643–648.
Krause, U., Bertrand, L., Maisin, L., Rosa, M., Hue, L., 2002. Signalling pathways and combinatory
effects of insulin and amino acids in isolated rat hepatocytes. Eur. J. Biochem. 269, 3742–3750.
Krinke, A.L., Jamroz, D., 1996. Signalling pathways and combinatory effects of insulin and amino acids
in isolated rat hepatocytes. Poultry Sci. 75, 705–710.
Lambert, B.D., Titgemeyer, E.C., Stokka, G.L., DeBey, B.M., Loest, C.A., 2002. Methionine supply to
growing steers affects hepatic activities of methionine synthase and betaine-homocysteine methyl-
transferase, but not cystathionine synthase. J. Nutr. 132, 2004–2009.
Lapierre, H., Bernier, J.F., Dubreuil, P., Reynolds, C.K., Farmer, C., Ouellet, D.R., Lobley, G.E., 1999.
The effect of intake on protein metabolism across splanchnic tissues in growing beef steers. Brit. J. Nutr.
81, 457–466.
Lash, L.H., Hagen, T.M., Jones, D.P., 1986. Exogenous glutathione protects intestinal epithelial cells
from oxidative injury. Proc. Natl. Acad. Sci. USA 83, 4641–4645.
Le Bacquer, O., Nazih, H., Blottiere, H., Meynial-Denis, D., Laboisse, C., Darmaun, D., 2001. Effects of
glutamine deprivation on protein synthesis in a model of human enterocytes in culture. Amer.
J. Physiol. 281, G1340–G1347.
Le Bacquer, O., Laboisse, C., Darmaun, D., 2003. Glutamine preserves protein synthesis and paracellu-
lar permeability in Caco-2 cells submitted to “luminal fasting”. Amer. J. Physiol. 285, G128–G136.
Le Floc’h, N.L., Obled, C., Seve, B., 1996. In vivo threonine oxidation in growing pigs fed on diets with
graded levels of threonine. Brit. J. Nutr. 75, 825–837.
Le Floc’h, N., Thibault, J.N., Seve, B., 1997. Tissue localization of threonine oxidation in pigs. Brit.
J. Nutr. 77, 593–603.
Ling, P.R., Gollaher, C., Colon, E., Istfan, N., Bistrian, B.R., 1995. IGF-I alters energy expenditure and
protein metabolism during parenteral feeding in rats. Amer. J. Clin. Nutr. 61, 116–120.
Lo, H.C., Ney, D.M., 1996. GH and IGF-I differentially increase protein synthesis in skeletal muscle and
jejunum of parenterally fed rats. Amer. J. Physiol. 271, E872–E878.
Lobley, G.E., Milne, V., Lovie, J.M., Reeds, P.J., Pennie, K., 1980. Whole body and tissue protein
synthesis in cattle. Brit. J. Nutr. 43, 491–502.
Lobley, G.E., Harris, P.M., Skene, P.A., Brown, D., Milne, E., Calder, A.G., Anderson, S.E., Garlick, P.J.,
Nevison, I., Connell, A., 1992. Responses in tissue protein synthesis to sub- and supra-maintenance
intake in young growing sheep: comparison of large-dose and continuous-infusion techniques. Brit.
J. Nutr. 68, 373–388.
Lobley, G.E., Connell, A., Revell, D.K., Bequette, B.J., Brown, D.S., Calder, A.G., 1996a. Splanchnic-
bed transfers of amino acids in sheep blood and plasma, as monitored through use of a multiple
U-13C-labelled amino acid mixture. Brit. J. Nutr. 75, 217–235.
Lobley, G.E., Connell, A., Revell, D., 1996b. The importance of transmethylation reactions to
methionine metabolism in sheep: effects of supplementation with creatine and choline. Brit. J. Nutr.
75, 47–56.
Lobley, G.E., Shen, X., Le, G., Bremner, D.M., Milne, E., Calder, A.G., Anderson, S.E., Dennison, N.,
2003. Oxidation of essential amino acids by the ovine gastrointestinal tract. Brit. J. Nutr. 89,
617–630.
Splanchnic protein and amino acid metabolism 191

Luiking, Y.C., Deutz, N.E., 2003. Isotopic investigation of nitric oxide metabolism in disease. Curr. Opin.
Clin. Nutr. Metab. Care 6, 103–108.
Lynch, C.J., 2001. Role of leucine in the regulation of mTOR by amino acids: revelations from structure-
activity studies. J. Nutr. 131, 861S–865S.
Lynch, C.J., Hutson, S.M., Patson, B.J., Vaval, A., Vary, T.C., 2002. Tissue-specific effects of chronic dietary
leucine and norleucine supplementation on protein synthesis in rats. Amer. J. Physiol. 283, E824–E835.
Mack, D.R., Michail, S., Wei, S., McDougall, L., Hollingsworth, M.A., 1999. Probiotics inhibit
enteropathogenic E. coli adherence in vitro by inducing intestinal mucin gene expression. Amer.
J. Physiol. 39, G941–G950.
MacLean, K.N., Janosik, M., Kraus, E., Kozich, V., Allen, R.H., Raab, B.K., 2002. Cystathionine
β-synthase is coordinately regulated with proliferation through a redox-sensitive mechanism in
cultured human cells and Saccharomyces cerevisiae. J. Cell. Physiol. 192, 81–92.
MacRae, J.C., 1993. Metabolic consequences of intestinal parasitism. Proc. Nutr. Soc. 52, 121–130.
Madej, M., Lundh, T., Lindberg, J.E., 1999. Activities of enzymes involved in glutamine metabolism in
connection with energy production in the gastrointestinal tract epithelium of newborn, suckling and
weaned piglets. Biol. Neonate 75, 250–258.
Malmezat, T., Breuille, D., Pouyet, C., Mirand, P.P., Obled, C., 1998. Metabolism of cysteine is modified
during the acute phase of sepsis in rats. J. Nutr. 128, 97–105.
Malmezat, T., Breuille, D., Capitan, P., Mirand, P.P., Obled, C., 2000a. Glutathione turnover is increased
during the acute phase of sepsis in rats. J. Nutr. 130, 1239–1246.
Malmezat, T., Breuille, D., Pouyet, C., Buffiere, C., Denis, P., Mirand, P.P., Obled, C., 2000b. Methionine
transsulfuration is increased during sepsis in rats. Amer. J. Physiol. 279, E1391–E1397.
Marchini, J.S., Nguyen, P., Deschamps, J.-Y., Maugere, P., Krempf, M., Darmaun, D., 1999. Effect
of intratvenous glutamine on duodenal mucosa protein synthesis in healthy growing dogs. Amer.
J. Physiol. 276, E747–E753.
Martensson, J., Jain, A., Meister, A., 1991. Glutathione is required for intestinal function. Proc. Natl.
Acad. Sci. USA, 87, 1715–1719.
Matthews, D.E., Marano, M.A., Campbell, R.G., 1993. Splanchnic bed utilization of leucine and pheny-
lalanine in humans. Amer. J. Physiol. 264, E109–E118.
McCowen, K.C., Bistrian, B.R., 2003. Immunonutrition: problematic or problem solving? Amer. J. Clin.
Nutr. 77, 764–770.
McDaniel, M.L., Marshall, C.A., Pappan, K.L., Kwon, G., 2002. Metabolic and autocrine regulation of
the mammalian target of rapamycin by pancreatic beta-cells. Diabetes 51, 2877–2885.
McNurlan, M.A., Garlick, P.J., 1981. Protein synthesis in liver and small intestine in protein deprivation
and diabetes. Amer. J. Physiol. 241, E238–E245.
McNurlan, M.A., Tomkins, A.M., Garlick, P.J., 1979. The effect of starvation on the rate of protein-
synthesis in rat liver and small intestine. Biochem. J. 178, 373–379.
Meek, S.E., Persson, M., Ford, G.C., Nair, K.A., 1998. Differential regulation of amino acid exchange
and protein dynamics across splanchnic and skeletal muscle beds by insulin in healthy human
subjects. Diabetes 47, 1824–1835.
Meijer, A.J., 2003. Amino acids as regulators and components of nonproteinogenic pathways. J. Nutr.
133, 2057S–2062S.
Mercier, S., Breuille, D., Mosoni, L., Obled, C., Patureau Mirand, P., 2002. Chronic inflammation alters
protein metabolism in several organs of adult rats. J. Nutr. 132, 1921–1928.
Merry, B.J., Lewis, S.E., Goldspink, D.F., 1992. The influence of age and chronic restricted feeding on
protein synthesis in the small intestine of the rat. Exp. Gerontol. 27, 191–200.
Metges, C.C., 2000. Contribution of microbial amino acids to amino acid homeostasis of the host. J. Nutr.
130, 1857S–1864S.
Miller, L.T., Watson, W.H., Kirlin, W.G., Ziegler, T.R., Jones, D.P., 2002. Oxidation of the glutathione
disulfide redox state is induced by cysteine deficiency in human colon carcinoma HT29 cells. J. Nutr.
132, 2303–2306.
Morris, S.M., 2002. Regulation of enzymes of the urea cycle and arginine metabolism. Annu. Rev. Nutr.
22, 87–105.
Mortimore, G.E., Poso, A.R., 1987. Intracellular protein catabolism and its control during nutrient
deprivation and supply. Annu. Rev. Nutr. 7, 539–564.
Mortimore, G.E., Poso, A.R., Lardeux, B.R., 1989. Mechanism and regulation of protein degradation in
liver. Diabetes Metab. Rev. 5, 49–70.
192 D. G. Burrin and B. Stoll

Mosharov, E., Cranford, M.R., Banerjee, R., 2000. The quantitatively important relationship between
homocysteine metabolism and glutathione synthesis by the transsulfuration pathway and its regula-
tion by redox changes. Biochemistry 39, 13005–13011.
Mosoni, L., Houlier, M.L., Mirand, P.P., Bayle, G., Grizard, J., 1993. Effect of amino acids alone or with
insulin on muscle and liver protein synthesis in adult and old rats. Amer. J. Physiol. 264, E614–E620.
Muramatsu, T., Coates, M.E., Hewitt, D., Salter, D.N., Garlick, P.J., 1983. The influence of the gut
microflora on protein synthesis in liver and jejunal mucosa in chicks. Brit. J. Nutr. 49, 453–462.
Murphy, J.M., Murch, S.J., Ball, R.O., 1996. Proline is synthesized from glutamate during intragastric
infusion but not during intravenous infusion in neonatal piglets. J. Nutr. 26, 878–886.
Nair, K.S., Ford, G.C., Ekberg, K., Fernqvist-Forbes, E., Wahren, J., 1995. Protein dynamics in whole
body and in splanchnic and leg tissues in type I diabetic patients. J. Clin. Invest. 95, 2926–2937.
Naka, S., Saito, H., Hashiguchi, Y., Lin, M.T., Furukawa, S., Inaba, T., Fukushima, R., Wada, N., Muto, T.,
1997. Alanylglutamine-enriched total parenteral nutrition improves protein metabolism more than
branched chain amino acid-enriched total parenteral nutrition in protracted peritonitis. J. Trauma 42,
183–190.
Newsholme, P., Procopio, J., Lima, M.M., Pithon-Curi, T.C., Curi, R., 2003. Glutamine and glutamate:
their central role in cell metabolism and function. Cell Biochem. Funct. 21, 1–9.
Nieto, R., Lobley, G.E., 1999. Integration of protein metabolism within the whole body and between
organs. In: Lobley, G.E., White, A., MacRae, J.C. (Eds.), Proceedings of the 8th International
Symposium on Protein Metabolism and Nutrition. EAAP Publication No. 96, p. 69.
Nissim, I., Brosnan, M.E., Yudkoff, M., Brosnan, J.T., 1999. Studies of hepatic glutamine metabolism in
the perfused rat liver with 15N-labeled glutamine. J. Biol. Chem. 274, 28958–28965.
Nkabyo, Y.S., Ziegler, T.R., Li, H.G., Watson, W.H., Jones, D.P., 2002. Glutathione and thioredoxin redox
during differentiation in human colon epithelial (Caco-2) cells. Amer. J. Physiol. 283, G1352–G1359.
Noda, T., Iwakiri, R., Fujimoto, K., Aw, T.Y., 2001. Induction of mild intracellular redox imbalance
inhibits proliferation of CaCo-2 cells. FASEB J. 15, 2131–2139.
Noda, T., Iwakiri, R., Fujimoto, K., Rhoads, C.A., Aw, T.Y., 2002. Exogenous cysteine and cystine pro-
mote cell proliferation in CaCo-2 cells. Cell Prolif. 35, 117–129.
Nolan, E.M., Masters, J.N., Dunn, A., 1990. Growth hormone regulation of hepatic glutamine synthetase
mRNA levels in rats. Mol. Cell Endocrinol. 69, 101–110.
Nyachoti, C.M., de Lange, C.F., McBride, B.W., Leeson, S., Gabert, V.M., 2000. Endogenous gut nitro-
gen losses in growing pigs are not caused by increased protein synthesis rates in the small intestine.
J. Nutr. 130, 566–572.
Nygren, J., Nair, K.S., 2003. Differential regulation of protein dynamics in splanchnic and skeletal
muscle beds by insulin and amino acids in healthy human subjects. Diabetes 52, 1377–1385.
Odedra, B.R., Bates, P.C., Millward, D.J., 1983. Time course of the effect of catabolic doses of corticos-
terone on protein turnover in rat skeletal muscle and liver. Biochem. J. 214, 617–627.
Ogier-Denis, E., Codogno, P., 2003. Autophagy: a barrier or an adaptive response to cancer. Biochim.
Biophys. Acta 1603, 113–128.
O’Leary, M.J., Koll, M., Ferguson, C.N., Coakley, J.H., Hinds, C.J., Preedy, V.R., Garlick, P.J., 2003.
Liver albumin synthesis in sepsis in the rat: influence of parenteral nutrition, glutamine and growth
hormone. Clin. Sci. (Lond.) 105, 691–698.
O’Sullivan, D., Brosnan, J.T., Brosnan, M.E., 2000. Catabolism of arginine and ornithine in the perfused
rat liver: effect of dietary protein and of glucagon. Amer. J. Physiol. 278, E516–E521.
Palmer, R.M., Pusztai, A., Bain, P., Grant, G., 1987. Changes in rates of tissue protein synthesis in rats
induced in vivo by consumption of kidney bean lectins. Comp. Biochem. Physiol. 88C, 179.
Papaconstantinou, H.T., Hwang, K.O., Rajaraman, S., Hellmich, M.R., Townsend, C.M. Jr., Ko, T.C.,
1998. Glutamine deprivation induces apoptosis in intestinal epithelial cells. Surgery 124,
152–160.
Pell, J.M., Bates, P.C., 1992. Differential actions of growth hormone and insulin-like growth factor-I on
tissue protein metabolism in dwarf mice. Endocrinology 130, 1942–1950.
Pias, E.K., Aw, T.Y., 2002. Apoptosis in mitotic competent undifferentiated cells is induced by cellular
redox imbalance independent of reactive oxygen species production. FASEB J. 16, 781–790.
Pink, D., Elango, R., Dixon, W.T., Ball, R.O., 2003. Intestinal catabolism of lysine in swine.
In: Ball, R.O. (Ed.), Proceedings of the 9th International Symposium on Digestive Physiology in Pigs.
University of Alberta, Department of Agriculture, Food and Nutritional Science, Edmonton, Alberta,
Vol. 2, pp. 131–133.
Splanchnic protein and amino acid metabolism 193

Platell, C., McCauley, R., McCulloch, R., Hall, J., 1993. The influence of parenteral glutamine and
branched-chain amino acids on total parenteral nutrition-induced atrophy of the gut. J. Parent. Enter.
Nutr. 17, 348–354.
Pollack, P.F., Koldovsky, O., Nishioka, K., 1992. The influence of parenteral glutamine and branched-chain
amino acids on total parenteral nutrition-induced atrophy of the gut. Amer. J. Clin. Nutr. 56, 371–375.
Ponter, A.A., Cortamira, N.O., Seve, B., Salter, D.N., Morgan, L.M., 1994. The effects of energy source
and tryptophan on the rate of protein synthesis and on hormones of the entero-insular axis in the
piglet. Brit. J. Nutr. 71, 661–674.
Preedy, V.R., Paska, L., Sugden, P.H., Schofield, P.S., Sugden, M.C., 1988. The effects of surgical stress
and short-term fasting on protein synthesis in vivo in diverse tissues of the mature rat. Biochem.
J. 250, 179–188.
Ratnam, S., Maclean, K.N., Jacobs, R.L., Brosnan, M.E., Kraus, J.P., Brosnan, J.T., 2002. Hormonal
regulation of cystathionine β-synthase expression in liver. J. Biol. Chem. 277, 42912–42918.
Reeds, P.J., Fjeld, C.R., Jahoor, F., 1994. Do the differences between the amino acid compositions
of acute-phase and muscle proteins have a bearing on nitrogen loss in traumatic states? J. Nutr.
124, 906–910.
Reeds, P.J., Burrin, D.G., Stoll, B., Jahoor, F., Wykes, L., Henry, J., Frazer, M.E., 1997. Enteral glutamate
is the preferential source for mucosal glutathione synthesis in fed piglets. Amer. J. Physiol.
273, E408–E415.
Reeds, P.J., Burrin, D.G., Stoll, B., van Goudoever, J.B., 1999. Consequences and regulation of gut
metabolism. In: Lobley, G.E., White, A., MacRae, J.C. (Eds.), Protein Metabolism and Nutrition.
EAAP Publication No. 96, p. 127.
Reid, M., Badaloo, A., Forrester, T., Morlese, J.F., Frazer, M., Heird, W.C., Jahoor, F., 2000. In vivo rates
of erythrocyte glutathione synthesis in children with severe protein-energy malnutrition. Amer.
J. Physiol. 278, E405–E412.
Rerat, A., Simoes-Nunes, C., Mendy, F., Vaissade, P., Vaugelade, P., 1992. Splanchnic fluxes of amino
acids after duodenal infusion of carbohydrate solutions containing free amino acids or oligopeptides
in the non-anaesthetized pig. Brit. J. Nutr. 68, 111–138.
Rhoads, M., 1999. Glutamine signaling in intestinal cells. J. Parent. Enter. Nutr. 23, S38–S40.
Saenz de Pipaon, M., VanBeek, R.H., Quero, J., Perez, J., Wattimena, D.J., Sauer, P.J., 2003. Effect of
minimal enteral feeding on splanchnic uptake of leucine in the postabsorptive state in preterm infants.
Pediat. Res. 53, 281–287.
Salleh, M., Ardawi, M., Majzoub, M.F., Newsholme, E.A., 1988. Effect of glucocorticoid treatment
on glucose and glutamine metabolism by the small intestine of the rat. Clin. Sci. (Lond.) 75,
93–100.
Samuels, S.E., Taillandier, D., Aurousseau, E., Cherel, Y., Maho, L., Arnal, M., Attaix, D., 1996.
Gastrointestinal tract protein synthesis and mRNA levels for proteolytic systems in adult fasted rats.
Amer. J. Physiol. 271, E232–E238.
Santangelo, F., 2002. The regulation of sulphurated amino acid junctions: fact or fiction in the field of
inflammation? Amino Acids 23, 359–365.
Schliess, F., Haussinger, D., 2002. The cellular hydration state: a critical determinant for cell death and
survival. Biol. Chem. 383, 577–583.
Schliess, F., Haussinger, D., 2003. Cell volume and insulin signaling. Int. Rev. Cytol. 225, 187–228.
Sen, C.K., 1998. Redox signaling and the emerging therapeutic potential of thiol antioxidants. Biochem.
Pharmacol. 55, 1747–1758.
Seve, I., Reeds, P.J., Fuller, M.S., Cadenhead, A., Hay, S.M., 1986. Protein synthesis and retention in
some tissues of the young pig as influenced by dietary protein intake after early-weaning: possible
connection to the energy metabiolism. Reprod. Nutr. Dev. 26, 849–861.
Shenoy, V., Roig, J.C., Chakrabarti, R., Kubilis, P., Neu, J., 1996. Ontogeny of glutamine synthetase in
rat small intestine. Pediat. Res. 39, 643–648.
Shoveller, A.K., Brunton, J.A., Pencharz, P.B., Ball, R.O., 2003. The methionine requirement is lower in
neonatal piglets fed parenterally than in those fed enterally. J. Nutr. 133, 1390–1397.
Souba, W.W., Smith, R.J., Wilmore, D.W., 1985. Effects of glucocorticoids on glutamine metabolism in
visceral organs. Metabolism 34, 450–456.
Southon, S., Livesey, G., Gee, J.M., Johnson, I.T., 1985. Differences in intestinal protein synthesis
and cellular proliferation in well-nourished rats consuming conventional laboratory diets. Brit. J. Nutr.
53, 87–95.
194 D. G. Burrin and B. Stoll

Southorn, B.G., Palmer, R.M., Garlick, P.J., 1990. Acute effects of corticosterone on tissue protein
synthesis and insulin-sensitivity in rats in vivo. Biochem. J. 272, 187–191.
Spaeth, G., Gottwald, T., Haas, W., Holmer, M., 1993. Glutamine peptide does not improve gut
barrier function and mucosal immunity in total parenteral nutrition. J. Parent. Enter. Nutr. 17,
317–323.
Stegink, L.D., Den Besten, L., 1972. Synthesis of cysteine from methionine in normal adult subjects:
effect of route of alimentation. Science 178, 514–516.
Stein, H.H., Nyachoti, M., 2003. Animal effects on ileal amino acid digestibility. In: Ball, R. (Ed.),
Proceedings of the 9th International Symposium on Digestive Physiology in Pigs. University
of Alberta, Department of Agriculture, Food, and Nutritional Science, Edmonton, Alberta. Vol. 2,
p. 223.
Stein, T.P., Yoshida, S., Schluter, M.D., Drews, D., Assimon, S.A., Leskiw, M.D., 1994. Comparison of
intravenous nutrients on gut mucosal protein synthesis. J. Parent. Enter. Nutr. 18, 447–452.
Stipanuk, M.H., Rotter, M.A., 1984. Metabolism of cysteine, cysteinesulfinate and cysteinesulfonate in
rats fed adequate and excess levels of sulfur-containing amino acids. J. Nutr. 114, 1426–1437.
Stoll, B., Henry, J., Reeds, P.J., Yu, H., Jahoor, F., Burrin, D.G., 1998. Catabolism dominates the
first-pass intestinal metabolism of dietary essential amino acids in milk protein-fed piglets. J. Nutr.
128, 606–614.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1999a. Substrate oxidation by the portal
drained viscera of fed piglets. Amer. J. Physiol. 277, E168–E175.
Stoll, B., Burrin, D.G., Henry, J., Jahoor, F., Reeds, P.J., 1999b. Dietary and systemic phenylalanine
utilization for mucosal and hepatic constitutive protein synthesis in pigs. Amer. J. Physiol. 276,
G49–G57.
Stoll, B., Chang, X., Fan, M.Z., Reeds, P.J., Burrin, D.G., 2000a. Enteral nutrient intake level
determines intestinal protein synthesis and accretion rates in neonatal pigs. Amer. J. Physiol. 279,
G288–G294.
Stoll, B., Chang, X., Jiang, R., van Goudoever, J.B., Reeds, P.J., Burrin, D.G., 2000b. Enteral carbohy-
drate and lipid inhibit small intestinal proteolysis in neonatal pigs. FASEB J. 14, A558.
Stoll, B., Price, P., Reeds, P.J., Henry, J., Burrin, D.G., 2000c. Feeding an elemental diet versus a milk-
based formula does not decrease intestinal mucosal growth in infant pigs. J. Pediat. Gastroenterol.
Nutr. 31, S169.
Stubbs, A.K., Wheelhouse, N.M., Lomax, M.A., Hazlerigg, D.G., 2002. Nutrient-hormone interaction in
the ovine liver: methionine supply selectively modulates growth hormone-induced IGF-I gene expres-
sion. J. Endocrinol. 174, 335–341.
Tamada, H., Nezu, R., Imamura, I., Matsuo, Y., Takagi, Y., Kamata, S., Okada, A., 1992. The dipeptide
alanyl-glutamine prevents intestinal mucosal atrophy in parenterally fed rats. J. Parent. Enter. Nutr.
16, 110–116.
Tashiro, T., Sugiura, T., Morishima, Y., Shimoda, N., Yamamori, H., Takagi, K., Nakajima, N.,
1999. Effect of IGF-1 on protein metabolism in burned rats. J. Parent. Enter. Nutr. 23, 5 Suppl., S93.
Torrallardona, D., Harris, C.I., Fuller, M.F., 2003a. Lysine synthesized by the gastrointestinal microflora
of pigs is absorbed, mostly in the small intestine. Amer. J. Physiol. 284, E1177–E1180.
Torrallardona, D., Harris, C.I., Fuller, M.F., 2003b. Pigs’ gastrointestinal microflora provide them with
essential amino acids. J. Nutr. 133, 1127–1131.
van der Schoor, S.R.D., van Goudeoever, J.B., Stoll, B., Henry, J.F., Rosenberger, J.R., Burrin, D.G.,
Reeds, P.J., 2001. The pattern of intestinal substrate oxidation is altered by protein restriction in pigs.
Gastroenterology 121, 1167–1175.
van der Schoor, S.R., Reeds, P.J., Stoll, B., Henry, J.F., Rosenberger, J.R., Burrin, D.G., van Goudoever,
J.B., 2002. The high metabolic cost of a functional gut. Gastroenterology 123, 1931–1940.
van Goudoever, J.B., Stoll, B., Henry, J.F., Burrin, D.G., Reeds, P.J., 2000. Adaptive regulation of intes-
tinal lysine metabolism. Proc. Natl. Acad. Sci. 97, 11620–11625.
van Klinken, B.J., Dekker, J., Büller, H.A., de Bols, C., Einerhand, A.W., 1997. Biosynthesis of mucins
(MUC2-6) along the longitudinal axis of the human gastrointestinal tract. Amer. J. Physiol. 273,
G296–G302.
van Sluijters, D.A., Dubbelhuis, P.F., Blommaart, E.F., Meijer, A.J., 2000. Amino-acid-dependent signal
transduction. Biochem. J. 351, 545–550.
Vary, T.C., Kimball, S.R., 1992. Regulation of hepatic protein synthesis in chronic inflammation and
sepsis. Amer. J. Physiol. 262, C445–C452.
Splanchnic protein and amino acid metabolism 195

Visek, W.J., 1978. The mode of growth promotion by antibiotics. J. Anim. Sci. 46, 1447.
von Allmen, D., Hasselgren, P.-O., Higashiguchi, T., Frederick, J., Zamir, O., Fischer, J.E., 1992.
Increased intestinal protein synthesis during sepsis and following the administration of tumour necro-
sis factor α or interleukin-1α. Biochem. J. 286, 585–589.
Wakabayashi, Y., Yamada, E., Yoshida, T., Takahashi, N., 1995. Effect of intestinal resection and arginine-
free diet on rat physiology. Amer. J. Physiol. 269, G313–G318.
Wang, J.-Y., Viar, M.J., Blanner, P.M., Johnson, L.R., 1996. Expression of the ornithine decar-
boxylase gene in response to asparagine in intestinal epithelial cells. Amer. J. Physiol. 271,
G164–G171.
Wang, Q., Meyer, T.A., Boyce, S.T., Wang, J.J., Sun, X., Tiao, G., Fischer, J.E., Hasselgren, P.-O., 1998.
Endotoxemia in mice stimulates production of complement C3 and serum amyloid A in mucosa of
small intestine. Amer. J. Physiol. 275, R1584–R1592.
Weber, F.L. Jr., Fresard, K.M., Veach, G.L., 1989. Stimulation of jejunal mucosal protein synthesis by
luminal glucose: effects with luminal and vascular leucine in fed and fasted rats. Gastroenterology 96,
935–937.
Wester, T.J., Davis, T.A., Fiorotto, M.L., Burrin, D.G., 1998. Exogenous growth hormone stimulates
somatotropic axis function and growth in neonatal pigs. Amer. J. Physiol. 274, E29–E37.
Wilmore, D.W., Shabert, J.K., 1998. Role of glutamine in immunologic responses. Nutrition 14,
618–626.
Windmueller, H.G., Spaeth, A.E., 1974. Uptake and metabolism of plasma glutamine by the small intes-
tine. J. Biol. Chem. 249, 5070–5079.
Windmueller, H.G., Spaeth, A.E., 1975. Intestinal metabolism of glutamine and glutamate from the
lumen as compared to glutamine from blood. Arch. Biochem. Biophys. 171, 662–672.
Windmueller, H.G., Spaeth, A.E., 1976. Metabolism of absorbed aspartate, asparagine, and arginine by
rat small intestine in vivo. Arch. Biochem. Biophys. 175, 670–676.
Windmueller, H.G., Spaeth, A.E., 1978. Identification of ketone bodies and glutamine as the major
respiratory fuels in vivo for postabsorptive rat small intestine. J. Biol. Chem. 253, 69–76.
Windmueller, H.G., Spaeth, A.E., 1980. Respiratory fuels and nitrogen metabolism in vivo in small
intestine of fed rats: quantitative importance of glutamine, glutamate, and aspartate. J. Biol. Chem.
255, 107–112.
Wolff, J.E., Bergman, E.N., Williams, H.H., 1972. Net metabolism of plasma amino acids by liver and
portal-drained viscera of fed sheep. Amer. J. Physiol. 223, 438–446.
Wu, G., 1997. Synthesis of citrulline and arginine from proline in enterocytes of postnatal pigs. Amer.
J. Physiol. 272, G1382–G1390.
Wu, G., 1998. Intestinal mucosal amino acid catabolism. J. Nutr. 128, 1249–1252.
Wu, G., Morris, S.M. Jr. 1998. Arginine metabolism: nitric oxide and beyond. Biochem. J. 336,
1–17.
Wu, G., Knabe, D.A., Yan, W., Flynn, N.E., 1995. Glutamine and glucose metabolism in enterocytes of
the neonatal pig. Amer. J. Physiol. 268, R334–R342.
Wu, G., Flynn, N.E., Knabe, D.A., 2000a. Enhanced intestinal synthesis of polyamines from proline in
cortisol-treated piglets. Amer. J. Physiol. 279, E395–E402.
Wu, G., Flynn, N.E., Knabe, D.A., Jaeger, L.A., 2000b. A cortisol surge mediates the enhanced
polyamine synthesis in procine enterocytes during weaning. Amer. J. Physiol. 279, R554–R559.
Wu, G., Haynes, T.E., Li, H., Yan, W., Meininger, C.J., 2001. Glutamine metabolism to glucosamine
is necessary for glutamine inhibition of endothelial nitric oxide synthesis. Biochem. J. 353,
245–252.
Wykes, L.J., Fiorotto, M., Burrin, D.G., Del Rosario, M., Frazer, M.E., Pond, W.G., Jahoor, F., 1996.
Chronic low protein intake reduces tissue protein synthesis in a pig model of protein malnutrition.
J. Nutr. 126, 1481–1488.
Yen, J.T., Pond, W.G., 1990. Effect of carbadox on net absorption of ammonia and glucose into hepatic
portal vein of growing pigs. J. Anim. Sci. 68, 4236–4242.
Yen, J.T., Nienaber, J.A., Pond, W.G., 1987. Effect of neomycin, carbadox and length of adaptation to
calorimeter on performance, fasting metabolism and gastrointestinal tract of young pigs. J. Anim. Sci.
65, 1243–1248.
Yen, J.T., Nienaber, J.A., Hill, D.A., Pond, W.G., 1989. Oxygen consumption by portal vein-
drained organs and by whole animal in conscious growing swine. Proc. Soc. Exp. Biol. Med. 190,
393–398.
196 D. G. Burrin and B. Stoll

Yu, Y.M., Young, V.R., Tompkins, R.G., Burke, J.F., 1995. Comparative evaluation of the quantitative
utilization of parenterally and enterally administered leucine and L-[1-13C,15N]leucine within the
whole body and the splanchnic region. J. Parent. Enter. Nutr. 19, 209–215.
Zafarullah, M., Li, W.Q., Sylvester, J., Ahmad, M., 2003. Molecular mechanisms of N-acetylcysteine
actions. Cell. Mol. Life Sci. 60, 6–20.
Zou, C.G., Banerjee, R., 2003. Tumor necrosis factor-α-induced targeted proteolysis of cystathionine
β-synthase modulates redox homeostasis. J. Biol. Chem. 278, 16802–16808.
8 Nitrogen metabolism by splanchnic tissues
of ruminants

C. K. Reynolds

Department of Animal Sciences, The Ohio State University, OARDC,


1680 Madison Avenue, Wooster, OH 44691-4096, USA

Ruminant amino acid metabolism is differentiated from the nonruminant by the extensive
development of the stomach and dietary fermentation that occurs there. These tissues are
extremely active metabolically, and their metabolism reduces the net availability of absorbed
amino acids. However, for essential amino acids, this metabolism largely reflects the seques-
tration of arterial amino acids, and not the utilization of amino acids during their absorption,
which represents a relatively small portion of total use by the PDV. The use of essential amino
acids by the PDV largely represents synthesis of constitutive and secreted proteins, and
endogenous losses to the gut, as well as oxidation of branched-chain amino acids. Absorptive
use is more extensive for nonessential amino acids, particularly those used as oxidative fuels.
Fermentation in the rumen results in a substantial recycling of urea to the gut lumen and reab-
sorption as ammonia, which is subsequently converted to urea in the liver. The costs of urea
synthesis from ammonia appear to be relatively small in terms of oxidative metabolism, and
contrary to earlier suggestions it does not appear to require additional amino acid nitrogen. In
contrast, consumption of excess protein can increase heat production in both the PDV and
liver, perhaps as a consequence of surplus amino acid oxidation. Liver removal of amino acids
reflects liver requirements and supply relative to body requirements, but in ruminants a net
release of branched-chain amino acids is often observed, and leucine is oxidized in the PDV
and other peripheral tissues. Finally, amino acids also make important contributions to liver
glucose production in ruminants, but apart from alanine, this appears to reflect the availability
of excess amino acid carbon, and not a metabolic priority, even in very early lactation.

1. INTRODUCTION
The ability of ruminants to derive their nutritional requirements from forages and byproduct
feeds is a consequence of their extensive fore-stomach development and fermentative capacity.
This pregastric microbial symbiosis characterizes ruminant digestion and has unique effects on
their protein and amino acid metabolism compared to nonruminants. Microbial fermentation

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
197 © 2005 Elsevier Limited. All rights reserved.
198 C. K. Reynolds

provides essential amino acids, even in the absence of protein from their diet, allows the uti-
lization of nonprotein nitrogen from blood urea for microbial amino acid synthesis and
absorption by the host, and thus enables extensive recycling of nitrogenous compounds
between the gut lumen, blood, and body tissues (Lapierre and Lobley, 2001). This enables the
ruminant to survive, and reproduce, under conditions where forage quality and thus food protein
are scarce. However, the extensive development, metabolism, and fermentation within the
stomach are not without consequences for the ruminant animal. Microbial protein digestion
and urea catabolism give rise to extensive absorption of ammonia, which must be detoxified
through urea synthesis and amination reactions, largely occurring in the liver. In addition, the
extensive development of the musculature and epithelium of the reticulo-rumen and omasum,
and their functional activity, give rise to a high rate of protein turnover and amino acid
requirement (MacRae et al., 1997a).
Other considerations for the nitrogen economy of the ruminant include the effects of the
fermentative process on the form and pattern in which energy is absorbed. Extensive carbohydrate
fermentation in the rumen provides a more continuous pattern of energy absorption, primarily in
the form of volatile fatty acids, as typically little starch escapes rumen fermentation and is avail-
able for digestion to glucose absorbed from the small intestine. As a consequence, the ruminant
relies primarily on liver gluconeogensis to meet glucose requirements, and glucose is continually
released from the liver, with little apparent diurnal fluctuation in liver glycogen flux (Bergman,
1975). The constant requirement for glucose precursors is met to a large extent by propionate
absorbed from the rumen and hindgut, as well as amino acid carbon. In addition, the flow of pro-
tein to the small intestine, and amino acid absorption, exhibit less variation than in nonruminants.
Most current feeding standards and models of nutrient utilization used for rationing protein
in beef and dairy cattle are based on estimates of dietary protein degradation in the rumen, the
synthesis of microbial protein using available energy and nitrogenous compounds, and the net
flow of feed and microbial and endogenous protein to the abomasum. After reaching the small
intestine, protein digestion and amino acid absorption proceed, much as in nonruminants.
While far from perfect, dynamic, mechanistic models of rumen digestion are available for the
prediction of metabolizable protein flow to the small intestine (Russell et al., 1992; Dijkstra
et al., 2002). However, most of these approaches rely on empirical relationships between esti-
mates of total or individual amino acid flow to the small intestine, assumptions about the basal
flow of endogenous amino acid secretions into the small intestine, and a fixed efficiency of
transfer from the lumen of the small intestine to their appearance in a product (Sniffen et al.,
1992; NRC, 2001). Extensive metabolism of absorbed amino acids occurs in the tissues of the
gut and liver, and the extent to which this metabolism determines the quantity and pattern of
amino acids reaching peripheral tissues, and the effects of diet composition and intake level on
fractional amino acid utilization, are not certain. Therefore, a greater understanding of these
processes, and the contributions of endogenous recycling to amino acid supply, is needed to
refine current empirical models of the efficiency of utilization of amino acids reaching the
small intestine in ruminants, and the development of more enlightened, mechanistic predic-
tions of nutrient utilization for production. Ultimately, these models will enable more precise
rationing of amino acids for productive purposes and thus reductions in the total amounts of
protein fed, and subsequent environmental losses arising from the production of beef and milk.

2. SPLANCHNIC TISSUE METABOLISM


While the word “splanchnic” generally applies to the viscera, the term “total splanchnic” has
historically been used by physiologists as a collective term for the tissues of the portal-drained
Nitrogen metabolism by splanchnic tissues of ruminants 199

viscera (PDV) and liver combined (e.g. Elwyn, 1970). The PDV are those tissues drained by the
hepatic portal vein, and in ruminants include the gastrointestinal tract (reticulo-rumen, omasum,
abomasum, small intestine, cecum, and large intestine), pancreas, spleen, and associated
omental, mesenteric, and other adipose tissue. The PDV and liver are tissues in vascular series,
which are integrated anatomically and functionally through their vascular and neural connec-
tions. This integrated metabolism determines the flow of nutrients from food to other body
tissues, through the gastrointestinal tract’s role in digestion and absorption, and the liver’s
metabolic role as integrator of nutrient supply with requirement. In addition, the splanchnic
tissues are extremely active metabolically, accounting for as much as 50% of total body
oxygen consumption, but a considerably lower proportion of body mass (Reynolds, 1995).
This high rate of metabolism requires a high rate of blood flow, and PDV and liver blood
flow (fig. 1) and oxygen consumption (Reynolds, 2002) in cattle increase with diet dry matter
intake (DMI). The high rate of oxygen consumption by the splanchnic tissues is in part a con-
sequence of a high rate of protein turnover and secretion, and PDV and liver mass, and thus
amino acid requirements, increase with greater DMI and metabolizable energy (ME) supply
(Burrin et al., 1990; Lobley, 1994). In cattle, daily PDV and liver heat production, estimated
from oxygen consumption, increases with greater ME both on a total (fig. 2) or metabolic body
size (body weight 0.75) basis (fig. 3). As much as 38% of cardiac output is distributed to the
liver (Huntington et al., 1990), thus while the splanchnic tissues produce a disproportionate
amount of the body’s oxidative metabolism and heat, they have access to a disproportionate
amount of available nutrients, both during absorption into blood and transfer to the vena cava,
as well as from the arterial blood pool (Reynolds, 2002).

Fig. 1. Portal vein and liver blood flow in growing, lactating, and dry mature cattle. Data are averages of
hourly measurements for a given animal and treatment (n = 335). Data sources described by Reynolds (1995,
2002), Maltby et al. (1993), and Benson et al. (2002).
200 C. K. Reynolds

Fig. 2. Portal-drained visceral (PDV) and liver blood heat production and metabolizable energy (ME) intake
in growing, lactating, and dry mature cattle. Data are daily averages of hourly measurements for a given
animal and treatment (n = 335). Data sources described by Reynolds (1995, 2002), Maltby et al. (1993), and
Benson et al. (2002).

Fig. 3. Portal-drained visceral (PDV) and liver blood heat production and metabolizable energy (ME) intake
in growing, lactating, and dry mature cattle. Data are daily averages of hourly measurements for a given
animal and treatment (n = 335) scaled to metabolic body size (body weight0.75). Data sources described by
Reynolds (1995, 2002), Maltby et al. (1993), and Benson et al. (2002).
Nitrogen metabolism by splanchnic tissues of ruminants 201

2.1. Measurement of net splanchnic metabolism

The use of multicatheterization procedures to measure the quantitative metabolism of nitrogenous


compounds by the splanchnic tissues of ruminants in vivo has been previously described in
detail (Katz and Bergman, 1969; Bergman, 1975; Huntington et al., 1989; Seal and Reynolds,
1993; Reynolds, 1995). Based on initial studies in dogs (Shoemaker, 1964), pioneering work in
sheep by E.N. Bergman of Cornell University elegantly described basic aspects of the inter-
organ flow of amino acids and other metabolites between the ruminant PDV, liver, kidneys, and
hindlimbs (Bergman and Heitmann, 1980; Bergman, 1986). The basic procedure is relatively
simple: permanent indwelling catheters are surgically established in appropriate blood vessels
which enable the measurement of venous–arterial concentration difference (VA) and blood
flow across the tissue of interest, which are multiplied mathematically to obtain a measurement
of net nutrient appearance in venous blood (a positive VA) or removal from arterial blood
(a negative VA). Blood flow is typically obtained by measuring the downstream dilution of a dye
(usually ρ-aminohippurate) infused into a distal mesenteric vein, but can also be measured
electronically using electromagnetic or ultrasound probes. Historically, the usefulness of elec-
tronic probes for measurement of portal vein blood flow was limited by their longevity and
requirement for a round vessel (i.e. an artery). The more recent development of “transit-time”
ultrasound probes (Transonics®, Ithaca, NY) has enabled measurement of total blood flow in
blood vessels with irregular shape using long-term (months) in vivo preparations. However,
until the recent development of a slimmer probe, the bulky size of the probe body and
anatomy of the portal vein have limited the usefulness of transit-time probes for measuring
portal vein blood flow in cattle (Lindsay and Reynolds, 2003).
While simple in concept, the obtaining of valid, statistically sound measurements of dietary
effects or physiological state on splanchnic nutrient flux goes far beyond the maintenance of
catheter patency. Considerations include laminar flow in the portal vein, heterogeneous portal
blood distribution to the liver, the frequency of sampling, postprandial fluctuations, the measure-
ment of small VA coupled with large blood flow rates, etc. The PDV represents a diverse,
heterogeneous, collection of tissue types, and the small intestinal enterocytes where amino acid
absorption occurs represent a small fraction of the total PDV tissues. Therefore, as a consequence
of their anatomical location, the majority of the PDV tissues must to a large extent rely on the
supply of amino acids in arterial blood for their requirements (MacRae et al., 1997a; Reynolds,
2002). Measurements of nutrient flux across sections of the PDV can also be obtained by plac-
ing catheters in vessels draining specific tissue beds, such as the mesenteric-drained viscera
(MDV), but depending on the species of interest and the location of the sampling tip relative to
the ileocecal vein, the measurement may or may not include the cecum and large intestine (Seal
and Reynolds, 1993). As cattle, but not sheep, have a collateral branch of the mesenteric vein
draining the ileum (Habel, 1992), measurements of total small intestinal flux, excluding the con-
tributions of the hindgut, are exceedingly difficult to obtain in cattle without ligating the collateral
branch. Precise placement of the mesenteric sampling catheter between the convergence of the
collateral and jejunal branches and the ileocecal branch is required, and complete mixing of PAH
and venous blood in such a short distance makes measurements of blood flow difficult.

2.2. Isotopic labeling approaches

Measurements of net amino acid flux across the PDV reflect the mathematical summation of
a number of metabolic processes, including absorption from the lumen of the small intestine
into the mesenteric veins, utilization during absorption by the enterocytes, release from
protein turnover, or synthesis, by a variety of PDV tissues, and removal from arterial blood.
202 C. K. Reynolds

Similarly, liver is not a homogeneous tissue, and simultaneously removes individual amino acids
from the portal vein and hepatic artery and releases them into the hepatic veins. To obtain the
true or “gross” rate of metabolite release and removal by a tissue (often called “unidirectional
flux”), isotopic labeling of the arterial blood pool is used to obtain an estimate of gross nutrient
removal from blood (fractional extraction times total arterial supply, or “arterial use”), which
is added to net flux to calculate the gross release into venous blood (Bergman, 1975). This
still underestimates utilization during absorption from the gut lumen, before reaching venous
blood. This “absorptive use” can be measured by infusing labeled metabolite into the small
intestine (usually the duodenum or jejunum) and measuring its quantitative appearance in the
portal (or mesenteric) vein (MacRae et al., 1997a; Yu et al., 2000; Lindsay and Reynolds, 2003).
However, this measurement must be corrected for the extraction of absorbed label from arterial
blood, which is accomplished by simultaneously, or on a separate occasion, labeling the arterial
blood pool (MacRae et al., 1997a). Otherwise, measurements of absorptive use or “first-pass”
metabolism by the small intestine will be overestimated.
Isotopic labeling of the carbon in individual amino acids also allows the measurement of oxi-
dation by specific tissues (Lobley et al., 1995, 2003; Lapierre et al., 1999, 2002), depending on the
amino acid labeled, the specific carbon labeled, and the metabolic route of oxidation. However,
when 13C labeling is used, consideration must be given to the amount of tracer required to label
CO2 relative to tracee turnover, especially under nutritional conditions that reduce oxidation. In
addition, the PDV release of fermentative and salivary CO2, as well as the removal of arterial
blood CO2 (Lobley et al., 2003), complicate the interpretation of CO2 flux across the PDV
(Lindsay and Reynolds, 2003). In addition to measurements of amino acid oxidation, the use of
multiple labeled amino acids, on separate occasions or differentially labeled, allows estimation
of the transfer of carbon among individual amino acids and metabolites within specific tissues
(Wolff and Bergman, 1972a; Bergman, 1975; Lindsay and Reynolds, 2003).

2.3. Interorgan amino acid exchanges in maintenance-fed sheep

Using a combination of multicatheterization and isotopic labeling procedures, as well as multi-


ple gut cannulation, Bergman and his colleagues conducted a detailed series of studies
describing the metabolism of amino acids by the visceral tissues and hind limbs of sheep (for
reviews see Bergman and Heitmann, 1980; Bergman, 1986). The data provide a picture of the
basic patterns of interorgan amino acid exchange in maintenance-fed or 3-day fasted sheep fed
alfalfa pellets, and provided a basis for subsequent studies of the effects of diet composition and
physiological state on amino acid metabolism in sheep and cattle. However, the data were often
obtained from a limited number of animals sampled multiple times to increase replication (see
Wolff et al., 1972), and measurements appear to have been made with limited time for recovery
from surgery. Regardless, these concerns do not diminish the contribution that this research
represents. The data represent a pioneering effort, and have stood the test of time. As emphasized
in their reviews, the work of Bergman and his colleagues highlighted a number of basic concepts,
which will be explored in the sections that follow in light of more recent observations.

3. AMINO ACID UTILIZATION BY THE PORTAL-DRAINED VISCERA


3.1. Absorptive use of amino acids

On a net basis, the disappearance of many amino acids from the lumen of the small intestine
of 2 sheep was considerably greater than their simultaneous net appearance in the portal vein
Nitrogen metabolism by splanchnic tissues of ruminants 203

Table 1
Apparent disappearance of amino acids from the small intestine (SI, g/d) in ruminants and
their net appearance across the portal-drained viscera (PDV) or mesenteric-drained viscera
(MDV) as a fraction of disappearance from the SIa

Sheep (n = 2)b Sheep (n = 3)c Lactating dairy cowsd

Amino acid SI, g/d PDV/SI SI, g/d PDV/SI MDV/SI SI, g/d PDV/SI MDV/SI

Leucine 5.02 0.21 7.77 0.65 1.02 127.7 0.62 0.92


Valine 3.73 0.33 5.47 0.57 0.95 77.9 0.51 1.11
Lysine 3.60 0.66 5.78 0.64 1.09 107.9 0.55 0.76
Threonine 2.82 0.38 4.67 0.73 0.99 65.2 0.43 1.15
Isoleucine 3.38 0.43 5.08 0.58 1.13 81.7 0.62 1.02
Phenylalanine 3.22 0.64 5.06 0.76 1.11 69.6 0.76 1.00
Histidine 1.49 0.11 23.6 0.95 1.27
Methionine 1.44 0.60 36.3 0.67 1.01
Arginine 2.84 0.48 79.6 0.63 1.03
Alanine 3.59 0.85 88 0.80 1.16
Aspartate 5.97 0.02 153.4 0.08 0.03
Asparagine – 0.12 – 0.19 0.37
Cysteine 0.61 0.48 15.1 (1.62) 0.28
Glutamate 6.23 (0.01) 191.1 0.08 0.11
Glutamine – (0.68) – (0.19) 0.20
Glycine 3.61 0.52 88.8 0.42 0.57
Proline 2.10 0.39 64.5 0.09 0.49
Serine 2.10 0.83 58.2 0.75 1.23
a Disappearance calculated as duodenum minus ileal flow, uncorrected for endogenous amino acid flow in the

ileum. Recoveries of asparagine and glutamine calculated relative to aspartate and glutamate disappearance,
respectively. Negative recoveries (net removal from arterial blood) given in parentheses.
b Tagari and Bergman (1978). Averages for measurements in 2 sheep fed high- and medium-protein diets.
c MacRae et al. (1997b). Averages for measurements in 3 sheep fed 800 or 1200 g alfalfa pellets per day.
d Berthiaume et al. (2001). Unlike the data for sheep, measurements of SI, PDV, and MDV were not obtained

simultaneously in the same animals. Data for SI are from 2 cows, whilst data for PDV are apparently from
3 separate cows, 2 of which were also sampled for MDV.

(table 1; Tagari and Bergman, 1978). This was especially true for some of the nonessential
amino acids, such as glutamate and aspartate, which are present in large amounts in duodenal
digesta, as well as some of the essential amino acids. The data suggested that there was a con-
siderable absorptive use of amino acids by the enterocytes of the small intestine, as observed
previously for volatile fatty acid absorption by rumen epithelia (Bergman, 1975). This inter-
pretation was supported by the work of Windmeuller and Spaeth (1980), which showed
that glutamate, glutamine, and aspartate are important energy substrates for the small intestinal
enterocyte of rats. More recent studies in pigs have confirmed that there is an extensive use of
glutamate during its absorption from the small intestine (Stoll et al., 1999). For alanine and serine,
the apparent recovery in the portal vein was much higher, but this in part reflects synthesis by the
PDV from products of glycolysis. Alanine, which is synthesized by the PDV and peripheral
muscle from pyruvate, is often the amino acid released by the PDV in the largest amount, across
a variety of nutritional and physiological states (Wolff et al., 1972; Bergman and Heitmann,
1980). This in part reflects a transfer of N from the catabolism of other amino acids in the
PDV to the liver for ureagenesis, and alanine is typically removed by the liver at rates equal to
net PDV release or greater, on a net basis (Bergman, 1986; Reynolds, 1995; Lobley et al., 2001).
204 C. K. Reynolds

The extensive absorptive use of amino acids suggested by Tagari and Bergman (1978), and
other estimates of “first-pass absorptive use” of amino acids, are in part attributable to
anatomical and technical considerations for the measurements reported. As discussed previ-
ously (Reynolds, 2002), the high rates of glutamate and aspartate disappearance from the
small intestine in part reflects the deamination of glutamine and asparagine during the acid
hydrolysis of digesta prior to amino acid analysis, and resulting overestimate of their flow to
the duodenum. For other amino acids, their low net recovery in the portal vein largely reflects
their arterial use by other PDV tissues, such as the stomach. In sheep (MacRae et al., 1997b)
and cattle (Berthiaume et al., 2001), measurements of net amino acid release by the MDV
were considerably higher (30–50%) than their net release by the PDV (table 1). This is a con-
sequence of the utilization of amino acids from arterial blood by those tissues not drained by
the mesenteric vein. These data strongly suggests that the low recovery of absorbed amino
acids across the PDV, on a net basis, is due to sequestration of amino acids from the arterial
blood pool. In these studies (MacRae et al., 1997b; Berthiaume et al., 2001), comparison of
small intestinal disappearance and net MDV appearance of essential amino acids showed
nearly equal, or lower, rates of disappearance from the small intestine compared to appear-
ance in the mesenteric vein, on a net basis, but these estimates underestimate true rates of
disappearance to the extent that endogenous secretions appear in the ileum. For nonessential
amino acids (Berthiaume et al., 2001), the same comparisons of net intestinal disappearance
and MDV appearance (table 1) compare more favorably with the findings of Tagari and
Bergman (1978), with a low net recovery of aspartate, glutamate, cysteine, proline, and
glycine, and a greater appearance of alanine and serine suggesting the synthesis of the latter
two amino acids by tissues of the MDV.

3.2. Comparison of absorptive and arterial essential amino acid use

Direct measurements of the absorptive use of a mixture of essential amino acids were
obtained by MacRae et al. (1997a) in sheep equipped with multiple intestinal cannulas and
catheters enabling measurement of small intestinal disappearance and PDV flux of amino
acids, along with dual-site isotope infusions to allow simultaneous (consecutive samplings)
measurements of absorptive and arterial use of a mixture of 13C-labeled essential amino acids.
The measurements confirmed that across the total PDV, arterial use of most essential amino
acids (leucine, valine, lysine, threonine, isoleucine, and histidine) accounts for the majority
(75–87%) of total PDV use or “sequestration”. For phenylalanine, a greater proportion (51%)
of total PDV utilization was accounted for by absorptive use. Using a similar approach to
study both PDV and MDV utilization of dual-labeled leucine in sheep (Yu et al., 2000), the
MDV accounted for only a small proportion (12%) of total arterial use of leucine by the PDV.
Of the leucine sequestered during absorption, oxidation accounted for only 2%, thus most
leucine sequestered during absorption is used for synthesis of constitutive and secreted
proteins. Using a similar approach, measurements of arterial and absorptive use of leucine
and phenylalanine were obtained in late-lactation dairy cows fed the same ration at two levels
of DMI (16 or 20 kg/day) and two different levels of DMI in the subsequent dry period (8 or
12 kg/day; Reynolds et al., 2001a). These data confirmed that arterial use of these amino acids
accounted for the majority of total PDV use, but absorptive use accounted for a greater
proportion of total PDV phenylalanine use than measured for leucine. Absorptive use was
negligible at lower DMI within each physiological state, and increased with greater intake,
perhaps reflecting an increase in gut mass relative to body requirements. A low absorptive use
of phenylalanine was also evident in other studies in lactating dairy cows at restricted intakes
Nitrogen metabolism by splanchnic tissues of ruminants 205

(Reynolds et al., 2000), where the increase in PDV release of phenylalanine at the end of a
6-day abomasal infusion of a mixture of essential amino acids equivalent to 800 g of milk
protein was 77% of the phenylalanine infused on a net basis, and 99% after correction for
arterial use.
Taken together, these data confirm that there is substantial utilization of amino acids by the
tissues of the PDV. For many nonessential amino acids, with some notable exceptions, their
utilization during absorption is substantial, and they make important contributions to the energy
and synthetic process of the enterocytes of the small intestine. Those contributions, in the
nonruminant small intestine, are discussed in more detail in other chapters. For most of the
essential amino acids, absorptive use accounts for 25%, or less, of total PDV sequestration.
Therefore, metabolism by the PDV tissues largely reflects use of amino acids from the arterial
pool, and absorptive use has less effect on the quantity and profile of amino acids reaching the
portal vein during the “first pass” of absorption than suggested by simple net flux measurements.
For phenylalanine, absorptive use accounted for a greater proportion of total PDV sequestration,
but this reflected less quantitative arterial use relative to the other essential amino acids, and
not more absorptive use per se (MacRae et al., 1997a; Reynolds et al., 2001a).

3.3. Arterial use of essential amino acids by the PDV

In sheep, total PDV sequestration of essential amino acids varied from 32% (histidine) to 67%
(valine) of whole-body irreversible loss (IRL), indicating a substantial contribution of PDV
tissues to whole-body protein synthesis (MacRae et al., 1997a). In dairy cows, total PDV
sequestration of leucine (27–40%) and phenylalanine (22–37%) represented a smaller pro-
portion of body IRL (table 2), but the proportion was increased by greater DMI in both dry
and lactating cows, yet was not affected by stage of lactation and associated differences in
average DMI (Reynolds et al., 2001a). This suggests that PDV sequestration of these amino
acids, as a fraction of total body IRL, is affected by their supply from the diet relative to
requirement. Although these measurements were not corrected for oxidation, or release of
4-methyl-2-oxopentanoate (MOP), it is likely that a large portion of the leucine sequestered
by the PDV is used for protein synthesis. Numerous studies have shown that the PDV of
sheep and cattle accounts for a major portion of body protein synthesis (24–35%), depending
on the technology used and interpretation of the results obtained (Lobley, 1994; Lobley et al.,
1995, 1996; Lapierre et al., 1999, 2002). Therefore, the tissues of the PDV have a substantial
requirement for essential amino acids, which must to a large extent be met by extraction from
the arterial supply.
The relationship between PDV protein turnover and essential amino acid use is illustrated
by the close agreement between the proportions of essential amino acids sequestered by the
PDV and their relative concentration in constitutive proteins (MacRae et al., 1997a). In addi-
tion, branched-chain amino acids are to a large extent oxidized by extrahepatic tissues, but it
is apparent that leucine, and perhaps other branched-chain amino acids, are oxidized by the
PDV (Lapierre et al., 1999, 2002; Lobley et al., 2003). This oxidation appears to be sensitive
to leucine supply relative to requirement (van der Schoor et al., 2001). In dairy cows, PDV
oxidation of leucine tended to increase when a higher protein diet was fed (Lapierre et al.,
2002), but the effect of protein status on PDV oxidation of other amino acids has not been
determined in ruminants, to my knowledge, and deserves further research. For phenylalanine,
little oxidation occurs in the PDV (Reynolds et al., 2000; Lobley et al., 2003), but liver clear-
ance and hydroxylation is extensive relative to other essential amino acids (Reynolds et al.,
2000; Lobley et al, 2001). This in part reflects the conversion of phenylalanine to tyrosine, as
206 C. K. Reynolds

Table 2
Effects of intake level during lactation and the subsequent dry period on leucine and
phenylalanine kinetics and endogenous leucine flow in the duodenum and ileum of 3 dairy
cows (Reynolds et al., 2001b)

Dry Lactating P<

Low High Low High SEM Stage DMI Inter a

DMI, kg/d 8.0 12.0 14.9 19.4 0.4 – – –


Milk yield, kg/d – – 14.3 16.5 2.9 – 0.200 –
Leucine metabolism, mmol/h
Body irreversible loss rate 51.5 63.2 73.0 94.4 5.0 0.905 0.015 0.393
Total PDV sequestrationb 14.8 29.3 25.3 40.3 3.2 0.153 0.021 0.884
Duodenal flow 35.4 46.6 82.0 106.9 3.3 0.030 0.010 0.204
Endogenous duodenal flow 3.2 3.4 6.4 8.9 0.5 0.040 0.025 0.042
Ileal flow 9.8 15.2 21.1 26.2 1.1 0.307 0.003 0.611
Endogenous ileal flow 1.5 2.2 3.0 3.6 0.5 0.693 0.129 0.848
Phenylalanine metabolism, mmol/h
Body irreversible loss rate 20.4 25.4 28.5 33.6 1.5 0.898 0.013 0.832
Total PDV sequestrationb 6.4 10.4 6.3 17.7 1.7 0.070 0.017 0.163
a Interaction of DMI and stage of lactation.
b Sum of arterial and absorptive use.

well as its use for protein synthesis, but largely represents oxidative metabolism. In lactating
dairy cows in mid-lactation, gross liver removal was equal to 70–80% of total body IRL
(Reynolds et al., 2000). Of the phenylalanine removed, 33–57% was oxidized, while only
4–8% was converted to tyrosine which was subsequently released into blood.

3.4. Endogenous amino acid secretions

In addition to constitutive protein synthesis and oxidation, essential amino acid sequestration
by the PDV must also support the synthesis of protein appearing in the lumen of the gut as
endogenous secretions (or sloughage of constitutive proteins). Depending on their site of
entry into the gut lumen, these endogenous amino acids are reabsorbed or lost in the faeces.
Tagari and Bergman (1978) did not account for endogenous amino acid flow in the ileum in
their study, and thus underestimated total amino acid disappearance. Measurements of
endogenous amino acid flow within the lumen of the gut are difficult to obtain and are often
assumed to be a fixed, basal amount from the regression of amino acid flow on intake
(MacRae et al., 1997a). Other approaches for estimating endogenous protein or amino acid
flow in the gut are based on statistical interpretation of digesta amino acid composition
(Larsen et al., 2000), or the isotopic labeling of precursor pools and gut contents, often using
15N- or 13C-labeled leucine (see Ouellet et al., 2002).

Using long-term intravenous 15N-labeled leucine infusion to estimate total endogenous N flow
into specific components of duodenal and fecal nitrogen flow, and mathematical modeling of
N exchanges, Ouellet et al. (2002) recently estimated that total endogenous N in the duodenum,
arising from urea and amino acid N transfer to bacterial N, or appearing directly as free endoge-
nous N, accounted for 24% of total N flow into the duodenum, and 4% of total N loss in the
feces. Using 13C-labeled leucine infused into the jugular vein for 7 days, endogenous leucine
represented 8% of duodenal leucine flow to the duodenum (table 2; Reynolds et al., 2001a),
Nitrogen metabolism by splanchnic tissues of ruminants 207

a number that is in agreement with measurements based on 14C-labeled leucine in sheep


(Marsden et al., 1988). Endogenous leucine flow in the ileum accounted for 14.5% of total
leucine flow, and the fraction of duodenal and ileal leucine flow of endogenous origin was
unaffected by stage of lactation or DMI, thus endogenous leucine flow increased linearly with
increased DMI (table 2). Larsen et al. (2000) also estimated an increase in endogenous amino
acid flow into the duodenum with increased DMI. These data show that, rather than comprising
a basal amount equal to flow in fasted animals, endogenous amino acid flow increases with
intake and digesta flow. While the values obtained through the use of 13C- or 14C-leucine label-
ing are considerably lower than estimates of the total endogenous contribution to N flow in the
duodenum of dairy cows (Ouellet et al., 2002), they are similar to the estimate of the contri-
bution of free endogenous sources (Ouellet et al., 2002), and thus the differences surely reflect
technical differences between the two methods for estimating endogenous amino acid flow.
The appearance of endogenous amino acids in the rumen and small intestine represents a
route of N recycling, which has been estimated to account for 30–40% of N absorption as
amino acids (Lapierre and Lobley, 2001). Based on their models, Ouellet et al. (2002) esti-
mated that endogenous secretions account for as much as 30% of total PDV sequestration of
amino acids. In our study (table 2), endogenous leucine flow in the duodenum and ileum
accounted for 19–37% of total leucine sequestration by the PDV of dry and lactating dairy
cows. While the reabsorption of endogenous N represents an opportunity for reuse of N, it
also incurs a cost in terms of the rate of protein synthesis in gut tissues (Reeds et al., 1999;
van der Schoor et al., 2002). In ruminants, the efficiency of utilization of absorbed essential
amino acids for growth (largely muscle accretion) is typically lower (50–55%) than in non-
ruminants (70–75%), and this may well be a consequence of the disproportionately high rates
of body protein turnover occurring in the gut compared to nonruminants (MacRae et al.,
1997a). These high rates of protein turnover and endogenous secretion appear to be one cost of
supporting the extensive development of metabolically active rumen tissues and the consumption
of high-fiber diets that increase gut fill.

4. AMMONIA ABSORPTION AND UREA SYNTHESIS


Initial studies of net amino acid absorption across the PDV in sheep fed alfalfa pellets (Wolff
et al., 1972) also highlighted the extensive absorption of ammonia N compared to amino acid N.
Across the PDV, net ammonia absorption accounted for a greater fraction of dietary N intake
(48%) than did the net absorption of N as amino acids (26%). This in part reflects the diet fed,
as the alfalfa pellets fed were relatively high in protein (20%), which is highly rumen-degradable.
However, across a range of dietary treatments, net PDV absorption of ammonia N is highly
correlated with N or digestible N intake, and typically exceeds the absorption of total amino
acid N (Seal and Reynolds, 1993; Reynolds, 1995; Lapierre and Lobley, 2001; Lindsay and
Reynolds, 2003).
In a recent summary of published data from cattle (Lindsay and Reynolds, 2003), increased
net PDV release of ammonia N accounted for 41% of incremental N intake, whilst simulta-
neous increases in α-amino N or total amino acid absorption accounted for 31% of the
increments in N intake. The ammonia absorbed is derived from microbial degradation of
nitrogenous compounds, including feed protein and endogenous urea and proteins, as well as
any ammonia arising from metabolism within PDV tissues. As emphasized by Lapierre and
Lobley (2001), a substantial portion of ammonia absorbed is derived from blood urea. They
estimated that, on average, roughly two-thirds of urea synthesized is transferred to the lumen
of the gut via saliva and direct blood transfer, which may involve a specific transporter
208 C. K. Reynolds

(Waterlow, 1999). Of the urea N transferred to the gut, 40% was recycled as absorbed ammonia,
whilst 50% was absorbed as amino acids synthesized in the rumen. The remainder was lost
in the feces, presumably in the form of microbial protein synthesized in the hindgut. As
microbial protein synthesis is a major route of urea and ammonia N utilization in the gut
lumen, dietary N supply and digestion is not the only determinant of ammonia absorption,
and the supply of fermentable energy and other factors affecting microbial growth also
impacts net ammonia appearance in the portal vein. For example, starch infusion into either
the rumen or abomasum of lactating dairy cows decreased net PDV release of ammonia
(Reynolds et al., 1998), and abomasal starch infusion increased fecal excretion of N, presumably
as a consequence of incomplete starch digestion in the small intestine and increased micro-
bial protein synthesis in the hindgut (Reynolds et al., 2001b).

4.1. Costs of excess protein digestion and urea synthesis in the liver

Ammonia absorbed into the portal vein is efficiently cleared by the liver and detoxified by
incorporation into urea and to a lesser extent other nitrogenous compounds such as glutamine
(Lobley et al., 2000). When considered in isolation, the urea cycle requires four phosphate
bonds per mole of urea synthesized, and on this basis ureagenesis in the liver appears to
account for a substantial portion of liver oxygen use (Reynolds et al., 1991b; Lobley et al.,
1995; Reynolds, 1995; Milano et al., 2000). This accounting is supported by calorimetry studies
in which increased heat production was measured in sheep infused into the abomasum with
ammonia or urea, or intravenously with urea (Martin and Blaxter, 1965). These studies sug-
gested a greater systemic cost of urea synthesis than the theoretical cost of converting
ammonia to urea, which was attributed to increased cycling of urea and ammonia between
the blood and gut lumen (Martin and Blaxter, 1965). In addition, calorimetry studies with
dairy cattle suggested an energetic cost of consuming digestible protein in excess of require-
ments equivalent to 30 kJ ME per g N (Tyrrell et al., 1970). However, these increases in heat
production (i.e. oxygen consumption) were measured on a whole-body basis.
In multicatheterized cattle fed isonitrogenous diets differing in forage:concentrate ratio at
equal ME intakes, digested N, ammonia absorption and liver removal, and liver urea synthesis
were greater when the high-forage diet was fed, but liver oxygen consumption was not affected
significantly (table 3; Reynolds et al., 1991a,b). Liver urea release and oxygen consumption
increased with intake of both rations, but this was associated with an increase in ME, glucose
synthesis, and presumably liver mass. In other work, adding urea to a high-protein diet fed at
maintenance to beef cattle caused a large increase in ammonia absorption and liver urea
release, but again had no significant effect on liver oxygen consumption (table 3; Maltby et al.,
1993). Subsequent studies in sheep in which ammonia delivery to the liver and ureagenesis
were increased by ammonia infusion (table 3) have also failed to show a statistically signifi-
cant increase in liver oxygen consumption (Lobley et al., 1995, 1996), although numerical
trends for increased oxygen uptake were observed in one study when ammonia absorption
was increased by 165% (Milano et al., 2000). These observations suggest that increases in
ammonia absorption and subsequently liver urea synthesis do not require increased oxidative
metabolism in the liver. This may reflect shifts in other metabolic processes within the liver,
to provide the ATP required for ureagenesis without increasing oxidative metabolism in total.
An alternative explanation is that the net cost of urea synthesis is lower when the ATP gain
of fumarate metabolism is included in the balance sheet, which reduces the energy cost of urea
formation by 75% (Reynolds et al., 1991b). This accounting does not explain the observed
increase in body oxygen consumption reported in sheep infused with ammonium bicarbonate
Nitrogen metabolism by splanchnic tissues of ruminants 209

(Martin and Blaxter, 1965), but differences in the relative amount of ammonia infused may
be a factor (Milano et al., 2000).
In contrast to the studies cited in the preceding discussion, increases in dietary protein
level, via the addition of soybean meal to a corn-based diet fed to growing beef steers, sig-
nificantly increased body heat production (Reynolds et al., 1992a), in line with the results of
Tyrrell et al. (1970). In this case, increases in body heat production with higher dietary pro-
tein intake were associated with increases in ammonia absorption (table 3), as well as
increased absorption of some amino acids (Reynolds et al., 1995a), and the increases in body
oxygen consumption were attributable to relatively small, but significant, increases in both
liver and PDV oxygen consumption. The significant increase in liver oxygen use in this later
study may in part reflect the larger number of animals sampled, and low variation observed.
However, in sheep fed low-protein concentrate-based diets supplemented with urea or two
sources of protein, liver oxygen consumption was increased when the supplemental protein

Table 3
Effects of diet intake and composition on net portal-drained visceral ammonia absorption
(PDV NH3N) and net liver release of urea N and use of oxygen in beef cattle

kg/d mmol/h

Animal and diet DMI N intake PDV NH3N Liver urea N Liver O2 use

Growing heifersa
Low intake
75% alfalfa 4.75 0.133 186 354 719
75% corn:SBMb 3.60 0.098 143 235 621
High intake
75% alfalfa 7.78 0.209 340 593 1205
75% corn:SBM 6.51 0.174 250 491 1192
Diet effect, P < 0.001 0.001 0.067 0.080 0.460
Mature steersc
75% alfalfa 5.67 0.153 253 363 926
75% alfalfa + urea 5.80 0.209 398 537 947
Diet effect, P < 0.001 0.001 0.001 0.004 0.457
Growing steersd
Medium intake
75% corn 5.10 0.096 121 199 821
75% corn:SBM 4.98 0.131 212 323 865
High intake
75% corn 6.87 0.128 165 289 1091
75% corn:SBM 6.96 0.179 261 473 1180
Diet effect, P < 0.718 0.001 0.001 0.001 0.051
Mature sheep–mesenteric vein NH3N infusione
Alfalfa pellets 0.71 0.018 27.1 37.7 92.4
+ 12.6 mmol/h NH4Cl 0.71 0.018 39.5 61.9 110.4
NH3 effect, P < 0.001 0.071 0.535
Mature sheep–mesenteric vein NH3N infusionf
Grass pellets 1.09 0.035 37.0 69.0 151.2
+ 7.5 mmol/h N4CO3 1.09 0.035 47.8 84.9 194.4
Grass pellets–barley 1.08 0.031 42.6 71.2 173.4
+ 7.5 mmol/h N4CO3 1.08 0.031 57.7 85.8 188.4
NH3 effect, P < 0.001 0.001 0.252

Continued
210 C. K. Reynolds

Table 3—Cont’d
Effects of diet intake and composition on net portal-drained visceral ammonia absorption
(PDV NH3N) and net liver release of urea N and use of oxygen in beef cattle

kg/d mmol/h

Animal and diet DMI N intake PDV NH3N Liver urea N Liver O2 use

Mature sheep–mesenteric vein NH3N infusiong


Grass pellets 0.78 0.017 20.6 42.6 96.6
+ 9 mmol/h NH4CO3 0.78 0.017 35.3 54.7 96.0
+ 24 mmol/h NH4CO3 0.78 0.017 54.5 81.4 120.2
NH3 effect, P < 0.001 0.002 0.130
Sheep–95% concentrateh
Control 1.10 0.012 6.5 37.4 178
+ Urea 1.13 0.021 12.8 52.3 185
+ SBM 1.19 0.021 11.7 55.2 229
+ BFM 1.01 0.018 11.8 68.3 239
Protein effect,i P < 0.57 0.43 0.40 0.18 0.01
a Reynolds et al. (1991a,b).
b Soybean meal.
c Maltby et al. (1993).
d Reynolds et al. (1992a).
e Lobley et al. (1995). Sheep received 5-day mesenteric vein infusions of low (1.5 mmol/h) or high

(14.1 mmol/h) NH4Cl.


f Lobley et al. (1996). Sheep received 4-day mesenteric vein infusions of low (1.5 mmol/h NH Cl)
4
or high (1.5 mmol/h NH4Cl plus 7.5 mmol/h NH4CO3) ammonium salt.
g Milano et al. (2000). Sheep received 4-day mesenteric vein infusions of 0, 9, or 24 mmol/h NH CO .
4 3
h Ferrell et al. (2001). Sheep were fed a low-protein control concentrate (6.6% crude protein) supplemented

(to 11.2% crude protein) with urea, SBM, or blood and feather meal (BFM).
i Comparison of urea with SBM and BFM protein supplements.

was fed, but not urea (Ferrell et al., 2001; table 3). This was also associated with a trend for
greater net PDV release of α-amino N when the protein-based supplements were fed.
Therefore, increases in oxygen consumption observed in animals fed protein in excess of
requirement may reflect an increased oxidation of amino acids, rather than an energy cost of
urea synthesis from ammonia per se. As emphasized by Bergman and Heitmann (1980),
interorgan cycles involving a number of amino acids shuttle N and carbon between the liver
and peripheral tissues, and provide spatial separation of the urea cycle. With the exception of
branched-chain amino acids, oxidation of essential amino acids occurs to a large extent in the
liver, but oxidation also occurs in peripheral tissues such as muscle, which is the primary site
of branched-chain amino acid catabolism (Layman, 2003). It is apparent that PDV tissues par-
ticipate in the oxidative metabolism of leucine and other amino acids (Lapierre et al., 1999,
2002; Lobley et al., 2001, 2003).
Urea synthesis and excretion essentially represents the catabolism of amino acids available
in excess of their requirement for anabolic and metabolic functions, and the fraction of
ammonia N absorbed that is not recycled to the gut. In the hepatocyte, one N in urea arises
directly from ammonia, whilst the other is derived from aspartate, which is derived from glu-
tamate and thus indirectly from ammonia. Previous studies in multicatheterized cattle
observed an increase in liver amino acid removal under conditions of increased ammonia
absorption, suggesting that the synthesis of aspartate from ammonia via glutamate might be
limiting urea synthesis, thus increasing liver catabolism of amino acids to support ureagene-
sis (Reynolds, 1992; Parker et al., 1995). However, variables other than ammonia supply to
Nitrogen metabolism by splanchnic tissues of ruminants 211

the liver were also altered in the studies forming the basis of this hypothesis (e.g. Reynolds
et al., 1991b). In the intervening period a detailed, systematic series of studies has been
conducted by G.E. Lobley and his colleagues at the Rowett Research Institute to explore
the effects of increased ammonia supply on liver metabolism in sheep, under experimental con-
ditions that sought to control variation in other metabolic variables affecting liver metabolism,
and incorporating novel isotopic labelling experimentation and mathematical modeling (Lobley
et al., 1995, 1996; Milano et al., 2000; Milano and Lobley, 2001). Initial indications suggested
an effect of mesenteric vein NH4Cl infusion on leucine oxidation (Lobley et al., 1995).
However, more recent studies in which shifts in acid–base status were avoided by infusing
NH4HCO3 (Lobley and Milano, 1997; Milano et al., 2000; Milano and Lobley, 2001) found
no indication that increased amino acid deamination is required for the synthesis of urea under
conditions of increased ammonia load, at least in maintenance-fed sheep.

4.2. Regulation of urea synthesis

Since its description seventy years ago, the regulation of the ornithine cycle and liver urea
synthesis has been the subject of numerous reviews and much discussion (e.g. Waterlow,
1999; Lobley et al., 2000), and will not be considered in detail in the present chapter.
Generally, liver removal of amino acids is determined by the functional requirements of the
liver (e.g. constitutive and export protein synthesis) and the availability of amino acids in the
blood pool relative to body requirements. The extent to which ureagenesis in the liver is
directly regulated by factors other than plasma amino acid concentrations has been the sub-
ject of debate, but after decades of research, current thinking is that “it is very unlikely that
the signal (integrating protein intake and urea cycle activity) is an alteration in the plasma
concentration either of total amino-N or any single amino acid” (Waterlow, 1999). In growing
steers treated with growth hormone-releasing factor, increases in body N retention, at equal
N intake, were accompanied by decreases in liver removal of α-amino and ammonia N and
release of urea, but with no change observed in blood α-amino N concentration (Lapierre
et al., 1992; Reynolds et al., 1992b). Similar responses were observed in growing steers treated
with growth hormone (Bruckenthal et al., 1997). In nonruminants, growth hormone treatment
caused a decrease in the activity of urea cycle enzymes in the liver, but the effects may have
been an indirect consequence of increased protein retention in vivo (McLean and Gurney,
1963; Palekar et al., 1981). Abomasal infusion of glucose decreased liver urea production and
increased N retention in maintenance-fed sheep (Obitsu et al., 2000), presumably through
effects of insulin, but the effects on urea production may also reflect indirect responses to
increased amino acid retention. Similarly, abomasal infusion of starch for 2 weeks markedly
increased tissue N retention in late-lactation, pregnant dairy cows (Reynolds et al., 2001b).

5. LIVER AMINO ACID METABOLISM


In maintenance-fed sheep (Wolff et al., 1972; Bergman and Heitmann, 1980), liver removal
of a number of amino acids accounted for substantial portions of their net PDV absorp-
tion and release, but as emphasized by Bergman (1986), the animals were not in positive
tissue N balance. Regardless, these studies identified a number of interorgan cycles involving
specific groups of amino acids. For the glucose precursors alanine, serine, and glycine, their
net removal by the liver was greater than their net absorption, such that their net total splanchnic
flux was negative. A negative total splanchnic flux for these amino acids represents the
contributions of peripheral tissues to their liver metabolism, and the hind limbs released them
212 C. K. Reynolds

on a net basis, reflecting the catabolism of glucose and other amino acids, such as leucine
(Layman, 2003). The branched-chain amino acids were removed by the liver, but in the lowest
amount relative to net PDV release, such that their net total splanchnic release was more positive
than for other amino acids, and they were removed by the hindlimbs. Glutamate was released
by the liver, whilst glutamine was removed, and the opposite occurred in the hindlimbs, repre-
senting a means of transporting ammonia N from peripheral tissues to the liver for urea
synthesis. A similar interorgan “cycle” was apparent for arginine and ornithine plus citrulline,
which shuttled N from the kidney and hindlimbs to the liver as arginine. These data suggest
that some of this arginine was synthesized via the action of arginine synthetase in peripheral
tissues, using ornithine and citrulline which were released by the liver and subsequently
removed by the kidney and hind limbs (Bergman and Heitmann, 1980). Again, this spatial
separation of the urea cycle provides opportunity for metabolic flexibility within and among
tissues of the body. On average, the liver removed more amino acid N than appeared across
the PDV (Wolff et al., 1972), but this reflects the level of feed intake and N status of the ani-
mals, and the fact that net absorption underestimates true rates of absorption to the extent that
amino acids are utilized by the PDV.
With greater intake, the relationship between liver removal of amino acids and their true or
net absorption across the PDV will vary, depending in part on the energy status of the animal,
which determines protein requirement in growing animals. Similarly, relationships between
PDV and liver flux of amino acids in dairy cattle will be determined by stage of lactation and
relative milk protein production. In attempting to ascertain the fractional clearance of amino
acids during their absorption and passage through the liver, information needed for the devel-
opment of mechanistic models of amino acid utilization, amino acids mixtures (or proteins)
have been infused into the abomasum, to increase absorption, or mesenteric veins, to mimic
increased absorption, over short (hours) or long (days) periods (e.g. Reynolds and Tyrrell,
1991; Reynolds et al., 1995b, 1999, 2000; Bruckenthal et al., 1997; Wray-Cahen et al., 1997;
Lobley et al., 2001). The interpretation of these results must consider a variety of factors, such
as the amounts infused relative to requirements, the time allowed for adaptation to the infused
amino acids, the form in which the amino acids are provided (protein vs free amino acids),
and the balance of amino acids provided (see Wolfe and Miller, 2002). Essential amino acids
are not required in isolation, thus provision of a limiting amino acid often creates a deficiency
of a second limiting amino acid. Therefore, the efficiency of utilization of one amino acid,
and thus liver clearance, is dependent on a balanced supply of other amino acids relative to
body requirements.
The inverse of this concept of a balanced amino acid supply is illustrated by a model for
the creation of a deficiency of a specific amino acid through the feeding of a low-protein diet
to lactating goats, and the infusion of a mixture of essential amino acids into the abomasum
which is balanced relative to milk protein composition but for the absence of a specific amino
acid, such as histidine (Bequette et al., 2000). Under these circumstances, the metabolic adap-
tations of the mammary gland to maximize histidine supply for milk protein synthesis
(increased blood flow and mammary histidine extraction) are nothing short of amazing.
Similarly, the metabolic and production responses of the gut and liver to mesenteric vein infusion
of mixtures of amino acids based on milk protein composition were affected by the presence
or absence of nonessential amino acids (Reynolds et al., 1995b). In other studies, the response
of net PDV amino acid release and liver removal of essential amino acids (Reynolds et al.,
1999), and PDV sequestration of leucine and phenylalanine (Caton et al., 2001), differed
when the same quantity of essential amino acids was provided as a component of casein,
which also provided supplemental nonessential amino acids, or a mixture of free essential
Nitrogen metabolism by splanchnic tissues of ruminants 213

amino acids. Differences in whole-body and splanchnic leucine metabolism were also
observed in rats fed supplemental free amino acids compared to the same mixture of amino
acids provided as casein (Daenzer et al., 2001).
Despite these factors, some general patterns of liver metabolism are apparent across studies
in cattle and sheep (Lobley et al., 1996, 2001; Lobley and Milano, 1997; Wray-Cahen et al.,
1997; Lapierre et al., 1999; Reynolds et al., 1999, 2000, 2001a; Blouin et al., 2002). As a pro-
portion of net PDV absorption, liver removal of alanine, serine, glycine, phenylalanine, and
histidine is typically high. This in part reflects the role of the first three as glucose precursors
(Bergman and Heitmann, 1980), the role of glycine in detoxification processes, and the high
concentration of phenylalanine and histidine in plasma protein synthesized by the liver
(Lobley et al., 2001). In contrast, net liver removal of the branched-chain amino acids is typically
low relative to their rate of absorption, as observed in maintenance-fed sheep (Bergman, 1986).
In cattle, the liver typically releases leucine and often valine and isoleucine (Blouin et al., 2002).
Their net removal was increased when their supply to the liver was increased by abomasal
casein infusion, but not a mixture of free essential amino acids, which caused an increase in
their net liver release (Reynolds et al., 1999).
Similar responses were observed when mixtures of free amino acids, based on the compo-
sition of milk protein, were infused into the mesenteric vein of early-lactation dairy cows fed
a low-protein diet for 3 days (Reynolds et al., 1995b). Infusion of the mixture of essential
amino acids increased net liver release of leucine and tended to increase the release of valine,
while the same amino acid mixture infused along with nonessential amino acids had no effect
on their net flux across the liver. On a net basis, the liver releases leucine in dogs as well, but
this could be attributable in part to liver removal and transamination of MOP (Abumrad et al.,
1982). In dairy cows, net liver removal of MOP is insufficient to account for the leucine
released (Lapierre et al., 2002), suggesting other sources such as peptides or blood-borne
proteins arising in other tissues and degraded in the liver (Elwyn, 1970). i-Valerate is a product
of leucine catabolism in the liver, and the synthesis of branched amino acids from branched-
chain volatile fatty acids in the ruminant liver has been proposed (van der Walt, 1993). In
lactating dairy cows, the absorption and liver removal of i-valerate cows (e.g. Reynolds et al.,
2003) is typically much greater than net leucine release (e.g. Lapierre et al., 2002). Leucine
delivery to peripheral tissues is important not only as an essential amino acid for protein
synthesis, but also as a regulator of protein synthesis and other anabolic processes (Abumrad
et al., 1982), in part through modification of insulin signalling (Layman, 2003).

5.1. Gluconeogenesis

In sheep, Wolff and Bergman (1972b) estimated that amino acid carbon accounts for from
11% to 30% of liver glucose synthesis, based on measured transfers of alanine, glutamate,
aspartate, glycine, and serine 14C to glucose (11% of glucose production), or the maximal
potential net contribution of all the plasma amino acids removed by the liver (30% of glucose
production). There is no question that the carbon from these glucogenic amino acids makes
an important contribution to liver glucose synthesis in ruminants. Net liver removal of other
precursors (propionate, lactate, glycerol, i-butyrate, and n-valerate) is seldom adequate to
account for all of the glucose released by the liver, but the extent to which amino acid supply
limits glucose production is not certain.
Glucose production by the liver is regulated by requirement, thus the provision of addi-
tional precursor that is efficiently removed by the liver, such as propionate or alanine,
typically decreases liver lactate removal, without affecting glucose production in fed animals
214 C. K. Reynolds

(Reynolds, 1995). However, in early-lactation dairy cows, glucose production is often con-
sidered to be limited by precursor supply from the diet, as occurs in fasting, and increases in
pyruvate carboxylase activity have suggested a greater contribution of alanine, and by con-
jecture other amino acids, to glucose production in early lactation (Drackley et al., 2001).
However, in early-lactation cows fed low-protein diets, mesenteric vein infusion of a mixture
of nonessential amino acids for 3 days had no significant effect on liver glucose production,
and significantly reduced milk yield (Reynolds et al., 1995b). In contrast, abomasal infusions
of casein, or an equivalent mixture of essential amino acids, increased liver glucose produc-
tion (Reynolds et al., 1999). This suggests that the response to casein was not due to the
provision of nonessential amino acids, but perhaps an effect of essential amino acid supply
on liver metabolism or glucose requirement. Even in very early lactation, the net balance of
liver glucose release and precursor removal does not support the concept of an obligatory
increase in nonessential amino acid contributions to liver glucose synthesis, apart from an
increased potential contribution of alanine (Reynolds et al., 2003). Between 9 days before
calving and 11 days after calving, the potential contribution of alanine to liver glucose release
doubled (from 2.5% to 5%), but increases in net liver removal of propionate, lactate, alanine,
and glycerol were sufficient to account for all of the increase in liver glucose release. This
suggests that the amino acid contributions to liver glucose synthesis in early lactation arise
from the obligatory deamination of amino acids in the liver, rather than being a metabolic
requirement of early lactation.

6. CONCLUSIONS
The extensive development of the ruminant forestomach distinguishes their N economy from
that of nonruminants. The microbial fermentation that occurs there markedly alters the pro-
file and form of protein and amino acids presented to the small intestine for digestion
and absorption, and provides opportunities for extensive recycling of N between the body and
lumen of the gut. This recycling occurs via exchanges of ammonia and urea with blood and
luminal pools, endogenous gut and secretory N entry to the gut lumen, microbial protein syn-
thesis, and the subsequent digestion and absorption of microbial and endogenous amino
acids. The costs of this exquisite microbial symbiosis include the costs of urea synthesis,
which may be less then hypothesized at the level of the liver, and an extensive utilization of
absorbed amino acids from arterial blood, which masks their net appearance from the lumen
of the gut into the portal vein. Liver metabolism of amino acids includes substantial require-
ments for liver functions and the integration of the supply of nitrogenous compounds from
the diet with body requirements. A more detailed understanding of these processes within the
splanchnic tissues and their response to changes in diet composition and intake, relative to
nutrient requirements for production, is needed to improve current empirical models of the
efficiency of amino acid utilization for production. Such data are forthcoming; however, the
prediction of productive responses to changes in amino acid supply from the small intestine
can not be based simply on estimates of nutrient supply. The genetic and environmental
propensity of the animal for the productive use of their nutrient supply will ultimately deter-
mine their N economy.

7. FUTURE PERSPECTIVES
The ability of current feed-rationing and nutrient requirement systems to predict productive
response of ruminants to changes in the supply of protein to the small intestine is currently
Nitrogen metabolism by splanchnic tissues of ruminants 215

limited by an oversimplification of postabsorptive metabolism. Future approaches will need


to address the quantitative metabolism of individual amino acids, which is highly integrated,
complex, flexible, and adaptive. Ultimately, the efficiency with which nitrogenous com-
pounds are utilized will be determined by the propensity of the animal fed for productive
nutrient use. However, the exact mechanisms by which nutrient absorption and tissue require-
ments are integrated are poorly understood. For example, the production of urea from amino
acids is largely a consequence of the availability of amino acids in excess of requirement, yet
the signal by which this oversupply is communicated to the urea cycle remains a mystery.
Although the urea cycle was the first metabolic cycle delineated, and despite years of research
on its regulation, it remains a potentially fruitful area for future research. Unraveling the costs
of excess protein intake in terms of the integration of amino acid catabolism and urea gene-
sis is particularly relevant. The processes of amino acid catabolism do not occur solely in the
liver for all amino acids, yet the response of amino acid metabolism in the gut and other extra-
hepatic tissues to increased protein intake are still poorly described. The regulation of
branched-chain amino acid metabolism and their role in regulating metabolic processes
deserves particular attention in this regard. In addition, the metabolic source of the branched-
chain amino acids released by the liver of ruminants should be determined. As restrictions on
N losses from animal production facilities increase, so to will the need to more precisely formu-
late diets to meet the requirements for specific amino acids. This will only be achieved through
a greater understanding of the metabolic integration of dietary supply with tissue demand.

REFERENCES
Abumrad, N.N., Wise, K.L., Williams, P.E., Abumrad, N.A., Lacy, W.W., 1982. Disposal of α-ketoiso-
caproate: roles of liver, gut and kidneys. Amer. J. Physiol. 243, E123–E131.
Benson, J.A., Reynolds, C.K., Aikman, P.C., Lupoli, B., Beever, D.E., 2002. Effects of abomasal long
chain fatty acid infusion on splanchnic nutrient metabolism in lactating diary cows. J. Dairy Sci. 85,
1804–1814.
Bequette, B.J., Hanigan, M.D., Calder, A.G., Reynolds, C.K., Lobley, G.E., MacRae, J.C., 2000. Kinetics
of amino acid transport in the mammary gland of lactating goats when histidine supply is limiting for
milk production. J. Dairy Sci. 83, 765–775.
Bergman, E.N., 1975. Production and utilisation of metabolites by the alimentary tract as measured in
portal and hepatic blood. In: McDonald, I.W., Warner, A.C.I. (Eds.), Digestion and Metabolism in the
Ruminant. The University of New England, Armidale, Australia, p. 292.
Bergman, E.N., 1986. Splanchnic and peripheral uptake of amino acids in relation to the gut. Fed. Proc.
45, 2277–2282.
Bergman, E.N., Heitmann, R.N., 1980. Integration of whole-body amino acid metabolism. In: Buttery,
P.J., Lindsay, D.B. (Eds.), Protein Deposition in Animals. Butterworth, London, pp. 69–84.
Berthiaume, R., Dubreuil, P., Stevenson, M., McBride, B.W., Lapierre, H., 2001. Intestinal disappearance
and mesenteric and portal appearance of amino acids in dairy cows fed ruminally protected methion-
ine. J. Dairy Sci. 84, 194–203.
Blouin, J.P., Bernier, J.F., Reynolds, C.K., Lobely, G.E., Dubreuil, P., Lapierre, H., 2002. Effect of diet qual-
ity on splanchnic fluxes of nutrients and hormones in lactating diary cows. J. Dairy Sci. 85, 2618–2630.
Bruckenthal, I., Huntington, G.B., Baer, C.K., Erdman, R.A., 1997. The effect of abomasal infusion of
casein and recombinant somatotropin hormone injection on nitrogen balance and amino acid fluxes
in portal-drained viscera and net hepatic and total splanchnic blood in Holstein steers. J. Anim. Sci.
75, 1119–1129.
Burrin, D.G., Ferrell, C.L., Britton, R.A., Bauer, M., 1990. Level of nutrition and visceral organ size and
metabolic activity in sheep. Brit. J. Nutr. 64, 439–448.
Caton, J.S., Reynolds, C.K., Bequette, B.J., Lupoli, B., Aikman, P.C., Humphries, D.J., 2001. Effects of
abomasal casein or essential amino acid infusions on splanchnic leucine and phenylalanine metabo-
lism in lactating dairy cows. J. Dairy Sci. 84, Suppl. 1, 363.
216 C. K. Reynolds

Daenzer, M., Petzke, K.J., Bequette, B.J., Metges, C.C., 2001. Whole-body and splanchnic amino acid
metabolism differ in rats fed mixed diets containing casein or its corresponding amino acid mixture.
J. Nutr. 131, 1965–1972.
Dijkstra, J., Mills, J.A.N., France, J., 2002. The role of dynamic modelling in understanding the micro-
bial contribution to rumen function. Nutr. Res. Rev. 15, 67–90.
Drackley, J.K., Overton, T.R., Douglas, G.N., 2001. Adaptations of glucose and long-chain fatty acid
metabolism in liver of dairy cows during the periparturient period. J. Dairy Sci. 84, E. Suppl.
E100–E112.
Elwyn, D.H., 1970. The role of the liver in regulation of amino acid and protein metabolism. In:
Munro, H.N. (Ed.), Mammalian Protein Metabolism. Academic Press, New York, p. 523.
Ferrell, C.L., Freetly, H.C., Goetsch, A.L., Kreikemeier, K.K., 2001. The effect of dietary nitrogen and
protein on feed intake, nutrient digestibility, and nitrogen flux across the portal-drained viscera and
liver of sheep consuming high-concentrate diets ad libitum. J. Anim. Sci. 79, 1322–1328.
Habel, R.E., 1992. Guide to the Dissection of Domestic Ruminants. Habel, R.E., Ithaca, NY.
Huntington, G.B., Reynolds, C.K., Stroud, B., 1989. Techniques for measuring blood flow in the splanch-
nic tissues of cattle. J. Dairy Sci. 72, 1583–1595.
Huntington, G.B., Eisemann, J.H., Whitt, J.M., 1990. Portal blood flow in beef steers: comparison of tech-
niques and relation to hepatic blood flow, cardiac output and oxygen uptake. J. Anim. Sci. 68, 1666–1673.
Katz, M.L., Bergman, E.N., 1969. Simultaneous measurements of hepatic and portal venous blood flow
in the sheep and dog. Amer. J. Physiol. 216, 946–952.
Lapierre, H., Lobley, G.E., 2001. Nitrogen recycling in the ruminant: a review. J. Dairy Sci. 84, E. Suppl.
E223–E236.
Lapierre, H., Reynolds, C.K., Elsasser, T.H., Gaudreau, P., Brazeau, P., Tyrrell, H.F., 1992. Effects of
growth hormone-releasing factor and intake on energy metabolism in growing beef steers: net
hormone metabolism by portal-drained viscera and liver. J. Anim. Sci. 70, 742–751.
Lapierre, H., Bernier, J.F., Dubreuil, P., Reynolds, C.K., Farmer, C., Ouellet, D.R., Lobley, G.E., 1999.
The effect of intake level on splanchnic protein metabolism in growing beef steers. Brit. J. Nutr.
81, 457–466.
Lapierre, H., Blouin, J.P., Bernier, J.F., Reynolds, C.K., Dubreuil, P., Lobley, G.E., 2002. Effect of diet qual-
ity on leucine metabolism across splanchnic tissues in lactating dairy cows. J. Dairy Sci. 85, 2631–2641.
Larsen, M., Madsen, T.G., Weisbjerg, M.R., Hvelplund, T., Madsen, J., 2000. Endogenous amino acid
flow in the duodenum of dairy cows. Acta Agric. Scand. Sect. A. Anim. Sci. 50, 161–173.
Layman, D.K., 2003. The role of leucine in weight loss diets and glucose homeostasis. J. Nutr. 133,
261S–267S.
Lindsay, D.B., Reynolds, C.K., 2003. Metabolism of the portal-drained viscera. In: Dyjkstra, J., France, J.,
Forbes, J.M. (Eds.), Quantitative Aspects of Ruminant Digestion and Metabolism. CABI,
Wallingford, UK. Submitted.
Lobley, G.E., 1994. Amino acid and protein metabolism in the whole body and individual tissues of rumi-
nants. In: Asplund, J.M. (Ed.), Principles of Protein Nutrition in Ruminants. CRC Press, London, p. 147.
Lobley, G.E., Bremner, D.M., Brown, D.S., 2001. Response in hepatic removal of amino acids by the
sheep to short-term infusions of varied amounts of an amino acid mixture into the mesenteric vein.
Brit. J. Nutr. 85, 689–698.
Lobley, G.E., Connell, A., Lomax, M.A., Brown, D.S., Milne, E., Calder, A.G., Faringham, D.A.H., 1995.
Hepatic detoxification of ammonia in the ovine liver: possible consequences for amino acid catabo-
lism. Brit. J. Nutr. 75, 217–235.
Lobley, G.E., Milano, G.D., 1997. Regulation of hepatic nitrogen metabolism in ruminants. Proc. Nutr.
Soc. 56, 547–563.
Lobley, G.E., Milano, G.D., van der Walt, J.G., 2000. The liver: integrator of nitrogen metabolism. In:
Cronje, P.B. (Ed.), Ruminant Physiology: Digestion, Metabolism, Growth and Reproduction. CAB
International, Wallingford, UK, pp. 149−168.
Lobley, G.E., Shen, X., Le, G., Bremmer, D.M., Milne, E., Calder, A.G., Anderson, S.E., Dennison, N.,
2003. Oxidation of essential amino acids by the ovine gastro-intestinal tract. Brit. J. Nutr. 89, 617–629.
Lobley, G.E., Weijs, P.J.M., Connell, A., Calder, A.G., Brown, D.S., Milne, E., 1996. The fate of absorbed
and exogenous ammonia as influenced by forage or forage-concentrate diets in growing sheep. Brit.
J. Nutr. 76, 231–248.
MacRae, J.C., Bruce, L.A., Brown, D.S., Calder, A.G., 1997a. Amino acid use by the gastrointestinal tract
of sheep given lucerne forage. Amer. J. Physiol. 273, G1200–G1207.
Nitrogen metabolism by splanchnic tissues of ruminants 217

MacRae, J.C., Bruce, L.A., Brown, D.S., Farningham, D.A.H., Franklin, M., 1997b. Absorption of amino
acids from the intestine and their net flux across the mesenteric- and portal-drained viscera of lambs.
J. Anim. Sci. 75, 3307–3314.
Maltby, S.A., Reynolds, C.K., Lomax, M.A., Beever, D.E., 1993. The effect of increased absorption of
ammonia and arginine on splanchnic metabolism of beef steers. Anim. Prod. 56, 462.
Marsden, M., Bruce, C.I., Bartram, C.G., Buttery, P.J., 1988. Initial studies on leucine metabolism in the
rumen of sheep. Brit. J. Nutr. 60, 161–171.
Martin, A.K., Blaxter, K.L., 1965. The energy cost of urea synthesis in sheep. In: Blaxter, K.L. (Ed.),
Energy Metabolism of Farm Animals. EAAP Publ. No. 11. Academic Press, London, pp. 83–91.
McLean, P., Gurney, M.W., 1963. Effect of adrenalectomy and of growth hormone on enzymes concerned
with urea synthesis in rat liver. Biochem. J. 87, 96–104.
Milano, G.D., Hotston-Moore, A., Lobley, G.E., 2000. Influence of hepatic ammonia removal on ure-
agenesis, amino acid utilization and energy metabolism in the ovine liver. Brit. J. Nutr. 83, 307–315.
Milano, G.D., Lobley, G.E., 2001. Liver nitrogen movements during short-term infusion of high levels of
ammonia into the mesenteric vein of sheep. Brit. J. Nutr. 86, 507–513.
NRC, 2001. Nutrient Requirements of Dairy Cattle, 7th Revised Edition. National Academy Press,
Washington, DC.
Obitsu, T., Bremmer, D., Milne, E., Lobley, G.E., 2000. Effect of abomasal glucose infusion on alanine
metabolism and urea production in sheep. Brit. J. Nutr. 84, 157–163.
Ouellet, D.R., Demers, M., Zuur, G., Lobley, G.E., Seoane, J.R., Nolan, J.V., Lapierre, H., 2002. Effect
of dietary fiber on endogenous nitrogen flows in lactating dairy cows. J. Dairy Sci. 85, 3013–3025.
Palekar, A.G., Collipp, P.J., Maddaiah, V.T., 1981. Growth hormone and rat liver mitochondria: Effects
on urea cycle enzymes. Biochem. Biophys. Res. Commun. 100, 1604–1610.
Parker, D.S., Lomax, M.A., Seal, C.J., Wilton, J.C., 1995. Metabolic implications of ammonia produc-
tion in the ruminant. Proc. Nutr. Soc. 54, 549–563.
Reeds, P.J., Burrin, D.G., Stoll, B., van Goudoever, J.B., 1999. Consequences and regulation of gut
metabolism. Proceedings of the VIIIth International Symposium on Protein Metabolism and
Nutrition. EAAP Publ. No. 96, Wageningen Press, Wageningen, The Netherlands, pp. 127–153.
Reynolds, C.K., 1992. Nitrogen metabolism by ruminant liver. J. Nutr. 122, 850–854.
Reynolds, C.K., 1995. Quantitative aspects of liver metabolism in ruminants. In: engelhardt, W.,
Leonhard-Marek, S., Breves, G., Giesecke, D. (Eds.), Ruminant Physiology: Digestion, Metabolism,
Growth and Reproduction: Proceedings of the 8th International Symposium on Ruminant Physiology.
Ferdinand Enke Verlag, Stuttgart, Germany, p. 351.
Reynolds, C.K., 2002. Economics of visceral tissue metabolism in ruminants: toll keeping or internal rev-
enue service? J. Anim. Sci. 80, E. Suppl., E74–E84.
Reynolds, C.K., Aikman, P.C., Lupoli, B., Humphries, D.J., Beever, D.E., 2003. Splanchnic metabolism
of dairy cows during the transition from late gestation through early lactation. J. Dairy Sci. 86,
1201–1217.
Reynolds, C.K., Bequette, B.J., Caton, J.S., Humphries, D.J., Aikman, P.C., Lupoli, B., Sutton, J.D.,
2001a. Effects of intake and lactation on absorption and metabolism of leucine and phenylalanine by
splanchnic tissues of dairy cows. J. Dairy Sci. 84, Suppl. 1, 362.
Reynolds, C.K., Cammell, S.B., Humphries, D.J., Beever, D.E., Sutton, J.D., Newbold, J.R., 2001b.
Effects of post-rumen starch infusion on milk production and energy metabolism in dairy cows.
J. Dairy Sci. 84, 2250–2259.
Reynolds, C.K., Casper, D.P., Harmon, D.L., Milton, C.T., 1992a. Effect of CP and ME intake on vis-
ceral nutrient metabolism in beef steers. J. Anim. Sci. 70, Suppl. 1, 315.
Reynolds, C.K., Crompton, L.A., Firth, K., Beever, D., Sutton, J., Lomax, M., Wray-Cahen, D., Metcalf, J.,
Chettle, E., Bequette, B., Backwell, C., Lobley G., MacRae, J., 1995b. Splanchnic and milk protein
responses to mesenteric vein infusion of 3 mixtures of amino acids in lactating dairy cows. J. Anim. Sci.
73, Suppl. 1, 274.
Reynolds, C.K., Harmon, D.L., Prior, R.L., Casper, D.P., Milton, C.T., 1995a. Splanchnic metabolism of amino
acids in beef steers fed diets differing in CP content at two ME intakes. J. Anim. Sci. 73, Suppl. 1, 270.
Reynolds, C.K., Humphries, D.J., Cammell, S.B., Benson, J., Sutton, J.D., Beever, D.E., 1998. Effects
of abomasal wheat starch infusion on splanchnic metabolism and energy balance of lactating dairy
cows. In: McCracken, K.J., Unsworth, E.F., Wylie, A.R.G. (Eds.), Energy Metabolism of Farm
Animals: Proceedings of the 14th Symposium on Energy Metabolism, CAB International,
Wallingford, UK, p. 39.
218 C. K. Reynolds

Reynolds, C.K., Lapierre, H., Tyrrell, H.F., Elsasser, T.H., Staples, R.C., Gaudreau, P., Brazeau, P.,
1992b. Effects of growth hormone-releasing factor and intake on energy metabolism in growing beef
steers: net nutrient metabolism by portal-drained viscera and liver. J. Anim. Sci. 70, 752–763.
Reynolds, C.K., Lupoli, B., Aikman, P.C., Benson, J.A., Humphries, D.J., Crompton, L.A., France, J.,
Beever, D.E., MacRae, J.C., 2000. Effects of diet protein level and abomasal amino acid infusions on
splanchnic metabolism in lactating dairy cows. J. Dairy Sci. 83, Suppl. 1, 299.
Reynolds, C.K., Lupoli, B., Aikman, P.C., Humphries, D.J., Crompton, L.A., Sutton, J.D., France, J.,
Beever, D.E., MacRae, J.C., 1999. Effects of abomasal casein or essential amino acid infusions on
splanchnic metabolism in lactating dairy cows. J. Anim. Sci. 77, Suppl. 1, 266.
Reynolds, C.K., Tyrrell, H.F., 1991. Effects of mesenteric vein L-alanine infusion on liver metabolism in
heifers fed on diets differing in forage-to-concentrate ratio. Brit. J. Nutr. 66, 437–450.
Reynolds, C.K., Tyrrell, H.F., Reynolds, P.J., 1991a. Effects of diet forage-to-concentrate ratio and intake
on energy metabolism in growing beef heifers: whole body energy and nitrogen balance and visceral
heat production. J. Nutr. 121, 994–1003.
Reynolds, C.K., Tyrrell, H.F., Reynolds, P.J., 1991b. Effects of diet forage-to-concentrate ratio and intake
on energy metabolism in growing beef heifers: net nutrient metabolism by visceral tissues. J. Nutr.
121, 1004–1015.
Russell, J.B., O’Connor, J.D., Fox, D.G., Van Soest, P.J., Sniffen, C.J., 1992. A net carbohydrate and
protein system for evaluating cattle diets: I. Ruminal fermentation. J. Anim. Sci. 70, 3551–3561.
Seal, C.J., Reynolds, C.K., 1993. Nutritional implications of gastrointestinal and liver metabolism in
ruminants. Nutr. Res. Rev. 6, 185–208.
Shoemaker, W.C., 1964. Methods and techniques for measurement of hepatic physiology and metabolism.
In: Rouiller, C. (Ed.), The Liver: Morphology, Biochemistry, Physiology. Academic Press, New York,
pp. 243–266.
Sniffen, C.J., O’Connor, J.D., Van Soest, P.J., Fox, D.G., Russell, J.B., 1992. A net carbohydrate and protein
system for evaluating cattle diets: II. Carbohydrate and protein availability. J. Anim. Sci. 70, 3562–3577.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1999. Substrate oxidation by the portal
drained viscera of fed piglets. Amer. J. Physiol. 277, E168–E175.
Tagari, H., Bergman, E.N., 1978. Intestinal disappearance and portal blood appearance of amino acids in
sheep. J. Nutr. 108, 790–803.
Tyrrell, H.F., Moe, P.W., Flatt, W.P., 1970. Influence of excess protein intake on energy metabolism of
the dairy cow. In: Schurch, A., Wenks, C. (Eds.), Energy Metabolism of Farm Animals. EAAP Publ.
No. 13. Juris Verlag, Zürich, Switzerland, pp. 69–72.
van der Schoor, S.R.D., van Goudoever, J.B., Stoll, B., Henry, J.F., Rosenberger, J.R., Burrin, D.G.,
Reeds, P.J., 2001. The pattern of intestinal substrate oxidation is altered by protein restriction in pigs.
Gastroenterology 121, 1167–1175.
van der Schoor, S.R.D., Reeds, P.J., Stoll, B., Henry, J.F., Rosenberger, J.R., Burrin, D.G., van Goudoever, J.B.,
2002. The high metabolic cost of a functional gut. Gastroenterology 123, 1931–1940.
van der Walt, J.G., 1993. Nitrogen metabolism of the ruminant liver. Aust. J. Agric. Res. 44, 381–403.
Waterlow, J.C., 1999. The mysteries of nitrogen balance. Nutr. Res. Rev. 12, 25–54.
Windmueller, H.G., Spaeth, A.E., 1980. Respiratory fuels and nitrogen metabolism in vivo in small intes-
tine of rats: quantitative importance of glutamine, glutamate and aspartate. J. Biol. Chem. 255, 107–112.
Wolfe, R.R., Miller, S.L., 2002. Protein metabolism in response to ingestion pattern and composition of
proteins: introduction. J. Nutr. 132, 3207S.
Wolff, J.E., Bergman, E.N., 1972a. Metabolism and interconversions of five plasma amino acids by tissues
of the sheep. Amer. J. Physiol. 223, 447–454.
Wolff, J.E., Bergman, E.N., 1972b. Gluconeogenesis from plasma amino acids in fed sheep. Amer. J. Physiol.
223, 455–460.
Wolff, J.E., Bergman, E.N., Williams, H.H., 1972. Net metabolism of plasma amino acids by liver and
portal-drained viscera of fed sheep. Amer. J. Physiol. 223, 438–446.
Wray-Cahen, D., Metcalf, J.A., Backwell, F.R.C., Bequette, B.J., Brown, D.S., Sutton, J.D., Lobley, G.E.,
1997. Hepatic response to increased exogenous supply of plasma amino acids by infusion into the
mesenteric vein of Holstein-Friesian cows in late gestation. Brit. J. Nutr. 78, 913–930.
Yu, F., Bruce L.A., Calder, A.G., Milne, E., Coop, R.L., Jackson, F., Horgan, G.W., MacRae, J.C., 2000.
Subclinical infection with the nematode Trichostrongylus colubriformis increases gastrointestinal tract
leucine metabolism and reduces availability of leucine for other tissues. J. Anim. Sci. 78, 380–390.
PART III
Lipid metabolism
9 Hepatic fatty acid oxidation
and ketogenesis in young pigs1

J. Odle, P. Lyvers-Peffer, and X. Lin

Department of Animal Science, North Carolina State University,


Raleigh, NC 27695, USA

A primary limitation to efficient pork production is morbidity and mortality during the peri-
natal period. Because pigs are born with low energy reserves, their survival hinges on timely
consumption of milk. In contrast to carbohydrate-based fetal metabolism, the transition to a
milk-based diet necessitates rapid biochemical adaptations to accommodate the oxidation of
fatty acids that comprise more than 60% of milk energy. From research reported to date, the
degree to which neonatal pigs make these adaptations is questionable. In stark contrast to
other mammalian neonates, piglets do not demonstrate elevated ketogenesis despite high
milk-fat intake. Ketone bodies play a pivotal role in the transition from carbohydrate-based
metabolism to fat-based metabolism, providing an important alternative fuel for glucose-
dependent tissues. Impaired adaptation limits the piglets’ ability to oxidize fat which likely
contributes to the etiology of mortality. Therefore, this review considers the developmental
aspects of lipid oxidation in the young pig. The key regulatory enzymes previously elucidated
in rodents are reviewed, with inclusion of the limited knowledge available in pigs. Further
research in this area will hopefully assist in development of strategies (via nutritional and/or
exogenous hormonal manipulation) to enhance development of fatty acid oxidation and
ultimately improve piglet survival and growth.

1. INTRODUCTION
Impaired growth and high mortality of neonatal pigs pose significant challenges to the swine
industry. Postnatal mortality varies among production units, but has been recently estimated
by the Agricultural Statistics Service of the United States Department of Agriculture to average
approximately 12% of live births, and has shown only modest improvement over the past twenty
years (USDA, 2002). In addition, it is estimated that prenatal (embryo and fetal) mortality in

1 Supported in part by a grant from the USDA-NRI, No. 98-35206-6645.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
221 © 2005 Elsevier Limited. All rights reserved.
222 J. Odle et al.

swine may be as high as 25% so that, collectively, these data imply that the number of piglets
weaned per litter currently may be less than 65% of true potential. Associated problems of
slow growth (morbidity) further add to the inefficiency. The combined economic impact of
these losses is enormous. The cost is ultimately carried by the consumer in the price paid for
pork. For this reason, there is great impetus for identifying and studying the stressors respon-
sible for the high postnatal mortality. The etiology is complex, as a number of factors may
contribute, including nutritional deficiency, low immunocompetence and disease resistance,
hypothermia, and crushing by the dam. Many retrospective survey studies have attempted to
determine the relative importance of these stressors, but epidemiological/survey approaches
have contributed little useful information because the cause of death is difficult to determine
precisely, and interactions among the factors complicate interpretation. Consequently, if
progress is to be made, controlled experimentation is needed to better understand the devel-
oping piglet’s nutritional, immunological, and behavioral responses to its environment.
This review addresses a metabolic component of this multifactorial etiology, examining the
biochemical competency of piglets to oxidize fatty acids during early postnatal life. In par-
ticular, the ontogeny and regulation of hepatic fatty acid oxidation is highlighted owing to the
dramatically low level of ketogenesis expressed in neonatal pigs compared to other species.

2. THE NEED FOR RAPID DEVELOPMENT


OF FATTY ACID OXIDATION
Prior to birth, the fetus oxidizes predominantly glucose, lactate, and amino acids (Battaglia
and Meschia, 1978). At parturition, the newborn must elicit the behavioral responses neces-
sary to acquire milk from the dam. This requires effective competition among littermates and
occurs in a thermal environment that may be more than 10ºC below the animal’s critical tem-
perature (Stanier et al., 1984). Rohde Parfet and Gonyou (1988) have shown that >30 min
may lapse before the first milk is consumed and considerably longer time is required before
positive energy balance is regained. Owing to limited body reserves at birth, negative energy
balance quickly becomes life-threatening to the piglet. Survival therefore hinges on the timely
consumption of the dam’s milk which provides 60% of its calories as fat (Ferre et al., 1986).
These events necessitate rapid metabolic adaptations to shift from carbohydrate-based fetal
metabolism to fat-based postnatal metabolism.
Other mammalian species, faced with a similar challenge, demonstrate elevated ketogene-
sis during this transition (Girard et al., 1992). For example, ketogenesis measured in
hepatocytes from newborn rats increases 6-fold between 0 and 16 h of age (Ferre et al., 1983),
and blood ketone body concentrations may exceed 1.5 mM (Foster and Bailey, 1976) in the
suckling rat. Thus, neonatal hyperketonemia plays a significant role in the energy economy
of the neonate (Girard et al., 1992), sparing glucose and providing carbon for lipogenesis in
neural tissue. In contrast, piglets do not display hyperketonemia (Bengtsson et al., 1969;
Pegorier et al., 1981) (i.e. less than 0.2 mM) and this may further compromise their survival.
Available data indicate that piglets apparently digest and absorb milk fat with high efficiency
(digestion coefficients >95% at 2 days of age) and that a major portion of absorbed fat is
deposited in adipose tissues prior to weaning (e.g. pigs are <2% fat at birth and ~15% at
weaning). However, relatively little is known regarding the oxidative fate of lipid in the piglet
and even less is known regarding the regulation of lipid oxidation in this species. Before
reviewing the available literature addressing development of fatty acid oxidation in piglets,
the major regulatory features of fatty acid oxidation (elucidated in other species) are
described below.
Hepatic fatty acid oxidation and ketogenesis in young pigs 223

3. REGULATION OF FATTY ACID OXIDATION AND KETOGENESIS


The major metabolic pathways involved in hepatic fatty acid metabolism are summarized
schematically in fig. 1. In animal cells there are two fatty acid β-oxidation systems, one
located in the mitochondria and the second in the peroxisome. The mitochondrion is consid-
ered the primary site for fatty acid β-oxidation, while the peroxisome is considered an
alternative pathway. Under conditions that enhance peroxisome proliferation, the relative
contribution of the peroxisome to total fatty acid oxidation in the liver may be as high as 30%
in the rat (Kondrup and Lazarow, 1985) and 47% in the neonatal pig (Yu et al., 1997a).

Fig. 1. Pathways of hepatic lipid metabolism with emphasis on oxidative metabolism. Enzymes/pathways
are numbered as follows: (1) long-chain acyl-CoA synthetase, (2) acetyl-CoA carboxylase, (3) various
acyl-CoA transferases, (4) carnitine shuttle consisting of CPT I, translocase, and CPT II, (5) medium-chain
acyl-CoA synthetase, (6) mitochondrial hydroxymethylglutaryl-CoA synthase, (7) acyl-CoA dehydrogenase,
(8) long-chain acyl-CoA synthetase, (9) very long-chain acyl-CoA synthetase, (10) acyl-CoA oxidase,
(11) carnitine octanoyltransferase.
224 J. Odle et al.

The mitochondria and peroxisome are often found in close proximity to lipid droplets and
are believed to work in concert (Latruffe et al., 2001). The β-oxidation reactions of the two
systems are similar beginning with an initial dehydrogenation followed by a hydration, a
second dehydrogenation, and finally thiolytic cleavage to produce acetyl-CoA and an acyl-
CoA shortened by two carbons (see Reddy and Mannaerts, 1994, for review). Although the
reactions are similar, the actual proteins involved differ between the β-oxidation systems.

3.1. Biochemical dogma

Long-chain fatty acids (LCFA) are activated to their CoA thioesters via synthetases (fig. 1,
enzyme 1; EC 6.2.1.3) located in the ER and in the outer membrane of the mitochondria (Aas,
1971). To date, five genetic variants of the long-chain acyl-CoA synthetases have been cloned
from rodent species (Oikawa et al., 1998) and their differential regulation may influence the
metabolic fate of the activated fatty acids (Lewin et al., 2001). While preference is shown for
LCFA substrates, those of medium-chain length (MCFA, i.e. C6–C10) may be activated as
well. The fatty acyl-CoA (FA-CoA) may then be esterified via acyl-CoA transferases (fig. 1,
enzyme 3; EC 2.3.1.15) located in the ER, forming various triglycerides, cholesterol esters,
phospholipids, etc., which may be exported as lipoproteins (VLDL; Coleman et al., 2000).
These transferases have low affinity for MCFA-CoAs such that MCFA are obligate fuels (Bach
and Babayan, 1982). The FA-CoA may alternatively be transported into the mitochondria via
the coordinated activities of three membrane proteins: carnitine acyltransferase I (CAT, fig. 1,
enzyme 4; EC 2.3.1.21) catalyzing the formation of FA-carnitine from FA-CoA outside of the
mitochondrial matrix, translocase catalyzing the exchange/diffusion (antiport) of FA-carnitine
for free carnitine across the inner mitochondrial membrane, and carnitine acyltransferase II
(CAT II), similar to CAT I except residing on the matrix side of the inner membrane.
The CAT I and II activities are catalyzed by a family of acyltransferase enzymes. Proteins have
been identified with optimum activity toward C2, C8, and C16 FA-CoAs in various tissues.
The latter, referred to as carnitine palmitoyltransferase (CPT; discussed below) is likely of great-
est importance in the young pig given the predominantly long-chain fatty acid composition of
sow’s milk. Within the mitochondrial matrix, the FA-CoA are subjected to oxidation at the
β-carbon-yielding acetyl-CoA, which may be further oxidized to CO2 in the TCA cycle.
Alternatively, it may exit as acetylcarnitine (using the CAT system in reverse), may produce
ketone bodies (acetoacetate, β-OH-butyrate, and acetone), or may be hydrolyzed to free acetate
by acetyl-CoA hydrolase. Hydroxymethylglutaryl-CoA (HMG-CoA) synthase (fig. 1, enzyme 6;
EC 4.1.3.5) is presumably the rate-limiting enzyme in the ketogenic pathway (Williamson et al.,
1968). Cytosolic acetyl-CoA (derived from the mitochondria via CAT or citrate lyase or from the
peroxisome) may be carboxylated to malonyl-CoA by acetyl-CoA carboxylase (ACC, fig. 1,
enzyme 2; EC 6.4.1.2) within the lipogenic pathway wherein FA-CoA is synthesized de novo.
Very long-chain and long-chain fatty acids are also activated to their CoA thioesters pre-
ceding catabolism in the peroxisome. The peroxisomal membrane contains two synthetases,
a long-chain fatty acid synthetase positioned on the cytosolic side of the membrane (fig. 1,
enzyme 8), and a very long-chain fatty acid (VLCFA) synthetase positioned toward the per-
oxisomal matrix (fig. 1, enzyme 9). The location of the VLCFA synthetase is responsible for
the substrate specificity between the peroxisome and the mitochondrion. While the mito-
chondria can oxidize long, medium, and short-chain fatty acids, VLCFA are either poorly
oxidized by this organelle or not at all (Lazo et al., 1990).
Following CoA activation, the initial dehydrogenation step in the peroxisome is catalyzed
by multiple acyl-CoA oxidases (fig. 1, enzyme 10); two have been identified in the human
Hepatic fatty acid oxidation and ketogenesis in young pigs 225

(Wanders et al., 2001) and three in the rat (Van Veldhoven et al., 2001). While acyl-CoA
dehydrogenase, the first enzyme in mitochondrial β-oxidation, produces two ATP as electrons
are donated directly to coenzyme Q of the electron transport chain (ETC), the first step in
peroxisomal β-oxidation catalyzed by acyl-CoA oxidase results in the production of H2O2 as
electrons are passed directly to molecular oxygen. As a result, the peroxisome is approxi-
mately half as efficient as the mitochondrion in producing energy from the β-oxidation of
fatty acids. Acyl-CoA oxidase shows very little affinity for medium- and short-chain fatty
acids (see Reddy and Mannaerts, 1994, for review); as a result, fatty acids are only chain-
shortened in the peroxisome. In addition, the peroxisome, lacking TCA cycle enzymes,
cannot metabolize the acetyl-CoA to CO2, nor can it produce ketone bodies because the
enzymes of ketogenesis are also absent. Therefore, the end-products of peroxisomal fatty acid
oxidation include acetyl-CoA and chain-shortened acyl-CoA.
The transport of fatty acids and the subsequent end-products of their oxidation across the
peroxisomal membrane is a subject of much debate. It was originally believed that the per-
oxisomal membrane was highly permeable. However, isolation of the peroxisome results in a
loss of the structural integrity of the membrane (Wanders et al., 2001); therefore, many ear-
lier studies involving peroxisome permeability may have been misleading. Studies involving
S. cerevisiae provide evidence for the involvement of transport proteins (see Hettema and
Tabak, 2000, for review); in addition, peroxisomal half transporters have been identified in
humans, although their definitive role in the transport of fatty acids has not been fully eluci-
dated (Wanders and Tager, 1998).
Although a CAT protein has been identified in the peroxisome, it is not membrane-bound
and is therefore not implicated in the transport of FA across the membrane. The CAT identi-
fied in the peroxisome, or carnitine octanoyltransferase (COT, fig. 1, enzyme 11) as it is often
referred to in literature, has optimum activity toward fatty acids of medium chain length. It is
speculated that the peroxisomal COT catalyzes the conversion of the end-products of perox-
isomal β-oxidation to their carnitine esters. Subsequently, these carnitine esters may exit the
peroxisome and be directed toward the mitochondria where they may undergo complete
oxidation to CO2. Studies have shown that 4,8-dimethylnonanoyl-CoA derived from the
incomplete oxidation of pristanic acid in the peroxisome is indeed translocated to the mito-
chondria for complete oxidation (Verhoeven et al., 1998). Furthermore, the concerted action of
the two β-oxidation systems is further supported by the observation that natural and synthetic
ligands (i.e. ligands of the peroxisome proliferator-activated receptor or PPAR), which
increase peroxisome proliferation and peroxisomal enzymes, also increase mitochondrial
enzymes involved in fatty acid metabolism (i.e. CPT I and HMG-CoA synthase).

3.2. Regulation of carnitine palmitoyltransferase I (CPT I)

Using fed, fasted, and alloxan-diabetic rats, McGarry and Foster (1980) have reported con-
siderable evidence establishing the allosteric control of CPT I by malonyl-CoA. During
physiological states in which lipogenesis is occurring, ACC is activated and the associated
high level of malonyl-CoA serves to inhibit CPT I and thereby prevent the simultaneous and
futile oxidation of fatty acids by preventing their entry into the mitochondria. As such, regu-
lation at CPT I is thought to function in directing FA-CoA between esterification and
oxidative fates. Beyond changes in malonyl-CoA concentration, changes in the sensitivity of
CPT I to malonyl-CoA inhibition have also been reported in various physiological states
including the perinatal period in rabbits (Prip-Buus et al., 1990). Furthermore, low levels of
carnitine in tissues of colostrum-deprived neonates (Borum, 1983) could limit transport and
226 J. Odle et al.

thus oxidation of fatty acids. Milk, however, is high in carnitine (Kerner et al., 1984), and
suckling results in elevated hepatic carnitine postnatally (Robles-Valdes et al., 1976).
Medium-chain fatty acids also are a valuable probe in studying regulation at CPT I because
they can diffuse across the mitochondrial membranes and be activated by an alternative
acyl-CoA synthetase (fig. 1, enzyme 5; EC 6.2.1.2) located in the matrix (Groot et al., 1976).
Thus, medium-chain fatty acids may bypass, in part, regulation via CPT I and be oxidized
independently of carnitine. Medium-chain triglyceride utilization by neonatal pigs has been
studied extensively in our laboratory and has been previously reviewed (Odle, 1997, 1998).
Recent advances in regulation of CPT have accompanied cloning of the genes for CPT I
and II (see McGarry and Brown, 1997, for review). Two isoforms of CPT I exist: the L-form
(liver), which possesses relatively low sensitivity to malonyl-CoA inhibition, and the M-form
(muscle), possessing very high sensitivity to malonyl-CoA. Both forms show high inter-
species homology (>80%). In rats, hepatic concentrations of CPT I mRNA have been shown
to increase markedly (up to 5-fold) within 24 h after birth, while CPT II expression was con-
stitutive (Thumelin et al., 1994; Asins et al., 1995). Studying hepatocytes cultured from
perinatal rats, Chatelain et al. (1996) have shown that mRNA levels for CPT I (L) respond
markedly and rapidly to in vitro supplementation with clofibrate, linoleate, and dibutyryl-
cAMP, presumably through interaction with respective cis-acting elements and transcription
factors (e.g. PPAR-RXR, FFAR, and CREB, respectively).

3.3. Regulation of acetyl-CoA carboxylase (ACC)

Regulation of ACC plays a central role in controlling carbon flux through both anabolic and
catabolic pathways of fatty acid metabolism. As the rate-limiting step in de novo fatty acid
biosynthesis, and because of the allosteric influence of the product (malonyl-CoA) of this
enzyme on CPT I (described above), its regulation takes many forms including rapid
allosteric and phosphorylation/dephosporylation mechanisms as well as longer-term mecha-
nisms at the level of gene expression (see Kim, 1997, for review). Hormonal stimulation
(e.g. glucagon) results in increased intracellular cAMP and causes rapid inactivation of ACC
via phosphorylation at multiple sites. Recent findings have suggested that different isozymes
(designated α and β) of ACC, encoded for by different genes, may differentially regulate ana-
bolic and catabolic carbon flux. Indeed, the recently cloned β form (Ha et al., 1996) has an
additional 150 amino acids at the N-terminus that may direct it to insertion into the mito-
chondrial membrane where it may play a direct role in regulating CPT I. This may be of
particular importance in tissues (e.g. piglet liver) in which fatty acid synthesis is negligible
and yet CPT I is highly sensitive to malonyl-CoA inhibition.

3.4. Regulation of mitochondiral HMG-CoA synthase (mHMGCS)

Since the establishment of the CPT I control theory, a growing body of evidence has
accumulated suggesting other possible intramitochondrial regulatory sites, particularly in
neonates (Pegorier et al., 1983; Escriva et al., 1986; Decaux et al., 1988). Much of the early
evidence was indirect and speculative. However, researchers at Cambridge (Lowe and Tubbs,
1985a,b,c; Quant et al., 1989, 1990, 1991, 1993) reported compelling evidence that control of
mHMGCS activity is likely. Specifically, they have shown that its activity is regulated by a
succinylation–desuccinylation mechanism. In the first step of its normal catalytic cycle,
mHMGCS becomes acetylated at the active site by reaction with acetyl-CoA (its first sub-
strate). They have shown that succinyl-CoA (at physiological concentrations) may also react
Hepatic fatty acid oxidation and ketogenesis in young pigs 227

leading to competitive inhibition of the enzyme (Lowe and Tubbs, 1985c). Furthermore, var-
ious in vivo treatments which stimulate ketogenesis (fasting, glucagon or mannoheptulose
injection, alloxan diabetes, high-fat feeding, etc.) all increased mHMGCS activity by decreas-
ing its degree of succinylation (Quant et al., 1989). The enzyme was 40−50% succinylated
(and inactive) in the livers of normally fed rats and could be rapidly (within minutes) acti-
vated in vitro. This led the authors to speculate that the succinylation-control mechanism
could allow for rapid changes in ketogenic flux rate in vivo.
More recently (Quant et al., 1991), the control of ketogenesis in the neonatal rat has been
shown to be mediated, in part, by changes in the amount and activity of mHMGCS, presumably
at the level of gene expression (Casals et al., 1992; see Hegardt, 1999, for review). Indeed,
Thumelin et al. (1993) showed that hepatic mRNA concentrations for mHMGCS increased by
about 3-fold within 24 h of birth in rats, remained constant throughout suckling, and then
rapidly declined when animals were weaned onto a low-fat diet. Furthermore, mRNA in
cultured fetal hepatocytes increased by 4-fold within 4 h after exposure to glucagon. Additional
research (Ayte et al., 1993) has suggested that expression may be regulated in part by
methylation/demethylation of the 5′ flanking region of the gene and has identified CRE and
C-EBP as potential cis regulatory elements (Brady et al., 1993; Gomez et al., 1993) which could
mediate the glucagon effects. Subsequent research also identified the PPAR-RXR diad as an
important regulator of expression induced by clofibrate and fatty acids (Rodriguez et al., 1994).

3.5. Hormonal support of fatty acid oxidation and ketogenesis

Regardless of the underlying biochemical regulatory mechanism(s), the major hormonal


influence is most likely mediated by the insulin/glucagon ratio (McGarry and Foster, 1977).
When the ratio is low, as observed during the neonatal period, fasting, or diabetes, ketogenesis
is stimulated. The hormonal effect may be mediated through regulation of acetyl-CoA car-
boxylase (Borthwick et al., 1986), thus affecting malonyl-CoA levels and CPT I activity
and/or by decreasing succinyl-CoA levels and thereby activating mHMGCS (Quant et al.,
1989) and/or by affecting levels of TCA cycle intermediates. Similarly, hormonal alteration
of intracellular cAMP concentrations (Pegorier et al., 1989) may directly impact gene tran-
scription (as described previously) via interaction with cAMP response elements.
Furthermore, in vivo and in vitro exposure of rodent tissues to the peroxisome-proliferating
hypolipidemic drugs (e.g. clofibrate) and dehydroepiandrosterone (Brady et al., 1991) has
been shown to upregulate fatty acid oxidation and/or ketogenesis by increasing transcription
of CPT and/or mHMGCS genes.

3.6. Hepatic lipid metabolism in the piglet

Clearly, a fundamental understanding of the developmental aspects of lipid metabolism is


essential in order to optimize postnatal fat utilization by the piglet. Putative changes in meta-
bolic capacity within the first week of life are of greatest concern because 75% of mortality
occurs during this time period (USDA, 1991). Unfortunately, relatively little research has
been focused on these animals. The following review examines pig-specific studies.

3.7. Development of fatty acid oxidation in piglets

Postnatal increases in fatty acid oxidation have been reported (Wolfe et al., 1978), but
must be interpreted in light of the general increase in metabolic rate that occurs after birth
228 J. Odle et al.

(Odle et al., 1991b). For example, the oxidation of [U-14C]palmitate to CO2 and acid-soluble
products (considered to represent ketone bodies and/or TCA cycle intermediates) by liver
homogenates was reported to increase 4-fold between 0 and 7 days of age (Mersmann and
Phinney, 1973). However, marked increases in mitochondrial respiration from several TCA
cycle intermediates has likewise been reported (Mersmann et al., 1972) wherein oxygen con-
sumption increased by 5-fold between 6 and 12 h postpartum. Subsequent histological work
suggested rapid mitochondrial proliferation during this early neonatal period. Odle et al.
(1991b) have likewise observed developmental increases in hepatic fatty acid oxidation in
small and normal-birth-weight pigs during the first 48 h of life which could be largely
explained by increases in oxygen consumption. Thus, increases in fatty acid oxidation post-
natally may be accounted for, in part, by increased oxidative metabolism in general, and may
not necessarily infer an increased reliance upon fat as a fuel (Adams et al., 1997a).

3.8. Lack of ketogenesis in piglets

While piglets appear to display a hormonal profile (low insulin/glucagon) that would support
ketogenesis (Pegorier et al., 1981), and have ample substrate from the fat present in milk, they
do not display hyperketonemia (Bengtsson et al., 1969; Pegorier et al., 1981), despite elevated
plasma nonesterified fatty acids (Adams and Odle, 1993b). This starkly contrasts with other
mammalian species (e.g. rats, rabbits, etc.) which show pronounced hyperketonemia during
suckling (Foster and Bailey, 1976) as well as ruminant species which under extreme lacta-
tional stress can die from ketoacidosis. Because ketone bodies provide important
glucose-sparing carbon, aiding otherwise glucose-dependent tissues (e.g. neural tissues), their
absence may be detrimental to the survival of the piglet, which is keenly susceptible to hypo-
glycemia (Swiatek et al., 1968). Furthermore, insofar as fatty acid oxidation also is required
to support active gluconeogenesis, impaired fat oxidation also could contribute indirectly to
hypoglycemia (Lepine et al., 1993; Duee et al., 1996). In theory, low ketone concentrations
could be due to a low production rate (i.e. ketogenesis) and/or a high rate of utilization.
Using continuous-infusion isotope kinetics, we have observed limited β-OH-butyrate
oxidation rates in vivo (Tetrick et al., 1995). Piglets were arterially infused with 14C-β-OH-
butyrate at rates sufficient to supply 15%, 30%, 45%, and 60% of their estimated ATP
turnover. Michaelis–Menten analysis of measured oxidation rates versus plasma concentra-
tions showed that β-OH-butyrate could supply a maximum of 32% of the piglet’s total body
ATP turnover, but at physiological concentrations would supply <5% of the animal’s energy
need. More likely, an impaired rate of hepatic ketogenesis (Pegorier et al., 1983; Duee et al.,
1994) is the cause. Comparative in vivo research (Adams and Odle, 1993b) showed that the
relative ketogenic capacity (measured by regression of plasma β-OH-butyrate on plasma
octanoate concentrations following intraperitoneal injection of octanoate) of neonatal pigs is
greatly attenuated (by up to 1−2 orders of magnitude) compared to weanling or mature swine
and to neonatal or mature rabbits. Furthermore, using radio-HPLC to characterize products of
radiolabeled fatty acid oxidation, we have observed trivial accumulation of isotope in ketone
bodies from piglet liver compared with neonatal rat liver preparations (Adams et al., 1997a).
Concurrent with these findings, measurements of hepatic mHMGCS have shown 70% lower
activities in neonatal pigs than in neonatal rabbits (Adams and Odle, 1993a) or adult rats
(Duee et al., 1994).
Following the cloning of the pig gene (Adams et al., 1997b), most recent findings con-
firmed the attenuated expression of this enzyme (compared with rats) during suckling, but
showed that starvation led to increased mRNA concentrations. However, substantial increases
Hepatic fatty acid oxidation and ketogenesis in young pigs 229

in mRNA did not occur until 2−3 weeks of age, indicating a developmental lag in induction
compared with the rapid postnatal rise observed in the suckling rat (Thumelin et al., 1993).
Subsequent examination of the upstream regulatory region of the pig gene has failed to iden-
tify any idiosyncrasies that might impair expression in that it contains the anticipated PPAR
response element (Ortiz et al., 1999). Despite increased mRNA concentrations, enzyme activ-
ity still remained low. Current research findings indicate that pig mRNA codes for a
catalytically active protein, and low synthase activity is a result of attenuated translation.
Furthermore, the decrease in mRNA translation was not a result of alteration of the
polyadenylate tail which can influence both mRNA stability and subsequent translation into
protein (Barrero et al., 2001).

3.9. Regulation of fatty acid oxidation and ketogenesis in piglets


The degree to which the regulatory features of ketogenesis (reviewed previously) described
in other species extrapolates to the neonatal pig is not known. A putative limitation of fatty
acid oxidation at the level of CPT I was suggested by data from Honeyfield and Froseth
(1991), who reported a >10-fold increase in oxygen consumption when palmitoylcarnitine
was compared to palmitoyl-CoA in heart and liver homogenates from newborn pigs.
Likewise, using hepatocytes isolated from piglets at birth and from piglets fed or fasted 24 h
from birth, we (Odle et al., 1995) have obtained evidence consistent with the hypothesis that
CPT I is a potential regulatory site: (1) The molar oxidation rate of octanoate was 4 times
higher than that of palmitate. This is 2-fold higher than can be explained by the molar energy
difference between these fatty acids, and implies a putative limitation in the oxidation of
palmitate compared to octanoate, given the measured oxygen consumption rates were similar
in cells incubated with each fatty acid. While differences in the relative activities of the
medium- and long-chain acyl-CoA synthetases cannot be excluded, this observation could be
explained by a limitation of palmitoyl-CoA transport into the mitochondrion via CPT.
(2) Carnitine (a co-substrate for CPT) increased oxidation and decreased esterification of
palmitate, but had no effect on octanoate metabolism. (3) TDGA (an irreversible inhibitor of
CPT I) reduced oxidation and increased esterification of palmitate, but again had no effect on
octanoate metabolism.
Developmental changes in liver and muscle CPT activity during the neonatal period have
been reported (Bieber et al., 1973; Lin and Odle, 1995; Schmidt and Herpin, 1998). The activ-
ity doubled in the liver of pigs between 0 and 1 day of age but then plateaued to equal the
activity in mitochondria from 24-day-old animals. Interestingly, palmitoyl-CoA oxidation
continued to increase, doubling between 1 and 2 days postpartum. This suggests that some-
thing other than CPT activity was limiting oxidation as the animals aged. Perhaps decreases
in the sensitivity of CPT to malonyl-CoA (Duee et al., 1994; Lin and Odle, 1995; Schmidt
and Herpin, 1998) are responsible, as has been reported in rabbits (Prip-Buus et al., 1990).
Indeed, Pegorier et al. (1983) challenged the idea of control mediated via CPT I. In contrast
to our findings (Odle et al., 1991a,b, 1995; Lin et al., 1996), they reported low oxidation rates
for octanoate (vs oleate). In addition, they were unable to increase oleate oxidation by iso-
lated piglet hepatocytes incubated with glucagon (which should have decreased malonyl-CoA
levels). This may not be surprising because the liver is not a major site of lipogenesis in the
pig (Mersmann et al., 1973). In general, malonyl-CoA concentrations are higher in lipogenic
tissues as the synthesis of malonyl-CoA is the first committed step of lipogenesis. The role of
malonyl-CoA in nonlipogenic tissues, such as skeletal muscle, is regulation of CPT I activity.
Thus, in nonlipogenic tissues the concentration of malonyl-CoA is much less than what is
230 J. Odle et al.

observed in lipogenic tissues and CPT I exhibits greater sensitivity to malonyl-CoA. Recent
cloning of the pig M- and L-CPT I isoforms has revealed that L-CPT I is a natural chimera
of rat M- and L-CPT I isoforms (Nicot et al., 2001). Pig L-CPT I shows similar kinetics
toward carnitine as rat L-CPT I; however, IC50 for malonyl-CoA is similar to rat M-CPT I.
Thus, pig L-CPT I has a greater sensitivity to malonyl-CoA than rat L-CPT I.
In addition to the potential control by CPT I, regulation may also shift to an intramito-
chondrial site (e.g. mHMGCS). Research reported by Duee et al. (1994) verified the low rate
of ketogenesis in mitochondria from 2-day-old piglets, and implicated an intramitochondrial
constraint; namely, when the mitochondria were treated with malonate (a blocker of succinate
dehydrogenase) to decrease TCA cycle flux, the major end-product of palmitoylcarnitine
oxidation was presumed to be acetoacetate. In this case, the rate of palmitoylcarnitine utiliza-
tion was only 2.5 nmol/min/mg protein. However, when malonate was added, the end-product
was citrate and the rate of palmitoylcarnitine utilization increased by 3-fold.

3.10. Peroxisomal β-oxidation in piglets

We have characterized the postnatal development and tissue distribution of peroxisomal


fatty acid oxidation in piglets (Yu et al., 1997a,b, 1998). In general, activity measured in liver,
kidney, and heart as either antimycin/rotenone-insensitive palmitate oxidation or palmitoyl-
CoA dependent KCN-insensitive reduction of NAD, was high (i.e. 40−50% of total fat
β-oxidation) compared with published rat values (e.g. 25%) and developed rapidly after birth.
Hepatic activity of peroxisomal oxidation was increased when suckling-aged piglets were
fasted, presumably due to elevated plasma free fatty acids and/or glucagon. Most recent data
from our laboratory (Yu et al., 2001) showed that hepatic activities of fatty acid oxidase (an
enzyme unique to peroxisomes) as well as CPT were dramatically induced in piglets fed milk
replacer containing clofibrate. Clofibrate is a hypolipidemic drug that is known to induce
peroxisome (and possibly mitochondrial) biogenesis in rodents (Brady et al., 1991), through
interaction with the peroxisome proliferator-activated receptor (PPAR). We hypothesize that
the extra thermogenesis associated with peroxisomal β-oxidation may be important in the
suckling piglet’s maintenance of homeothermy.

4. FUTURE PERSPECTIVES
Research findings to date indicate impairment in ketone body synthesis for the neonatal pig
when compared to other species. This reduction in ketone bodies during a period when fatty
acids are the main source of energy implies that the young pig is inefficient in utilizing dietary
fat, and thus may impact piglet survival and performance. In order to optimize nutrition for
the neonatal pig, an understanding of the underlying mechanisms of lipid metabolism must
be elucidated. Further insight into the regulation of key lipid enzymes, including rate of syn-
thesis and degradation of mRNA and protein, and subsequent protein activity is required.
Recent research on CPT I and HMG-CoA synthase has shown key differences between these
enzymes in piglets when compared to the rat, a popular model in lipid research. These dif-
ferences highlight the need for more research in the neonatal pig; only then will the
mechanisms responsible for lipid oxidation be defined so that improvements in lipid metab-
olism can be made at a cellular level.
Research in the arena of molecular mechanisms of lipid oxidation will not provide overnight
insight into improving current practices for piglet survival and performance. However, as
swine production becomes increasingly specialized, the demand for the formulation of milk
Hepatic fatty acid oxidation and ketogenesis in young pigs 231

replacers that optimize early growth will increase also. The use of such milk replacers will
allow dietary manipulation that cannot be achieved through nutritional management of the
dam. Control over the piglet’s diet would allow for manipulation of the relative percent of
energy contributed by carbohydrates versus fat and could have a profound impact on piglet
morbidity and mortality. The specific fatty acids added to the diet also could be altered to
increase the amount of MCT (for example), which serve as obligatory fuels. Furthermore,
current research in the use of peroxisome proliferators indicate that their supplementation into
the diet may provide a useful tool in improving fatty acid utilization by the pig. Further
research in these areas will hopefully assist in development of strategies to enhance fatty acid
oxidation and ultimately improve piglet survival and growth.

REFERENCES
Aas, M., 1971. Organ and subcellular distribution of fatty acid activating enzymes in the rat. Biochim.
Biophys. Acta 23, 32–47.
Adams, S.H., Odle, J., 1993a. Does HMG-CoA synthase play a role in limitation of ketogenesis in neona-
tal pigs? FASEB J. 7, A379.
Adams, S.H., Odle, J., 1993b. Plasma β-hydroxybutyrate after octanoate challenge: attenuated ketogenic
capacity in neonatal swine. Amer. J. Physiol. 265, R761−R765.
Adams, S.H., Lin, X., Yu, X.X., Odle, J., Drackley, J.K., 1997a. Hepatic fatty acid metabolism in pigs
and rats: major differences in end-products, O2 uptake and β-oxidation. Amer. J. Physiol. 272,
R1641−R1646.
Adams, S.H., Alho, S.C., Asins, G., Hegardt, F.G., Marrero, P.F., 1997b. Gene expression of mitochon-
drial 3-hydroxy-3-methylglutaryl-CoA synthase in a poorly ketogenic mammal: effect of starvation
during the neonatal period of the piglet. Biochem. J. 324, 65−73.
Asins, G., Serra, D., Arias, G., Hegardt, F.G., 1995. Developmental changes in carnitine palmitoyltrans-
ferases I and II gene expression in intestine and liver of suckling rats. Biochem. J. 306, 379−384.
Ayte, J., Gomez, G.G., Hegardt, F.G., 1993. Structural characterization of the 3′ noncoding region of the gene
encoding rat mitochondrial 3-hydroxy-3-methylglutaryl coenzyme A synthase. Gene 123, 267−270.
Bach, A.C., Babayan, V.K., 1982. Medium-chain triglycerides: an update. Amer. J. Clin. Nutr. 36,
950−962.
Barrero, M.J., Alho, C.S., Oritz, J.A., Hegardt, F.G., Haro, D., Marrero, P.F., 2001. Low activity of mito-
chondrial HMG-CoA synthase in liver of starved piglets is due to low levels of protein despite high
mRNA levels. Arch. Biochem. Biophys. 385, 364−371.
Battaglia, F.C., Meschia, G., 1978. Principal substrates of fetal metabolism. Physiol. Rev. 58, 499.
Bengtsson, N.J., Gentz, J., Hakkarainen, J., Hellstrom, R., Persson, B., 1969. Plasma levels of FFA, glycerol,
β-hydroxybutyrate and blood glucose during the postnatal development of the pig. J. Nutr. 97, 311−315.
Bieber, L.L., Markwell, M.A.K., Blair, M., Helmrath, T.A., 1973. Studies on the development of carni-
tine palmitoyltransferase and fatty acid oxidation in liver mitochondria of neonatal pigs. Biochim.
Biophys. Acta 326, 145−154.
Borthwick, A.C., Edgell, N.J., Denton, R.M., 1986. Mechanisms involved in the short-term regulation of
acetyl-CoA carboxylase by insulin. Biochem. Soc. Trans. 14, 563−565.
Borum, P.R., 1983. Carnitine. Annu. Rev. Nutr. 3, 233−259.
Brady, L.J., Ramsay, R.R., Brady, P.S., 1991. Regulation of carnitine acyltransferase synthesis in lean and
obese zucker rats by dehydroepiandrosterone and clofibrate. J. Nutr. 121, 525−531.
Brady, P.S., Barke, R.A., Brady, L.J., 1993. Regulation of the 68-kDa hepatic carnitine palmitoy-
ltransferase. In: Berdanier, C.D. (Ed.), Nutrition and Gene Expression. CRC Press, Boca Raton, FL,
pp. 319−334.
Casals, N., Roca, N., Guerrero, M., Gil-Gomez, G., Ayte, J., Ciudad, C.J., Hegardt, F.G., 1992. Regulation
of the expression of the mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase gene: its role in the
control of ketogenesis. Biochem. J. 283, 261−264.
Chatelain, F., Kohl, C., Esser, V., McGarry, J.D., Girard, J., Pegorier, J.P., 1996. Cyclic AMP and fatty
acids increase carnitine palmitoyltransferase I gene transcription in cultured fetal rat hepatocytes.
Eur. J. Biochem.235, 789−798.
232 J. Odle et al.

Coleman, R.A., Lewin, T.M., Muoio, D.M., 2000. Physiological and nutritional regulation of enzymes of
triacylglycerol synthesis. Annu. Rev. Nutr. 20, 77−103.
Decaux, J.F., Robin, D., Robin, P., Ferre, P., Girard, J., 1988. Intramitochondrial factors controlling
hepatic fatty acid oxidation at weaning in the rat. FEBS Lett. 232, 156−158.
Duee, P.-H., Pegorier, J.P., Quant, P.A., Herbin, C., Kohl, C., Girard, J., 1994. Hepatic ketogenesis in
newborn pigs is limited by low mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase activity.
Biochem. J. 298, 207−212.
Duee, P.-H., Pegorier, J.P., Darcy-Vrillon, B., Girard, J., 1996. Glucose and fatty acid metabolism in the
newborn pig. In: Tumbleson, M.E., Schook, L.B. (Eds.), Advances in Swine in Biomedical Research
Plenum Press,New York, pp. 865−884.
Escriva, F., Ferre, P., Robin, D., Decaux, J.F., Girard, J., 1986. Evidence that the development of hepatic
fatty acid oxidation at birth in the rat is concomitant with an increased intramitochondrial CoA con-
centration. Eur. J. Biochem. 156, 603−607.
Ferre, P., Satabin, P., Decaux, J.F., Escriva, F., Girard, J., 1983. Development and regulation of ketogen-
esis in hepatocytes isolated from newborn rats. Biochem. J. 214, 937−942.
Ferre, P., Decaux, J.F., Issad, T., Girard, J., 1986. Changes in energy metabolism during the suckling and
weaning period in the newborn. Reprod. Nutr. Dev. 26, 619−631.
Foster, P.C., Bailey, E., 1976. Changes in hepatic fatty acid degradation and blood lipid and ketone body
content during development in the rat. Enzyme 21, 397−407.
Girard, J., Ferre, P., Pegorier, J.P., Duee, P.H., 1992. Adaptations of glucose and fatty acid metabolism
during the perinatal period and suckling-weaning transition. Physiol. Rev. 72, 507−562.
Gomez, G.G., Ayte, J., Hegardt, F.G., 1993. The rat mitochondrial 3-hydroxy-3-methylglutaryl-
coenzyme-A-synthase gene contains elements that mediate its multihormonal regulation and tissue
specificity. Eur. J. Biochem. 213, 773−779.
Groot, P.H.E., Scholte, H.R., Hulsmann, W.C., 1976. Fatty acid activation: specificity, localization and
function. Adv. Lipid Res. 14, 75−126.
Ha, J., Lee, J.K., Kim, D.S., Witters, L.A., Kim, K.H., 1996. Cloning of human acetyl-CoA carboxylase-β
and its unique features. Proc. Natl. Acad. Sci. USA 93, 11466−11470.
Hegardt, F.G., 1999. Mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase: a control enzyme in
ketogenesis. Biochem. J. 338, 569−582.
Hettema, E.H., Tabak, H.F., 2000. Transport of fatty acids and metabolite across the peroxisomal mem-
brane. Biochim. Biophys. Acta 1486, 18−27.
Honeyfield, D.C., Froseth, J.A., 1991. Evaluation of energy sources with and without carnitine in new-
born pig heart and liver. J. Nutr. 121, 1117−1122.
Kerner, J., Froseth, J.A., Miller, E.R., Bieber, L.L., 1984. A study of the acylcarnitine content of sows’
colostrum, milk and serum and newborn piglet tissues: demonstration of high amounts of isovaleryl-
carnitine in colostrum and milk. J. Nutr. 114, 854−861.
Kim, K.H., 1997. Regulation of mammalian acetyl-coenzyme A carboxylase. Annu. Rev. Nutr. 17, 77−99.
Kondrup, J., Lazarow, P.B., 1985. Flux of palmitate through the peroxisomal and mitochondrial β-oxidation
systems in isolated rat hepatocytes. Biochim. Biophys. Acta 835, 147−153.
Latruffe, N., Cherkaoui Malki, M., Nicolas-Frances, V., Jannin, B., Clemencet, M.-C., Hansmannel, F,
Passilly-Degrace, P., Berlot, J.-P., 2001. Peroxisome-proliferator-activated receptors as physiological sen-
sors of fatty acid metabolism: molecular regulation in peroxisomes. Biochem. Soc. Trans. 29, 305−309.
Lazo, O., Contreras, M., Singh, I., 1990. Topographical localization of peroxisomal acyl-CoA ligases: dif-
ferential localization of palmitoyl-CoA and lignoceroyl-CoA ligases. Biochemistry 29, 3981−3986.
Lepine, A.J., Watford, M., Boyd, R.D., Ross, D.A., Whitehead, D.M., 1993. Relationship between
hepatic fatty acid oxidation and gluconeogenesis in the fasting neonatal pig. Brit. J. Nutr. 70, 81−91.
Lewin, T.M., Kim, J.H., Granger, D.A., Vance, J.E., Coleman, R.A., 2001. Acyl-CoA synthetase isoforms
1, 4, and 5 are present in different subcellular membranes in rat liver and can be inhibited independently.
J. Biol. Chem. 276, 24674−24679.
Lin, X., Odle, J., 1995. Regulation of fatty acid oxidation by hepatic mitochondria from neonatal pigs via
carnitine palmitoyltransferase I. J. Anim. Sci. 73, Suppl. 1, 77.
Lin, X., Adams, S.H., Odle, J., 1996. Acetate represents a major product of heptanoate and octanoate
β-oxidation in hepatocytes isolated from neonatal piglets. Biochem. J. 318, 235−240.
Lowe, D.M., Tubbs, P.K., 1985a. 3-Hydroxy-3-methylglutaryl-coenzyme A synthase from ox liver:
purification, molecular and catalytic properties. Biochem. J. 227, 591−599.
Lowe, D.M., Tubbs, P.K., 1985b. 3-Hydroxy-3-methylglutaryl-coenzyme A synthase from ox liver:
properties of its acetyl derivative. Biochem. J. 227, 601−607.
Hepatic fatty acid oxidation and ketogenesis in young pigs 233

Lowe, D.M., Tubbs, P.K., 1985c. Succinylation and inactivation of 3-hydroxy-3-methylglutaryl-CoA synthase
by succinyl-CoA and its possible relevance to the control of ketogenesis. Biochem. J. 232, 37−42.
McGarry, J.D., Brown, N.F., 1997. The mitochondrial carnitine palmitoyltransferase system: from concept
to molecular analysis. Eur. J. Biochem. 244, 1−14.
McGarry, J.D., Foster, D.W., 1977. Hormonal control of ketogenesis. Arch. Intern. Med. 137, 495−501.
McGarry, J.D., Foster, D.W., 1980. Regulation of hepatic fatty acid oxidation and ketone body production.
Annu. Rev. Biochem. 49, 395−420.
Mersmann, H.J., Phinney, G., 1973. In vitro fatty acid oxidation in liver and heart from neonatal swine
(Sus domesticus). Comp. Biochem. Physiol. 44B, 219−223.
Mersmann, H.J., Goodman, J., Houk, J.M., Anderson, S., 1972. Studies on the biochemistry of mito-
chondrial and cell morphology in the neonatal swine hepatocyte. J. Cell. Biol. 53D, 335−347.
Mersmann, H.J., Phinney, G., Sanguinetti, M.C., Houk, J.M., 1973. Lipogenic capacity of liver from peri-
natal swine (Sus domesticus). Comp. Biochem. Physiol. 46B, 493−497.
Nicot, C., Hegardt, F.G., Woldegiorgis, G., Haro, D., Marrero, P.F., 2001. Pig liver palmitoyltransferase I,
with low Km for carnitine and high sensitivity to malonyl-CoA inhibition, is a natural chimera of rat
liver and muscle enzymes. Biochemistry 40, 2260−2266.
Odle, J., 1997. New insights into medium-chain triglyceride utilization by the neonate: observations from
a piglet model. J. Nutr. 127, 1061−1067.
Odle, J., 1998. Medium-chain triglycerides: a unique energy source for neonatal pigs. Pig News 20, 25−32.
Odle, J., Benevenga, N.J., Crenshaw, T.D., 1991a. Utilization of medium-chain triglycerides by neonatal
piglets: chain length of even- and odd-carbon fatty acids and apparent digestion/absorption and
hepatic metabolism. J. Nutr. 121, 605−614.
Odle, J., Benevenga, N.J., Crenshaw, T.D., 1991b. Postnatal age and the metabolism of medium- and
long-chain fatty acids by isolated hepatocytes from small-for-gestational-age and appropriate-for-
gestational-age piglets. J. Nutr. 121, 615−621.
Odle, J., Lin, X., van Kempen, T.A.T.G., Drackley, J.K., Adams, S.H., 1995. Carnitine palmitoyltrans-
ferase modulation of hepatic fatty acid metabolism and radio-HPLC evidence for low ketogenesis in
neonatal pigs. J. Nutr. 125, 2541−2549.
Oikawa, E., Iijima, H., Suzuki, T., Sasano, H., Sato, H., Kamataki, A., Nagura, H., Kang, M., Fujino, T.,
Suzuki, H., Yamamoto, T.T., 1998. A novel acyl-CoA synthetase, ACS5, expressed in intestinal
epithelial cells and proliferating preadipocytes. J. Biochem. 124, 679−685.
Ortiz, J.A., Mallolas, J., Nicot, C., Bofarull, J., Rodriguez, J.C., Hegardt, F.G., Haro, D., Marrero, P.F.,
1999. Isolation of pig mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase gene promoter: char-
acterization of a peroxisome proliferator-responsive element. Biochem. J. 337, 329−335.
Pegorier, J.P., Duee, P.H., Peret, J., Girard, J., 1981. Changes in circulating fuels, pancreatic hormones
and liver glycogen concentration in fasting or suckling newborn pigs. J. Dev. Physiol. 3, 203−217.
Pegorier, J.P., Duee, P.H., Girard, J., Peret, J., 1983. Metabolic fate of non-esterified fatty acids in
isolated hepatocytes from newborn and young pigs: evidence for a limited capacity for oxidation and
increased capacity for esterification. Biochem. J. 212, 93−97.
Pegorier, J.P., Garcia-Garcia, M.-V., Prip-Buus, C., Duee, P.-H., Kohl, C., Girard., J., 1989. Induction of
ketogenesis and fatty acid oxidation by glucagon and cyclic AMP in cultured hepatocytes from rabbit
fetuses. Biochem. J. 264, 93−100.
Prip-Buus, C., Pegorier, J., Duee, P., Kohl, C., Girard, J., 1990. Evidence that the sensitivity of carnitine
palmitoyltransferase I to inhibition by malonyl-CoA is an important site of regulation of hepatic fatty
acid oxidation in the fetal and newborn rabbit. Biochem. J. 269, 409−415.
Quant, P.A., Tubbs, P.K., Brand, M.D., 1989. Treatment of rats with glucagon or mannoheptulose
increases mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase activity and decreases succinyl-
CoA content in liver. Biochem. J. 262, 159−164.
Quant, P.A., Tubbs, P.K., Brand, M.D., 1990. Glucagon activates mitochondrial 3-hydroxy-3-methylglu-
taryl-CoA synthase in vivo by decreasing the extent of succinylation of the enzyme. Eur. J. Biochem.
187, 169−174.
Quant, P.A., Robin, D., Robin, P., Ferre, P., Brand, M.D., Girard, J., 1991. Control of hepatic mitochon-
drial 3-hydroxy-3-methylglutaryl-CoA synthase during the foetal/neonatal transition, suckling and
weaning in the rat. Eur. J. Biochem. 195, 449−454.
Quant, P.A., Robin, D., Robin, P., Girard J., Brand, M.D., 1993. A top-down control analysis in isolated
rat liver mitochondria: can the 3-hydroxy-3-methylglutaryl-CoA pathway be rate-controlling for keto-
genesis? Biochim. Biophys. Acta 1156, 135−143.
Reddy, J.K., Mannaerts, G.P., 1994. Peroxisomal lipid metabolism. Annu. Rev. Nutr. 14, 343−370.
234 J. Odle et al.

Robles-Valdes, C., McGarry, J.D., Foster, D.W., 1976. Maternal-fetal carnitine relationships and neona-
tal ketosis in the rat. J. Biol. Chem. 251, 6007−6012.
Rodríguez, J.C., Gomez, G.G., Hegardt, F.G., Haro, D., 1994. Peroxisome proliferator-activated receptor
mediates induction of the mitochondrial 3-hydroxy-3-methylglutaryl-CoA synthase gene by fatty
acids. J. Biol. Chem. 269, 18767−18772.
Rohde Parfet, K.A., Gonyou, H.W., 1988. Effect of creep partitions on teat-seeking behavior of newborn
piglets. J. Anim. Sci. 66, 2165.
Schmidt, I., Herpin, P., 1998. Carnitine palmitoyltransferase I (CPTI) activity and its regulation by mal-
onyl-CoA are modulated by age and cold exposure in skeletal muscle mitochondria from newborn
pigs. J. Nutr. 128, 886−893.
Stanier, M.W., Mount, L.E., Bligh, J., 1984. Energy Balance and Temperature Regulation. Cambridge
University Press, New York.
Swiatek, K.R., Kipnis, D.M., Mason, G., Chao, K., Cornblath, M., 1968. Starvation hypoglycemia in
newborn pigs. Amer. J. Physiol. 214, 400−405.
Tetrick, M.A., Adams, S.H., Odle, J., Benevenga, N.J., 1995. Contribution of D-(−)-3-hydroxybutyrate to
the energy expenditure of neonatal pigs. J. Nutr. 125, 264−272.
Thumelin, S., Forestier, M., Girard, J., Pegorier, J.P., 1993. Developmental changes in mitochondrial
3-hydroxy-3-methylglutaryl-CoA synthase gene expression in rat liver, intestine and kidney. Biochem. J.
292, 493−496.
Thumelin, S., Esser, V., Charvy, D., Kolodziej, M.P., Zammit, V.A., McGarry, J.D., Girard, J., Pegorier, J.P.,
1994. Expression of liver carnitine palmitoyltransferase I and II genes during development in the rat.
Biochem. J. 300, 583−587.
USDA, 1991. National swine survey: morbidity/mortality and health management of swine in the U.S.
National Animal Health Monitoring System APHIS/U.S. Department of Agriculture, Washington, DC.
USDA, 2002. Reference of swine health and environmental management in the United States, 2000, Part III.
U.S. Department of Agriculture/APHIS/VS/CEAH, National, Fort Collins, CO, No. N361.0902.
Van Veldhoven, P.P., Casteels, M., Mannaerts, G.P., Baes, M., 2001. Further insights into peroxisomal
lipid breakdown via α- and β-oxidation. Biochem. Soc. Trans. 29, 292−297.
Verhoeven, N.M., Roe, D.S., Kok, R.M., Wanders, R.J., Jakobs, C., Roe, C.R., 1998. Phytanic acid and
pristanic acid are oxidized by sequential peroxisomal and mitochondrial reactions in cultured fibro-
blasts. J. Lipid Res. 39, 66−74.
Wanders, R.J., Tager, J.M., 1998. Lipid metabolism in peroxisomes in relation to human disease. Mol.
Aspects Med. 19, 69−154.
Wanders, R.J.A., Vreken, P., Ferdinandusse, S., Jansen, G.A., Waterham, H.R., van Roermund, C.W.T.,
van Grunsven, E.G., 2001. Peroxisomal fatty acid α- and β-oxidation in humans: enzymology,
peroxisomal metabolite transporters and peroxisomal diseases. Biochem. Soc. Trans. 29, 250−267.
Williamson, D.H., Bates, M.W., Krebs, H.A., 1968. Activity and intracellular distribution of enzymes of
ketone-body metabolism in rat liver. Biochem. J. 108, 353−361.
Wolfe, R.G., Maxwell, C.V., Nelson, E.C., 1978. Effect of age and dietary fat level on fatty acid oxida-
tion in the neonatal pig. J. Nutr. 108, 1621−1634.
Yu, X.X., Drackley, J.K., Odle, J., Lin, X., 1997a. Response of hepatic mitochondrial and peroxisomal
β-oxidation to increasing palmitate concentrations in piglets. Biol. Neonate 72, 284−292.
Yu, X.X., Drackley, J.K., Odle, J., 1997b. Rates of mitochondrial and peroxisomal β-oxidation of palmitate
change during postnatal development and feed deprivation in liver, kidney and heart of pigs. J. Nutr.
127, 1814−1821.
Yu, X.X., Drackley, J.K., Odle, J., 1998. Food deprivation changes peroxisomal β-oxidation activity but
not catalase activity during postnatal development in pig tissues. J. Nutr. 128, 1114−1121.
Yu, X.X., Drackley, J.K., Odle, J., 2001. Differential induction of peroxisomal β-oxidation enzymes by
clofibric acid and aspirin in piglet tissues. Amer. J. Physiol. Regul. Integr. Comp. Physiol. 281,
R1553−R1561.
10 Essential fatty acid metabolism during
early development

S. M. Innis

Department of Paediatrics, University of British Columbia, Vancouver,


British Columbia, Canada V5Z 4H4

The n-6 and n-3 polyunsaturated fatty acids are essential nutrients that are required for growth
and normal cell function. These fatty acids are present in cells as the acyl moieties of phos-
pholipids which make up the structural matrix of cell and subcellular membranes, and
function directly, or as precursors to other molecules that modulate cell growth, metabolism,
inter- and intracellular communication and gene expression. The n-3 fatty acid docosahexaenoic
acid (22:6n-3) is accumulated in the retina and brain grey matter during development, but is
continually turned over, recycled and replenished by uptake from plasma during the dynamic
processes of signal transduction in the retina and neuronal membranes. Depletion of 22:6n-3
from retinal and neural membranes results in reduced visual function, behavioural abnormal-
ities, and alterations in the metabolism of neurotransmitters, and in membrane proteins,
receptors and ion channel activities. Large gaps still exist in understanding of dietary require-
mements for n-3 fatty acids at different stages of the life cycle, species differences in essential
fatty acid metabolism, and the process that controls the partitioning of n-3 fatty acids for
generation of energy and further metabolism for incorporation into membrane lipids.

1. OVERVIEW OF ESSENTIALITY OF n-6 AND n-3 FATTY ACIDS


Studies in animals over 70 years ago provided the first demonstration that dietary fat contains
components that are essential to normal growth and development. The signs of deficiency,
which included scaly skin, growth retardation, reproductive failure and histological abnor-
malities (table 1), were ascribed to the absence of n-6 fatty acids, and were reversed or
prevented by feeding the n-6 fatty acid linoleic acid (18:2n-6) (Innis, 1991). The importance
of a second class of polyunsaturated fatty acids, the n-3 fatty acids, did not emerge until the
1970s when altered electroretinograph (ERG) recordings and behaviour were found in rats
fed diets deficient in n-3 fatty acids (Benolken et al., 1973; Wheeler et al., 1975; Lamptey and
Walker, 1976). Despite the knowledge that the n-3 fatty acid docosahexaenoic acid (22:6n-3)

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
235 © 2005 Elsevier Limited. All rights reserved.
236 S. M. Innis

Table 1
Signs of n-6 and n-3 fatty acid deficiency

Essential fatty acid/n-6 fatty acid deficiency n-3 fatty acid deficiency

Decreased growth Reduced electroretinogram A and/or B wave recording


Scaly dermatoses Reduced visual acuity
Inflamed skin Reduced discrimination learning
Decreased skin pigmentation Reduced exploratory behaviour
Increased transepidermal water loss Increased stereotyped behaviour
Impaired wound healing
Alopecia
Tail necrosis
Fatty liver
Kidney degeneration
Reproductive failure

represents a major proportion of the polyunsaturated fatty acids in brain and retina (Fliesler
and Anderson, 1983; Sastry, 1985; Giusto et al., 2000), the acceptance of the essential role of
n-3 fatty acids, particularly in human nutrition, grew only slowly, perhaps because of the
absence of overt effects of n-3 fatty acid deficiency on growth or health. The formation and
extension of neural and retinal membranes, however, requires a large amount of 22:6n-3, all
of which must be derived from n-3 fatty acids in the diet. It is well appreciated that deficiency
of one or more key nutrients during brain development can, depending on the timing, sever-
ity and duration, decrease cell division, dendritic arborization and myelination with resultant
long-term effects on cognitive and behavioural functions (Dobbing et al., 1971; Dobbing,
1972; Wiggins, 1986; Morgane et al., 1992; Levitsky and Strupp, 1995). Because of this,
much of the recent interest on essential fatty acids in human growth and development has
focused on the requirements for n-3 fatty acids for visual and neural function. This chapter
reviews current research on polyunsaturated fatty acid metabolism in development, the supply
of n-6 and n-3 fatty acids before and after birth, and the role of n-3 fatty acids in the devel-
oping brain and retina.

2. ESSENTIAL FATTY ACID STRUCTURE AND METABOLISM


2.1. Structure and nomenclature

Fatty acids are identified using a systematic nomenclature which identifies the fatty acid by
the number of carbon atoms, and the number and position of any unsaturated double bonds
relative to the carboxyl group (table 2, fig. 1). More commonly, fatty acids are referred
to using a structural designation, or trivial name. The structural designation describes a
fatty acid by the number of carbons, the number of double bonds, and the position of the
first double bond from the methyl terminal carbon which is designated by the Greek letter
omega (ω) or “n”. The delta (Δ) notation is used to designate the position of a carbon from
the carboxyl terminus and is used to denote the site of action of the fatty acid desaturase
enzymes. Polyunsaturated fatty acids, which are often referred to as PUFA, are grouped into
families based on the position of the first methylene interrupted double bond from the methyl
(ω or n) end.
Essential fatty acid metabolism during early development 237

Table 2
Systematic names and abbreviations for essential n-6 and n-3 fatty acids

Systematic name Abbreviation Trivial name

n-6 family
9,12-octadecadienoic 18:2n-6 Linoleic
6,9,12-octadecatroenic 18:3n-6 γ-Linolenic
8,11,14-eicosatrienoic 20:3n-6 Dihomo-γ-linolenic
5,8,11,14-eicosatetraenoic 20:4n-6 Arachidonic
7,10,13,16-docosatetraenoic 22:4n-6 Adrenic
4,7,10,13,16-docosapentaenoic 22:5n-6
n-3 family
9,12,15-octadecatrienoic 18:3n-3 α-Linolenic
6,9,12,15-octadecatetraenoic 18:4n-3 Stearidonic
5,8,11,14,17-eicosapentaenoic 20:5n-3
4,7,10,13,16,19-docosahexaenoic 22:6n-3
6,9,12,15,18,21-tetracosahexanenoic 24:6n-3 Nisinic

2.2. Metabolism

The Δ12- and Δ15-desaturase enzymes necessary to form n-6 and n-3 fatty acids, respectively,
are present in plant but not in animal cells. Consequently, animals require a dietary source of
n-6 and n-3 fatty acids, all of which are ultimately derived from plants either directly, or after
transfer up the food chain. The parent 18-carbon n-6 and n-3 fatty acids are linoleic acid
(18:2n-6) and α-linolenic acid (18:3n-3). These fatty acids are considered essential dietary
nutrients for most animals.
Linoleic acid (18:2n-6) and 18:3n-3 can be further metabolized by desaturation and elon-
gation, as well as chain shortening, but cannot be interconverted (Innis, 1991) (fig. 2). The
major site of desaturation and elongation is the liver, although desaturase activity is also present
in some other cells. Desaturation and elongation proceed alternately, by Δ6-desaturation,
elongation and Δ5-desaturation, to yield arachidonic acid (20:4n-6) and eicosapentaenoic acid
(20:5n-3) from 18:2n-6 and 18:3n-3, respectively (fig. 2). These metabolites of 18:2n-6 and
18:3n-3 with 20 or more carbon atoms are often referred to as long-chain polyunsaturated
fatty acids, and abbreviated as LC-PUFA. For many years it was believed that docosapentaenoic
acid (22:5n-3) was desaturated by a microsomal Δ4-desaturase to form docosahexaenoic acid
(22:6n-3). Such a Δ4-desaturase, however, has not been found in animal cells (Moore et al., 1991;
Voss et al., 1991; Mohammed et al., 1995, 1997; Luthria et al., 1996). The pathway for

Fig. 1. Schematic representation of a fatty acid.


238 S. M. Innis

Fig. 2. Schematic representation of the major pathways of essential fatty acid desaturation and elongation.

synthesis of 22:6n-3 is now known to involve synthesis of a 24-carbon n-3 fatty acid by two
successive elongations of 20:5n-3 to form 24:5n-3, which is then desaturated at position 6 to
form tetracosahexaenoic acid (24:6n-3), and translocated to the peroxisomes where one cycle
of β-oxidation leads to formation of 22:6n-3 (Sprecher et al., 1995, 1999; Ferdinandusse et al.,
2001). The 22:6n-3 thus formed is then shuttled back to the endoplasmic reticulum, which is
the major site of phospholipid biosynthesis (Vance, 1990; Choy, 1997). Synthesis of the
n-6 docosapentaenoic acid (22:5n-6) from 20:4n-6 occurs in an analogous pathway (Sprecher
et al., 1995, 1999). The regulation and steps involved in the intracellular trafficking of fatty
acids between the endoplasmic reticulum and peroxisomes is still incompletely understood.
The desaturase enzymes necessary for synthesis of 20:4n-6 and 22:6n-3 are present in
animal but not in plant cells. Thus, omnivores and carnivores obtain 20:4n-6 and 22:6n-3
from the flesh of other animals, while herbivores consume only the precursors 18:2n-6 and
18:3n-3. The activity of the essential fatty acid desaturation and elongation pathway appears
to vary markedly among species, being higher in rodents and low in pigs and humans
(Pooviah et al., 1976; Yamazaki et al., 1992; Emken et al., 1994; Li et al., 2000; Pawlosky et al.,
2001). Rats, for example, have very high amounts of 22:6n-3 in liver phosphatidylethanolamine
(PE) when compared to humans and other primates (table 3). The felinae, which are obligate
carnivores, have very low activity of Δ6-desaturase, and thus require a dietary source of
Table 3
Essential n-6 and n-3 fatty acids in brain grey matter and liver phosphatidylethanolamine of different species

Brain Liver

18:2n-6 20:4n-6 22:4n-6 22:5n-6 22:6n-3 18:2n-6 20:4n-6 22:4n-6 22:5n-6 22:6n-3

Rat 0.3 11 5.1 0.6 21 6.0 21 0.6 0.4 22


Guinea pig 1.6 17 7.9 2.4 19 23 11 1.3 0.9 21
Hamster 1.5 9.6 4.2 0.2 22 11 16 0.3 0.5 10
Lemming 1.3 10 4.3 1.5 29 16 7.6 0.3 0.9 23
Bat 0.9 15 5.2 2.5 21 8 11 0.6 0.2 3.9
Vole 0.5 11 1.1 0.5 20 7.2 5.5 0.2 0.1 26
Warthog 0.9 12 6.3 1.2 23 22 12 0.6 0.2 1.6
Horse 1.0 12 7.1 1.3 19 17 6.2 0.6 0.4 3.6
Zebra 0.1 10 4.9 18 47 4.2 0.1 0.1 0.2
Elephant 1.1 11 7.2 2.1 25 10 10 0.4 0.1 0.6
Hartebeest 1.5 11 7.2 0.0 17 11 14 0.7 0.3 0.9
Red deer 0.8 10 6.9 1.2 25 5.1 11 0.1 0.1 0.3
Ox 1.2 11 6.3 0.8 22 4.0 16 2.3 0.6 0.3
Buffalo 2.3 12 6.5 1.1 16 9.7 13 0.3 0.1 2.3
Eland 1.7 14 7.0 1.2 21 8.5 14 0.6 0.2 1.8
Giraffe 1.4 14 6.2 0.9 24 12 12 1.1 0.8 0.8
Cat 0.9 13 6.7 1.1 27 5.5 18 1.3 2.0 22
Civet 1.2 14 5.9 1.3 26 3.4 15 3.6 1.2 20
Leopard 0.7 14 7.1 1.0 22 3.8 15 2.4 0.7 18
Marmoset 0.8 11 4.7 1.6 29 10 23 0.7 1.3 8.2
Vervet 0.1 6.8 5.3 1.2 21 9.2 17 0.3 0.4 5.9
Man 0.3 14 6.0 2.2 23 6.3 16 3.5 2.3 7.7
Dolphin 0.7 6.9 3.6 0.2 27 1.7 30 0.9 0.9 11
Adapted from Crawford et al. (1976).
240 S. M. Innis

20:4n-6 and 22:6n-3 (Rivers et al., 1975, 1976; MacDonald et al., 1983; Pawlosky et al.,
1997), which is consumed as part of their natural meat diet.
Many of the essential roles of n-6 and n-3 fatty acids are fulfilled by 20:4n-6 and 20:5n-3
and 22:6n-3, rather than their precursors 18:2n-6 and 18:3n-3, respectively. Linoleic acid
(18:2n-6) and metabolites in the pathway to 20:4n-6, however, have important roles in cho-
lesterol metabolism and in the skin, and 20:3n-6 is the precursor for synthesis of series 1
eicosanoids (Wertz et al., 1987; Ziboh and Chapkin, 1988; Innis, 1996). α-Linolenic acid, on
the other hand, is not known to have any essential functions other than as a precursor for syn-
thesis of 20:5n-3 and 22:6n-3. Acetyl-CoA derived from β-oxidation of 18:3n-3 is extensively
recycled for synthesis of saturated and monounsaturated fatty acids and cholesterol synthesis in
developing tissues (Cunnane et al., 1994, 1999; Sheaff-Greiner et al., 1996; Menard et al., 1998).
Desaturation of 18:2n-6 and 18:3n-3 is believed to involve the same enzymes. In vitro, the
Δ6-desaturase enzyme shows clear preference in order 18:3n-3 > 18:2n-6 > 18:1n-9 (Brenner
and Peluffo, 1966, 1969; Brenner et al., 1969). Owing to the abundance of 18:2n-6 in com-
monly used vegetable oils, such as safflower, corn and soybean oil (Chow, 2000), human diets
in many Western countries have much higher 18:2n-6 than 18:3n-3 (Simopoulos, 1999). The
high proportion of 18:2n-6 to 18:3n-3 in the diet has implications for reducing synthesis of
22:6n-3, which is important for brain and visual development, and for increasing the risk of
health problems associated with increased production of 20:4n-6-derived eicosanoids.
Synthesis of 22:6n-3 and 20:4n-6 is also reduced by products of the same and opposing series
of fatty acids. For example, high intakes of 20:5n-3 or 22:6n-3 from fish or fish oil reduces
tissue 20:4n-6, decreases the synthesis of n-6 fatty acid-derived eicosanoids, and increases the
synthesis of eicosanoids derived from 20:5n-3 (Fischer et al., 1989; Broughton and Morgan,
1994; Ferretti et al., 1998; Broughton and Wade, 2002). Changes in the balance of n-6 and
n-3 fatty acid-derived eicosanoids can have important effects on inflammation and immunity,
hemostatic and endothelial function, and reproductive functions including ovulation rate,
progesterone production by the corpus luteum, timing of luteolysis and gestational length
(von Schacky and Weber, 1985; Rogers et al., 1987; Kristensen et al., 1989; Tremoli et al.,
1995; Calder, 1998, 2001; Abayasekara and Wathes, 1999).
β-Oxidation of the 18-carbon essential fatty acids in the mitochondria depends on carnitine-
dependent translocation, and leads to generation of acetyl-CoA which then enters the tricar-
boxylic acid cycle. The first and rate-limiting step of β-oxidation of the longer-chain n-6 and
n-3 fatty acids in the peroxisomes is catalysed by straight-chain fatty acyl-CoA oxidase and
generates hydrogen peroxide (Wanders et al., 2001). In a similar process, 22:6n-3 and 22:4n-6
can be retroconverted to 20:5n-3 and 20:4n-6, respectively, thereby maintaining tissue pools
of these fatty acids.
In the absence of a dietary supply of n-6 and n-3 fatty acids, oleic acid (18:1n-9) derived
from the diet or synthesized de novo from acetyl-CoA, undergoes Δ6- and Δ5-desaturation
and elongation to form eicosatrienoic acid (20:3n-9), and concentrations of 20:4n-6 decrease
due to the absence of 18:2n-6 (Innis, 1991, 1996). The usual biochemical method for estab-
lishing essential fatty acid deficiency is to calculate the ratio of 20:3n-9 to 20:4n-6. This is
commonly referred to as the triene to tetraene ratio. An increase in the ratio of plasma 20:3n 9
to 20:4n-6 to >0.2 is considered to indicate essential fatty acid deficiency in humans (Holman
et al., 1991; Jeppensen et al., 1998). Dietary deficiency of n-3 fatty acids results in decreased
22:6n-3 and increased desaturation of n-6 series fatty acids, leading to increased 22:4n-6 and
22:5n-3 in brain and retinal membranes (Bourre et al., 1984; Innis, 1991). Refeeding deficient
developing and adult animals with 18:3n-3 results in recovery of neural cell membrane
22:6n-3, although the rate of recovery may be slower in the central nervous system than in
Essential fatty acid metabolism during early development 241

other organs (Youyou et al., 1986; Bourre et al., 1989a; Connor et al., 1990; Weisinger et al.,
1999; Moriguchi et al., 2001). Desaturation and elongation of n-9 fatty acids and the triene/
tetraene ratio is normal unless there is a concomitant deficiency of n-6 fatty acids. The com-
pensatory increased desaturation and elongation of n-6 fatty acids in n-3 fatty acid-deficient
animals results in maintenance of the normal total n-6 + n-3 polyunsaturated fatty acids in the
brain (Galli et al., 1971; Neuringer et al., 1986; Hrboticky et al., 1990), although this may not
be so for all neural membrane phospholipids (Murthy et al., 2002).

3. FUNCTIONAL ROLES OF ESSENTIAL FATTY ACIDS


Long-chain polyunsaturated fatty acids (those with 20 or more carbon chains and 3 or more
double bonds) are found predominately in phospholipids in which they form the hydrophobic
core of all cell and subcellular membranes. Linoleic acid (18:2n-6) is also present in mem-
brane phospholipids, adipose tissue triglycerides, plasma cholesterol esters, and in the
specialized lipids of the skin. Reviews of the role of n-6 fatty acids in maintaining the normal
epithelial cell–water barrier are available (Ziboh and Chapkin, 1988). The n-6 and n-3 fatty
acids, 20:3n-6, 20:4n-6 and 20:5n-3 are precursors for prostaglandins, hydroxy fatty acids,
leukotrienes and lipoxins, often collectively referred to as eicosanoids. These oxygenated
metabolites, which are formed via cyclo-oxygenase and lipoxygenase, are synthesized fol-
lowing a stimulus and act locally as autocoids, often initiating a cascade of events. In general,
metabolites formed from 20:5n-3 have weaker or opposing effects than the metabolites
formed from 20:4n-6. Several reviews on eicosanoid metabolism have been published
(Fischer, 1989; Kinsella and Lokesh, 1990; Funk, 2001).
The importance of n-6 and n-3 fatty acids in metabolic and physiological processes can be
summarized into three general mechanisms: the fatty acid moieties of membrane phospho-
lipids contribute to the physical properties of the membrane bilayer, with secondary effects
on the activity of membrane-associated proteins, receptors and ion channels; n-6 and n-3 fatty
acids are precursors for generation of membrane-derived signal molecules, as well as
eicosanoids; and n-6 and n-3 fatty acids have direct effects on gene expression. Docosahexaenoic
acid, (22:6n-3), unlike the n-6 fatty acids, has a highly specific distribution in tissues and
phospholipids. Concentrations of 22:6n-3 are particularly high in the amino phospholipids,
phosphatidylserine (PS) and phosphatidylethanolamine (PE) of neural grey matter, and in the
outer segments of rod and cone photoreceptors in the retina (Fliesler and Anderson, 1983;
Sastry, 1985; Giusto et al., 2000). Large amounts of 22:6n-3 are also present in specific phos-
pholipids in the heart and sarcoplasmic reticulum of skeletal muscle sarcolemma, and in
sperm (Fiehn and Pewter, 1971; Poulos et al., 1973; Gudbjarnason et al., 1978; Charnock et al.,
1983; Ollero et al., 2000). Severe restriction of n-3 fatty acids throughout gestation, lactation
and postnatal development, when there is a need for new tissue synthesis, results in reduced
tissue 22:6n-3, decreased visual function, decreased performance on discrimination learning
tasks, and increased stereotyped behaviour in rodents and non-human primates (Benolken et al.,
1973; Wheeler et al., 1975; Neuringer et al., 1984, 1986; Yamamoto et al., 1988; Bourre et al.,
1989b; Reisbick et al., 1990, 1994; Frances et al., 1996a,b; Okada et al., 1996; Gamoh et al., 1999;
Moriguchi et al., 2000; Greiner et al., 2001) (table 1). In adult animals, 22:6n-3 is aggres-
sively retained, even during longstanding and severe dietary n-3 fatty acid restriction (Tinoco
et al., 1979; Tinoco, 1982). Some species, including many fish, insects and pulmonates, require
a dietary source of 18:3n-3 for normal growth and feed efficiency (Tinoco et al., 1979; Tinoco,
1982). n-3 fatty acids do not appear to be essential for growth and feed efficiency in mammals,
although 22:6n-3 is involved in energy metabolism and calcium ion channel activity in the
242 S. M. Innis

heart (Kang and Leaf, 1996; Leaf et al., 1999; Leifert et al., 1999; McLennan, 2001; Ferrier
et al., 2002), which may explain the relation of high dietary intakes of n-3 fatty acids with
reduced heart rate variability and sudden death (Christensen et al., 1996, 1997; Nair et al.,
1997; Albert et al., 1998).
When released by phospholipases, membrane phospholipid n-6 and n-3 fatty acids become
available as unesterified fatty acids and function as important signal molecules, in addition to
serving as substrates for eicosanoid synthesis. Their functions include regulation of the activity
of protein kinases, G-proteins, adenylate and guanylate cyclases, phospholipases, ion channels
and multiple other proteins and receptors (Bernsohn and Spitz, 1974; Bourre et al., 1989b; Park
and Ahmed, 1992; Gerbi et al., 1994; Poling et al., 1995; Kang and Leaf, 1996; Litman and
Mitchell, 1996; Koenig et al., 1997; Huster et al., 1998; Leifert et al., 1999; Bonin and Khan,
2000; Litman et al., 2001). In addition, polyunsaturated fatty acids regulate the expression
of genes for regulatory proteins of lipid and carbohydrate metabolism through peroxisome
proliferator-activated receptor (PPAR)-dependent and independent mechanisms (Duplus et al.,
2000; Clarke, 2001; Jump, 2002), and influence leptin gene expression (Reseland et al., 2001).
New information also suggests that n-6 fatty acids are involved in adipogenesis during devel-
opment by pathways involving PPAR γ2, although polyunsaturated fatty acids also suppress
genes related to lipogenesis (Ntambi et al., 1988; Clarke et al., 1990; Reginato et al., 1998).
Recently, it has become clear that n-3 fatty acids alter the expression of genes related to endo-
cytosis, signal transduction, synaptic vesicle recycling and formation, lipid metabolism,
nuclear ligand-activated transcription factor receptors in brain, retinoic acid receptor (R × R)
(Khair-El-Din et al., 1996; Mata de Urquiza et al., 2000; Kitajka et al., 2002), and intestinal
nutrient absorption (Lampen et al., 2001).

3.1. n-3 essential fatty acids and visual function

The retina is an integral part of the central nervous system, which is composed of six cell
types – the photoreceptor cells, horizontal, bipolar, amacrine, interplexiform and ganglion cells –
and communicates directly with the brain via ganglion cells passing through the optic nerve.
The two photoreceptor cell types are the rods and cones. Rods are elongate and cylindrical
and function as dim light receptors, while cones are shorter and usually cylindrical, mediate
colour vision and function at relatively higher light intensities (Giusto et al., 2000). Rods and
cones are highly specialized differentiated neurons that contain a stack of photosensitve
membranes at the distal end (known as the outer segments), a central region containing mito-
chondria, golgi and nucleus, and a synaptic terminal. The outer segments are made up of
densely stacked disks, each of which is a double layer of infolded plasma membrane which
is highly enriched in 22 :6n-3. Vertebrate retina photoreceptor cells contain 50% protein and
50% lipid, with 90–95% of the lipid present as phospholipid and 4–6% as cholesterol (Giusto
et al., 2000). The major phospholipid species are PE, PC and PS. Within the outer segment
disks, as much as 80% of the polyunsaturated fatty acids are 22:6n-3, with species of PE, PS
and PC in which both fatty acids are present as 22:6n-3 (Fliesler and Anderson, 1983;
Aveldano, 1987; Giusto et al., 2000). In bovine rod outer segment membranes, about 30% of
the PC, 20% of the PE and 50% of the PS have a long-chain polyunsaturated fatty acid ester-
ified at both the sn-1 and the sn-2 position (Aveldano et al., 1983; Aveldano, 1987; Aveldano
and Sprecher, 1987). It is notable that this unusual and characteristic membrane enrichment
of 22:6n-3 is present even in the bovine retina, a herbivore species that obtains no dietary
22:6n-3. Specific pathways allow efficient recycling of 22:6n-3 from photoreceptor cells
during shedding (turnover) from the disk tips. This involves phagocytosis by retinal pigment
Essential fatty acid metabolism during early development 243

epithelium cells, then recycling of 22:6n-3 for reuse in synthesis of new disk membranes
(Rodriguez de Turco et al., 1999).
About 80–90% of the retina rod and cone outer segment protein is the visual pigment
rhodopsin (opsin plus the carotenoid 11-cis-retinal), which functions as a photon receptor
coupled to regulatory G-proteins (Giusto et al., 2000). A light-induced change in the confor-
mation of rhodopsin triggers a cascade of reactions that result in increased phosphodiesterase
activity and decreased cyclic GMP, which leads to closure of the photoreceptor membrane
sodium channels and hyperpolarization, followed by depolarization. These events correspond
to the positive A and negative B waves, respectively, of the electroretinograph (ERG). Recent
studies have provided evidence that 22:6n-3 may influence photoreceptor signal transduction by
influencing the ability of photons to transform rhodopsin from metarhodopsin I to the activated
metarhodopsin II state (Mitchell et al., 1992; Mitchell and Litman, 1998; Litman et al., 2001).
Retina concentrations of 22:6n-3 increase during gestation and reach adult concentrations by
the time of term birth in humans (Martinez, 1992). Dietary deficiency of n-3 fatty acids during
development results in reduced 22:6n-3 in the retina of animals (Hrboticky et al., 1991). In
addition to 22:6n-3, n-3 fatty acids with up to 36 carbons are present in small amounts in the
retina (Aveldano et al., 1983; Aveldano, 1987; Aveldano and Sprecher, 1987), and these are
reduced in rats fed a diet deficient in n-3 fatty acids throughout development (Suh et al., 1996,
2000). The role of these fatty acids in retinal function has not yet been elucidated, although
they may be related to rhodopsin kinetics.
Several studies have addressed the effect of decreased retina 22:6n-3 on retinal and visual
function in animals and human infants. Early studies in this field reported increased A and B
wave amplitudes in ERG responses of rats fed a fat-free diet (Benolken et al., 1973; Wheeler
et al., 1975; Anderson et al., 1976). Later, Neuringer et al. (1984, 1986) reported prolonged
recovery times of dark-adapted A and B wave responses, and reduced rod and cone A wave
responses in full-field ERG of monkeys fed a severely n-3 fatty acid-restricted diet through
fetal and neonatal development. The role of n-3 fatty acids in retinal and visual development
in human infants is discussed in section 8.

3.2. n-3 essential fatty acids and brain function

The brain contains the second highest concentration of lipid in the body, after adipose tissue,
with 50% lipid on a dry weight basis, 10% lipid on a wet weight basis (Sastry, 1985). Unlike
adipose, however, the brain contains minimal amounts of triglyceride, the lipids of brain
being almost entirely composed of the membrane structural components. About half of the
lipid is phospholipid, with about 20% cholesterol, 15–20% cerebrosides and smaller amounts
of sulphatides and gangliosides. The phospholipids of brain grey matter contain large
amounts of 20 : 4n-6 and 22: 4n-6, particularly in PI and PE, and high amounts of 22:6n-3 in
grey matter PE and PS (Sastry, 1985). Myelin, on the other hand, contains mainly saturated
and monounsaturated fatty acids (O’Brien and Sampson, 1965). The brain of herbivores, like
the visual photoreceptor cells, is enriched in 22:6n-3 despite the absence of a dietary intake
of preformed 22 : 6n-3, and only low amounts of 22:6n-3 in liver phospholipids (table 3).
Only small amounts of 18 : 2n-6 are present in neural phospholipids, usually less than 1% of
all the fatty acids, and concentrations of 18:3n-3 are even lower. This unusual characteristic
feature of brain suggests the presence of specific pathways for selective uptake of 20:4n-6
and 22:6n-3 from plasma against a considerable concentration gradient, and that 20:4n-6
and 22:6n-3 are important to normal neural metabolism. The enrichment of 22:6n-3 in
mammalian brain grey matter, together with the inability of animals to form n-3 fatty acids
244 S. M. Innis

de novo, has focused attention on the need to elucidate the role of 22:6n-3 in neural function
and requirements for n-3 fatty acids during brain growth and development.
The behaviour changes in developing animals fed diets deficient in n-3 fatty acids include
reduced performance in maze tasks, habituation, exploratory activity in novel environments
and brightness and olfactory-based discrimination learning (Yamamoto et al., 1988; Bourre
et al., 1989b; Enslen et al., 1991; Innis, 1991; Yonekubo et al., 1993; Frances et al., 1996a,b;
Greiner et al., 1997, 2001; Wainwright et al., 1998). Polydypsia, increased stereotypic (loco-
motor) activity and altered performance on a task of recognition looking memory have also
been reported for monkeys fed a diet very low in n-3 fatty acids (<0.1% dietary energy 18:3n-3)
through gestation and postnatal development (Reisbick et al., 1990, 1994). In addition to the
decrease in 22:6n-3, n-3 fatty acid deprivation results in increased n-6 fatty acids, including
20:4n-6, 22:4n-6 and 22:5n-6, in the brain of animals (Galli et al., 1971; Bourre et al., 1984;
Hrboticky et al., 1989, 1990; Favrelière et al., 1998) and human infants (Farquharson et al.,
1995). It needs to be considered whether or not functional and biochemical changes associ-
ated with n-3 fatty acid deficiency are due to a specific requirement for 22:6n-3, the increase
in n-6 fatty acids, or a combination of effects.
Although considerable information has been published to describe the changes in brain and
other tissues of animals fed diets varying in essential fatty acids, less is known on how changes
in n-3 fatty acids alter behaviour. Recent research has established that 22:6n-3 alters metab-
olism of several neurotransmitters, including dopamine, serotonin (5-HT), epinephrine,
acetylcholine, GABA and N-methyl-D-aspartate (NMDA) channel activity (Delion et al., 1994,
1996, 1997; Nishikawa et al., 1994; Hamano et al., 1996; Jones et al., 1997; Minami et al.,
1997; Young et al., 1998; Zimmer et al., 1998, 1999, 2000a,b, 2002; de la Presa Owens and
Innis, 1999a,b; McGahon et al., 1999; Itokazu et al., 2000). The cortical dopaminergic system
is important in modulation of learning, attention and motivation, and in the visual pathways
(Gava and McKean, 1977; Brozoski et al., 1979; Le Moal and Simon, 1991; Antal et al., 1997;
Basmak et al., 1999) and has been the focus of several studies with n-3 fatty acid-deficient
animals. Usually, these studies involve feeding a diet severely restricted in n-3 fatty acids
through several generations in order to deplete adipose tissue n-3 fatty acids that can be trans-
ported from mother to offspring during gestation and suckling.
Newer research has shown that the effects of n-3 fatty acid deficiency on the brain are
complex and region specific. The effects described include changes in dopamine concentra-
tion, vesicular monoamine transporter 2, dopamine D2 receptor, tyrosine hydroxylase (the rate-
limiting enzyme in dopamine synthesis), and dopamine storage pools (Delion et al., 1994, 1996,
1997; Yoshida et al., 1997; Zimmer et al., 1998, 1999, 2000a,b, 2002). Newer techniques of
dual-probe microdialysis have shown that although dopamine is decreased in frontal cortex
of n-3 fatty acid-deficient animals, dopamine may be increased in the nucleus accumbens
(Zimmer et al., 2002). This could suggest that the mesocorticolimbic area dopaminergic
system functions more, but the mesocortical pathway is less active in n-3 fatty acid-deprived
animals. Newer studies also suggest that the effects of n-3 fatty acid deficiency may be more
apparent after a learning task, or administration of drugs that deplete endogenous dopamine
storage pools (Yoshida et al., 1997; Zimmer et al., 1998, 2000a,b).
Dietary deficiency of n-3 fatty acids leading to decreased 22:6n-3 in the developing brain
is also associated with reduced serotonin and serotonin receptor binding in frontal cortex
(Delion et al., 1996; de la Presa Owens and Innis, 1999a,b). This is of interest because 5-HT2
receptors are believed to play an important role in behaviour (Leysen and Pauwels, 1990).
Cross-sectional studies in humans have described a negative relation between 5-hydroxy
indolaceteic acid (5-HIAA, which is the metabolite of serotonin) in cerebrospinal fluid and
Essential fatty acid metabolism during early development 245

plasma 22:6n-3 among adults with violent behaviours (Hibbeln et al., 1998). Other aggres-
sive and depressive affective disorders have also been postulated to be linked to altered n-3
fatty acid metabolism (Adams et al., 1995; Hamazaki et al., 1996; Edwards et al., 1998;
Hibbeln, 1998; Maes et al., 1999; Stoll et al., 1999). In vascular smooth muscle cells, enrich-
ment with 22:6n-3 results in failure to respond to serotonin (Pakala et al., 2000). In other
tissues, 20:5n-3 and 22:6n-3 both block the effect of growth factors that act through sig-
nalling pathways involving receptor tyrosine kinase (such as platelet-derived growth factor,
fibroblast growth factor, epidermal growth factor and insulin-like growth factor) and G-protein
coupled receptors (such as bomberin, bradykinin, vasopressin, thrombin, serotonin and throm-
boxane A2) (Kaminski et al., 1993). These findings have potential implications for linking the
effects of 22:6n-3 to altered behaviour, and for a wide spectrum of physiological processes in
developing animals.
The abundance of 22:6n-3 in synaptic terminals and cortical growth cone membranes sug-
gests n-3 fatty acids could be important for brain structural growth. Morphological analysis
of brain from rats fed a n-3 fatty acid-deficient diet through three generations have described
a lower cell body size in CA1 pyrimidal neurons at the septal location, although no differ-
ences were found in other hippocampal regions, or in neuron volume, density or number
(Ahmad et al., 2002a). Consistent with this, others have shown reduced nerve growth factor
concentrations in hippocampus of n-3 fatty acid-deficient rats (Ikemoto et al., 2000). Lower
neuron size in hypothalamus and parietal cortex of weanling rats, and in periform cortex of
mature rats with a 90% decrease in brain 22:6n-3, has also been reported (Ahmad et al.,
2002b). Non-randomized studies in children with disorders of peroxisomal biogenesis have
provided evidence that supplemental 22:6n-3 can lead to improvements in the marked deficits
in cognitive and visual function, and impaired myelination in these disorders (Martinez and
Vazquez, 1998; Martinez et al., 2000). In these disorders, the absence of peroxisomes results
in the inability to form 22:6n-3. A mechanism through which n-3 fatty acids could stimulate
oligodendrocyte metabolism and the synthesis of myelin has not been described.
Information to suggest that n-3 fatty acids influence synthesis and turnover of brain phos-
pholipids has also been published. Elucidation of this possibility is important because
decreased phospholipid synthesis or turnover could have profound effects on membrane
phospholipid-dependent signalling pathways. Studies with F2 generation 8-week-old rats
found decreased PS in olfactory bulb, brain cortex and brain mitochondria, while PS in liver
and adrenal were not affected (Hamilton et al., 2000). In other studies, PS was reduced by
28% and PC was increased in hippocampus phospholipids of rats fed a n-3 fatty acid-deficient
diet (Murthy et al., 2002). Further evidence for a role of 22:6n-3 in PS metabolism has come
from studies showing that intra-amniotic injection of ethyl-22:6n-3 increased both 22:6n-3
and PS in brain of fetal rats (Green and Yavin, 1995). PS synthesis was also increased in C6
glioma cells cultured with 22:6n-3 (Garcia et al., 1998). The significance of these findings
relates to the role of PS in signal transduction through a regulatory role in protein kinase C
(PKC) activation (Bell, 1986; Bell and Burns, 1991; Casamenti et al., 1991; Borghese et al.,
1993; Mosier and Newton, 1998). Some evidence to suggest that dietary essential fatty acids
affect the activity of enzymes involved in PC biosynthesis in brain synaptic membranes has
also been published (Hargreaves and Clandinin, 1987). Studies using radiolabelled tracers
have also provided evidence of decreased docosahexaenoyl-CoA, and phospholipid synthesis
and turnover in vivo in n-3 fatty acid-deficient animals (Gazzah et al., 1995; Contreras et al.,
2001). Chronic n-3 fatty acid deficiency, however, does not affect 20:4n-6 recycling in the brain,
which suggests that phospholipases involved in release of 20:4n-6 and 22:6n-3 may be regu-
lated independently (Contreras et al., 2001). Decreased turnover of membrane phospholipids
246 S. M. Innis

enriched in 22:6n-3, reflecting adaptation by the brain to conserve important membrane


polyunsaturated fatty acids during limited dietary availability, has important implications for
optimal signal transduction.

4. ESSENTIAL FATTY ACID TRANSFER IN GESTATION


All of the n-6 and n-3 fatty acids accumulated in the fetus must ultimately be derived from
the mother by placental transfer. Placental tissue contains Δ6- and Δ5-desaturases. The activity
of Δ6-desaturase in ovine placenta increases towards term (Shand and Noble, 1981; Choy et al.,
1997). Synthesis of 22:6n-3 by the placenta, however, has not been demonstrated. Whether
placental desaturase activity contributes 20:4n-6 to the foetus does not appear to be known.
Concentrations of 20:4n-6 and 22:6n-3 are much higher, whereas 18:2n-6 is lower in foetal
than maternal plasma lipids (Crawford et al., 1981; Elias and Innis, 2001) (fig. 3). Current
information suggests that placental transfer of 20:4n-6 and 22:6n-3 involves a multi-step
process of cell uptake and intracellular translocation that is facilitated by several membrane-
associated and cytosolic fatty acid binding proteins. Specific membrane binding of n-6 and
n-3 fatty acids, favouring 20:4n-6 and 22:6n-3 over 18:2n-6 or 18:3n-3, and preferential trans-
fer of n-6 and n-3 fatty acid compared to non-essential fatty acids by human placenta have
been reported (Campbell et al., 1996, 1998a,b; Haggarty et al., 1997; Dutta-Roy, 2000).
Despite the presence of pathways to facilitate the transfer of n-6 and n-3 fatty acids across
the placenta, the maternal intake of essential fatty acids during pregnancy can have a marked
effect on n-6 and n-3 fatty acid accretion in developing foetal tissues. Dietary deficiency of
n-3 fatty acids in gestation results in decreased 22:6n-3 and increased 22:4n-6 and 22:5n-6 in
growth cones, the amoeboid leading edge of the growing neurite, in the fetal rat brain while
high intakes of 20:5n-3 and 22:6n-3 result in increased 22:6n-3 and decreased n-6 fatty acids
(Innis and de la Presa Owens, 2001) (fig. 4). In the latter studies, differences in maternal

Fig. 3. The relative enrichment of 18:2n-6, 18:3n-3, 20:4n-6 and 22:6n-3 in fetal compared to maternal
plasma was calculated for each mother–fetal cord plasma pair as the difference in the given fatty acid in the
maternal compared to fetal plasma/maternal plasma × 100%. Values shown are means ± SEM, n=55. Adapted
from Elias and Innis (2001).
Fig. 4. Effect of maternal essential fatty acid intake on n-6 and n-3 fatty acids in fetal rat brain neuronal growth cone phosphatidylcholine (PC), phosphatidylethanolamine (PE), phos-
phatidylserine (PS) and phosphatidyliniositol (PI). Rats were fed safflower, soybean or fish oil throughout pregnancy and fetal brain was analysed at birth. The bars represent means + SEM,
values with a different superscript are different (P< 0.05). Adpated from Innis and de la Presa Owens (2001).
248 S. M. Innis

dietary fat intake also resulted in altered dopamine in the fetal rat brain. Levels of 20:4n-6 and
22:6n-3 in maternal plasma are also significantly correlated with the concentration of the
same fatty acid in newborn human infant plasma (Elias and Innis, 2001), and infants born
with higher blood levels of 22:6n-3 and 20:4n-6 maintain this advantage for several weeks
after birth (Foreman-van Drangelen et al., 1995; Guesnet et al., 1999). This adipose tissue
22:6n-3 and 20:4n-6 accumulated during foetal development contributes to the postnatal blood
pool of 22:6n-3 and 20:4n-6. Several studies have shown that supplementation of pregnant
women with 22:6n-3 increases 22:6n-3 in plasma and red blood cell lipids of infants at birth
(van Houwelingen et al., 1995; Connor et al., 1996; Helland et al., 2001), providing convincing
evidence that the maternal intake of 22:6n-3 is an important factor in the placental transfer of
this fatty acid.
Research is now starting to address the functional significance of the supply of 20:4n-6 and
22:6n-3 to the developing foetus. A positive relation between 20:4n-6 and birth weight, despite
high 18:2n-6, has been found in several studies (Koletzko and Braun, 1991; Elias and Innis,
2001). Although this is consistent with the essentiality of 20:4n-6 in growth (Mohrhauer and
Holman, 1963), the biological explanation for this is not known. Electroencephalography
(EEG) at 2 days, as a measure of infant CNS maturity, indicated infants with a more mature
EEG pattern have significantly higher 22:6n-3 in cord plasma phospholipids than infants with
a less mature EEG pattern (Helland et al., 2001). Other recent studies have found an inverse
relationship between maternal plasma 22:6n-3 and active sleep and sleep–wake transitions,
and a positive association with wakefulness in 2-day-old infants (Ceruku et al., 2002), also
suggesting that lower 22:6n-3 during gestation may be associated with delayed CNS maturation.
Evidence of long-term sequelae, should these be present, due to early differences in 22:6n-3
status at birth is still needed, as are intervention studies to rule out the possibility that these
associations are explained by differences in other dietary and lifestyle variables that accom-
pany differences in 22:6n-3 intake.
Long-chain polyunsaturated n-3 fatty acids, and the balance of n-6 to n-3 fatty acids in the
diet, also influences gestation length in animals and humans (Olsen et al., 1992, 1995;
Abayasekara and Wathes, 1999). In humans, high habitual intakes of 20:5n-3 and 22:6n-3,
and supplementation with 2.7 g/day fish oil have been associated with longer gestation, pos-
sibly due to suppression of n-6 fatty acid-derived eicosanoids that are involved in initiation of
parturition (Olsen et al., 1991, 1992). However, no association was found between length of
gestation and the intake of n-3 fatty acids among women for whom the 95% range of intake
was 0 to 0.75 g/day (Olsen et al., 1995). Typical intakes of long-chain n-3 fatty acids in
North America are in the range of 100–200 mg/person/day (Innis and Elias, 2003). Whether
the small increase in length of gestation in humans (of about 3–4 days) found at very high
intakes of 20:5n-3 + 22:6n-3 is of functional benefit to the infant is not known.

5. ESSENTIAL FATTY ACID TRANSFER IN MILK


The n-6 and n-3 fatty acid composition of milk varies considerably among different species,
probably reflecting differences in diet and lipid metabolism. The milks of ruminants, such as
cattle and goats, have low concentrations of all the n-6 and n-3 fatty acids, while concentra-
tion of 20:5n-3 and 22:6n-3 are particularly high in the milk of marine mammals (table 4).
In addition to the effects of the diet, species differences in the desaturation and elongation, or
β-oxidation, of 18:2n-6 and 18:3n-3 also influence the amount of 20:4n-6 and 22:6n-3 in
milk. For example, milk from rats fed a diet containing 18:2n-6 and 18:3n-3, but no 20:4n-6 or
22:6n-3, had 0.8% 20:4n-6 and 0.5% 22:6n-3. Milk from pigs fed the same fat source, on the
Essential fatty acid metabolism during early development 249

Table 4
Fatty acid composition of selected animal and human milk lipids

Cowa Goatb Pigc Catd Dogd Rate Whalee Dolphine Sea lione Humanf

<10:0 10.8 22.4 1.0 0.4 0.2 14.8 0.6


12:0 3.0 6.4 4.1 0.9 0.7 10.7 0.1 0.2 0.1 4.1
14:0 10.6 12.4 3.1 5.4 4.9 11.6 6.1 5.6 3.7 6.1
16:0 33.7 33.7 28.0 26.8 26.5 23.5 12.8 12.9 18.0 19.4
18:0 12.6 3.8 5.1 10.1 3.7 4.8 5.0 4.8 2.1 7.2
16:1 1.8 9.4 5.1 7.8 1.3 7.3 8.1 7.6 2.5
18:1 21.4 11.3 13.2 40.3 10.9 17.8 25.6 38.2 40.1 35.7
18:2n-6 2.9 2.7 5.8 9.2 10.4 0.3 1.1 1.5 12.1
20:4n-6 0.2 1.1 0.8 0.8 0.5 1.1 0.4
18:3n-3 0.3 0.3 0.6 1.3 2.1 0.8 0.8 1.1 1.5 1.4
20:5n-3 0.2 0.4 6.5 4.2 5.5 0.1
22:6n-3 0.1 0.5 9.6 6.0 7.1 0.2
Adapted from: a Jensen (2002); b Le Doux et al. (2002); c Foote et al. (1990); d Parodi (1982); e Unpublished data;
f Innis and King (1999).

other hand, had 0.6% 20:4n-6 and only 0.1% 22:6n-3. Human milk also typically has about
0.5–0.7% 20:4n-6 and 0.2–0.4% 22:6n-3, with the amount of 22:6n-3 decreasing to 0.1%
fatty acids among women following vegan diets with no preformed 22:6n-3 (Sanders et al.,
1978; Innis, 1992; Innis and King, 1999).
Studies in multiple species have shown that the amounts of 18:2n-6, 18:3n-3, 20:5n-3 and
22:6n-3 secreted in milk depends on the amount of the same fatty acid in the maternal diet.
As for other dietary situations, this can be expected to influence the amount of essential fatty
acids deposited in the tissues of the growing milk-fed neonate. Studies in animals, including
cattle, have shown that increasing the dietary intake of 18:2n-6 results in increased secretion
of 18:2n-6 in milk (Bayourthe et al., 2000; Morales et al., 2000). Similarly, supplementation
of diets with fish oil results in increased secretion of 20:5n-3 and 22:6n-3 in milk (Arbuckle
and Innis, 1993). The increased 20:5n-3 and 22:6n-3 in the milk of lactating sows fed 1%
weight diet as fish oil results in marked increases in 20:5n-3 and 22:6n-3 in plasma, red blood
cells and liver, and a smaller increase in 22:6n-3 in brain of the suckling piglets (table 5). The
implications of the polyunsaturated fatty acid supply in milk to the growth, behaviour and
health of livestock is worthy of consideration.
Autposy analysis has confirmed that, as in animals, the essential fatty acid composition
of the diet does influence the accretion of n-6 and n-3 fatty acids in developing human infant
tissues (Makrides et al., 1994; Farquharson et al., 1995; Jamieson et al., 1999). Because of this,
it is important to understand the effects of diet on the essential fatty acid composition of
human milk. The essential fatty acid composition of human milk shows considerable vari-
ability both among and within populations (Jensen, 1989, 1999; Innis, 1992). This variability
appears to be explained by differences in the maternal essential fatty acid intake. Human milk
in North America and Europe currently has 12–16% total fat as 18:2n-6, representing about
6–8% of the infant’s energy intake (Innis, 1992; Innis and King, 1999). Reports from the
1950s found 7% 18:2n-6 in human milk lipids, and clearly showed that a diet high in 18:2n-6
results in a marked increase in 18:2n-6 in the milk of lactating women (Insull and Ahrens,
1959; Insull et al., 1959). Whether the 2-fold increase in 18:2n-6 in human milk is explained
by changes in dietary fat over the last half century is unclear.
Table 5
Polyunsaturated fatty acids in milk lipids and suckling piglets from sows fed vegetable and fish oils

Sow milk Piglet plasma Liver PE Cerebrum PE Retina

Veg. oil +Fish oil Veg. oil + Fish oil Veg. oil + Fish oil Veg. oil + Fish oil Veg. oil + Fish oill

16:0 27.6 28.6 23.7 23.8 10.4 10.2 10.4 8.5 8.1 9.4
18:0 4.4 4.5 20.5 22.6* 31.8 26.5* 28.7 23.9 25.6 27.9
18:1 37.5 28.6 14.9 9.7* 9.8 5.8* 14.7 18.4 9.1 9.8
18:2n-6 11.0 20.7 20.8 19.7 8.7 10.0 0.9 0.7 1.3 0.9
20:4n-6 0.6 0.6 11.1 9.2* 24.0 19.8* 15.4 15.7 14.8 12.8
22:5n-6 0.1 0.2* 0.5 0.3 2.4 2.0 1.6 0.7
18:3n-3 1.1 2.6 0.1 0.5* 0.1 0.1 0.1 0.1 0.1 0.1
20:5n-3 0.1 0.4 0.4 1.1 0.4 1.4* 0.2 0.1 0.1 0.1
22:5n-3 0.2 0.4 1.8 1.2 2.8 1.9* 0.7 0.5 1.6 1.0
22:6n-3 0.1 1.5 2.6 7.9* 7.3 21.1* 13.1 16.6* 31.4 32.3
Adapted from Arbuckle et al. (1993) Sows were fed diets with 4% by weight fat as vegetable oil with 35% 18:2n-6 and 5.3% 18:3n-3, or vegetable plus fish oil with 47.7% 18:2n-6, 7.2% 18:3n-3,
0.7% 20:5n-3, 0.7% 20:5n-3 and 0.2% 22:6n-3 from 5 days before parturition through lactation. The piglet tissue fatty acids were determined at 15 days of age. PE, phosphatidylethanolamine.
*P < 0.05 between groups.
Essential fatty acid metabolism during early development 251

Human milk concentrations of 22:6n-3 also show considerable variability, by over an order
of magnitude among populations and individuals. The average amount of 22:6n-3 in human
milk has been reported to be 2.8% in Zhangzi, China, 1.4% fatty acids in Canadian Inuit,
1.0% in Japan, 0.8–0.9% on the Malay Peninsula, 0.5% in Norway, 0.2–0.4% in Canada,
Europe and the USA (Kneebone et al., 1985; Innis and Kuhnlein, 1988; Innis, 1992; Chulei
et al., 1995; Helland et al., 2001), and 0.1% among women in the UK following vegan diets
(Sanders et al., 1978). While 18:2n-6 has increased, the amount of 22:6n-3 in human milk in
Western countries has decreased. Analyses conducted over the last 15 years in Australia and
Canada show human milk 22:6n-3 has declined by 25–50% (Makrides et al., 1995b; Innis,
2002). Supplementation of lactating women with fish or fish oil increases the secretion of
20:5n-3 and 22:6n-3 into human milk (Harris et al., 1984; Henderson et al., 1992; Helland et al.,
2001). Supplementation of lactating women with 22:6n-3 from single-cell oils with >40%
22:6n-3 and <0.1% 20:5n-3, on the other hand, increases 22:6n-3 in human milk without
increasing 20:5n-3 (Makrides et al., 1996; Gibson et al., 1997; Fidler et al., 2000). Higher
amounts of 22:6n-3 in milk result in higher intakes and higher blood lipid levels of 22:6n-3
in breast-fed infants (Henderson et al., 1992; Gibson et al., 1997; Innis and King, 1999;
Jensen et al., 2000). The amount of 20:4n-6 in milk appears to be much more tightly regu-
lated, and does not seem to be reduced even with high intakes of 20:5n-3 and 22:6n-3 (Innis
and Kuhnlein, 1988; Makrides et al., 1996; Fidler et al., 2000; Jensen et al., 2000).
Recent studies using stable isotope tracer methodologies have estimated that 20–25% of
22:6n-3, 33% of 18:2n-6 and 12% of 20:4n-6 secreted in human milk are derived from the dietary
intake of the previous 48 hours (del Prado et al., 2000; Fidler et al., 2000). This may suggest that
adipose tissue provides a significant portion of the fatty acids secreted in milk. Consequently, the
essential fatty acid content of the diet fed during gestation may also be important to the essential
fatty acid quality of the milk in later lactation. The importance of the supply of essential fatty
acids in milk or milk substitutes to infant growth and development is discussed in section 8.

6. ESSENTIAL FATTY ACID ACCRETION IN THE BRAIN AND RETINA


Arachidonic acid (20:4n-6) and 22:6n-3 are accumulated in large amounts in the brain and
retina during brain growth, particularly the brain growth spurt when the relative increase in
brain weight is at its highest (Dobbing and Sands, 1979). Animals that are born with a rela-
tively mature brain, such as the monkey and guinea pig, known as precocial animals, have the
highest requirement for 20:4n-6 and 22:6n-3 for brain growth in utero. Altricial animals, such
as rats, which are born immature, have the largest period of brain growth after birth. The
increase in rat brain 20:4n-6 and 22:6n-3 is most rapid between days 10 and 15 after birth. At
birth, rat brain contains about 0.6 mg 20:4n-6 and 0.6 mg 22:6n-3. This increases to 2.46 mg
20:4n-6 and 2.48 mg 22:6n-3 by 10 days of age, then almost doubles over the next 5 days to
4.21 mg 20:4n-6 and 4.8 mg 22:6n-3 at 15 days of age, while the adult brain has 6.3 mg
20:4n-6 and 10.2 mg 22:6n-3 (Sinclair and Crawford, 1972). In the human and pig, brain
growth is intermediate about the time of birth. The human brain weighs about 100 g at the
beginning of the third trimester of gestation, 370 g at term birth, and 2000 g at 2 years of age.
The pig brain weighs about 5.8 g at 70 days of gestation, 10.7 g at 80 days, then increases
another 3-fold to 31 g at 110 days of gestation, and weighs about 46 g by 4 weeks after birth
(Pond et al., 2000). However, the rate of brain growth does not reflect the rate of maturation
within individual regions of the brain, or the sensitivity of particular developing anatomical,
biochemical or functional systems to the supply of essential fatty acids. For example, in the
human, the growth of dendritic arbors and peak formation of synapses extends from about
252 S. M. Innis

34 weeks of gestation through 24 months after birth, during which time new connections form
at a rate of up to almost 40,000 synapses/sec (Huttenlocher et al., 1982; Borgeois, 1997;
Huttenlocher and Dabholkar, 1997; Levitt, 2003).
Autopsy analyses of fetal and infant tissue from 22 weeks gestation to 17 weeks after birth
has been used to estimate the amounts of n-6 and n-3 fatty acids accumulated in developing
human tissues (Clandinin et al., 1980a,b, 1981). Estimates of intrauterine accretion of essen-
tial fatty acids indicate deposition of 552 mg/day n-6 fatty acids and 67 mg/day n-3 fatty acids
during the last trimester of gestation. Most of this is accumulated in white adipose tissue
(368 mg n-6 and 52 mg n-3 fatty acids/day). Accretion in the human fetal brain amounts to
5.8 mg n-6 and 3.1 mg n-3 fatty acids/day, representing about 1.1% and 4.65% of total body
accretion, respectively. It is not known if the brain is protected, for example at the expense of
adipose tissue during limited dietary availability of essential fatty acids.
The extent to which the developing brain and retina depend on plasma 22:6n-3, or are able
to convert 18:3n-3 to 22:6n-3, is important to understand the dietary requirements for n-3
fatty acids and the interpretation of differences in plasma 22:6n-3. Several studies have shown
that brain is able to take up and convert 14C-labelled 18:2n-6 and 18:3n-3 into longer-chain
metabolites (Sinclair and Crawford, 1972; Dhopeshwarkar and Subramanian, 1975a,b, 1976;
Cohen and Bernsohn, 1978; Anderson and Connor, 1988; Washizaki et al., 1994; Chang et al.,
1997). Conversion of 18:3n-3 and 22:5n-3 to 22:6n-3 in the retina and retinal pigment epithe-
lium has also been demonstrated (Wang and Anderson, 1993; Alvarez et al., 1994). In vitro,
brain cerebroendothelial cells take up 18:2n-6 and 18:3n-3 and can form 20:4n-6 and 22:4n-6,
and 20:5n-3 and 22:5n-3 (Moore et al., 1991; Moore, 1994) as well as 22:6n-3 in a pathway
involving carbon chain 24 intermediates (Delton-Vandenbroucke et al., 1997). However, cerebral
and cerebellar neuronal cells do not form 22:6n-3 (Moore, 2001). Astrocytes from suckling
rat brain, however, can form 22:6n-3, which after release can be taken up by neurons (Moore
et al., 1991; Moore, 2001). Although the pathway for 22:6n-3 formation is present, the major
product of 18:3n-3 metabolism in neonatal brain astrocytes is 22:5n-3, not 22:6n-3 (Innis and
Dyer, 2002, Williard et al., 2002); in vivo, brain has only minor amounts of 22:5n-3 (Sastry,
1985). In the absence of n-3 fatty acids, neonatal brain astrocytes cultured with 18:2n-6 accu-
mulate 20:4n-6 and 22:4n-6, rather than 22:5n-6 as occurs in n-3 fatty acid- deficient animals
(Innis and Dyer, 2002). These findings indicate that while astrocytes have the ability to form
22:6n-3 uptake of 22:6n-3 from plasma derived from synthesis in the liver, or from placental
transfer before birth or provided preformed in the diet after birth is likely to be quantitatively
more important for brain 22:6n-3 accretion.
The pathways of transfer of 22:6n-3 from plasma to the developing brain are still incom-
pletely understood. Several studies have shown that albumin-bound (unesterified) 20:4n-6
and 22:6n-3 is taken up by the brain (Washizaki et al., 1994; Jones et al., 1997). Other stud-
ies have shown that 18:2n-6 and 20:4n-6, but not 16:0, from physiological amounts of
radiolabelled 2-arachidonyl-lysophosphatidylcholine, 2-palmitoyl-lysophosphatidylcholine,
and 2-linoleoyl-lysophosphatidylcholine, rapidly appear in the brain, and that uptake of
18:2n-6, 20:4n-6 and 22:6n-3 occurs more readily from lysophospholipids than from unes-
terified fatty acids (Thies et al., 1992, 1994; Bernoud et al., 1999; Lagarde et al., 2001).
Lysophospholipids represent a significant (5–20%) proportion of total phospholipids in
mammalian plasma, and thus could play an important role in the transfer of 20:4n-6 and
22:6n-3 to extrahepatic tissues. Other recent studies have shown that HDL PE may deliver
long-chain polyunsaturated fatty acids to the brain via the sequential methylation of PE to PC
at the blood–brain barrier (Magret et al., 1996). Further elucidation of the role of different
plasma lipids in transporting essential fatty acids to the developing brain is important in order
Essential fatty acid metabolism during early development 253

to better identify the dietary and metabolic conditions that place developing infants at risk of
delayed neural and retinal development.

7. ESSENTIAL FATTY ACID METABOLISM IN DEVELOPMENT


Prior to birth and during suckling, all of the n-6 and n-3 fatty acids accumulated by the
developing foetus and neonate must be derived by transfer from the mother, either by pla-
cental transfer or from milk. The extent to which the foetus and young animal are able to use
18:2n-6 and 18:3n-3 or depend on a supply of preformed 20:4n-6 and 22:6n-3 is important in
understanding dietary essential fatty acid requirements during development. Both Δ6- and
Δ5-desaturase activities are present in liver and brain from early in development. Desaturase
activities are relatively low in fetal rat liver and higher in fetal brain (Sanders and Rana,
1987). In mouse, brain Δ6-desaturase activity decreased about 4-fold from 3 days before birth
to weaning, and activity in liver increased about 7-fold until 11 days after birth (Bourre and
Piciotti, 1992). Linolenic acid (18:3n-3) deficiency increased, and feeding with 20:5n-3 +
22:6n-3 decreased, activity of the Δ6-desaturase (Dinh et al., 1993). Similarly, pig liver and
brain at 63 days of gestation, 3 days postpartum (term birth, about 115 days) and at 12 weeks
of age are able to convert [1-14C]18:2n-6 and 20:3n-6 to tetraenes (20:4n-6, 22:4n-6) and
pentaenes (22:5n-6), but the rate of conversion in the fetal piglet liver and brain is 3–5-fold
lower than in the animals after birth, and also much lower in brain than liver (Clandinin et al.,
1985a,b). Li et al. (2000) found no synthesis of 22:6n-3 from [U-13C]18:3n-3 in liver of foetal
piglets at 70–72 or 110–112 days gestation. Instead, desaturation of 18:3n-3 was limited at
20:5n-3, suggesting that the final steps of elongation and peroxisomal chain shortening may
be limiting. Biosynthesis of 22:6n-3, however, increased rapidly over the first 14 days after
birth (Li et al., 2000). This is compatible with placental supply of 22:6n-3 before birth, and
the low amount of 22:6n-3 in sow milk. Stable isotope tracer methodology has been used
to show that foetal baboons can form 18:3n-3 from a dose of intravenous [U-13C]18:3n-3
(Su et al., 2001). However, only about 0.6% of the 18:3n-3 administered was recovered in brain
22:6n-3. Again, this is consistent with low desaturation of n-3 fatty acids during prenatal
development when placental transfer facilitates high concentrations of 22:6n-3 in foetal
plasma. Activity of Δ6- and Δ5-desaturase has also been shown in human foetal liver micro-
somes from as early as 17 weeks of gestation (Chambaz et al., 1985; Poisson et al., 1993). The
activity of the pathway to 22:6n-3 in human foetal liver prior to birth, however, is not known.
Studies over a decade ago established that blood lipid 20:4n-6 and 22:6n-3 are lower in
infants fed formula with 18:2n-6 and 18:3n-3 as the only n-6 and n-3 fatty acids than in
breast-fed infants (Putnam et al., 1982; Ponder et al., 1992). Higher amounts of 18:3n-3 in the
diet do not result in higher plasma or red blood cell 22:6n-3 in infants (Ponder et al., 1992).
Studies in this field interpreted these findings as evidence of “immature” activity of a putative
Δ4-desaturase believed to be responsible for synthesis of 22:6n-3 from 22:5n-3 (Putnam et al.,
1982; Carlson et al., 1993). However, 20:4n-6, which is formed by Δ6- and Δ5-desaturation
of 18:2n-6, is also lower in infants fed formula than in breast-fed infants. Advances in stable
isotope tracer technologies have now allowed the demonstration that conversion of isotopi-
cally labelled 18:3n-3 to 22:6n-3 and of 18:2n-6 to 20:4n-6 is as high or higher in preterm as
in term infants, and higher in infants not receiving preformed 20:4n-6 and 22:6n-3 from
human milk (Carnielli et al., 1996; Salem et al., 1996; Sauerwald et al., 1997; Uauy et al.,
2000; Pawlosky et al., 2001). The conversion of 18:3n-3 to 22:6n-3, however, appears to be
highly variable among individuals and could be as low as <1 to 4% 18:3n-3 converted to
22:6n-3 in humans (Emken et al., 1994; Pawlosky et al., 2001). These tracer methodologies
254 S. M. Innis

Table 6
Cerebral cortex long-chain polyunsaturated fatty acids in human infants fed human milk
or infant formula

PE PS

Formula 18:3n-3 Formula 18:3n-3

Fatty acid Human milk 1.5% 0.4% Human milk 1.5% 0.4%

20:4n-6 17.6 20.1* 19.6∗∗ 7.9 9.0 9.0


22:4n-6 12.0 12.6 14.3* 7.9 8.5 9.3
22:5n-6 3.2 4.8* 7.0* 5.3 7.7* 10.4**
22:6n-3 17.7 13.4* 11.6* 23.5 19.3* 14.4*
22 :5n-6/22:6n-3 0.18 0.36 0.60 0.22 0.40 0.72
*P <0.01, **P < 0.5
compared to human milk. The formula contained 1.5% 18:3n-3 and 16.0% 18:2n-6 or 0.4%
18:3n-3 and 14.5% 18:2n-6. Adapted from Farquharson et al. (1995). PE, phosphatidylethanolamine; PS,
phosphatidylserine.

have not yet been able to provide quantitative estimates of the contribution of 22:6n-3 syn-
thesis to accumulation at the level of the tissues. Data from autopsy tissue also provides
information on essential fatty acid metabolism in human infants. Infants fed formula con-
taining 0.4% or 1.5% 18:3n-3 but no long-chain polyunsaturated fatty acids had lower brain
22:6n-3 and higher 22:4n-6 and 22:5n-6 in cerebral cortex PE and PS than infants who were
breast-fed (Farquharson et al., 1995). Brain 22:5n-6 was also higher and the 22:6n-3 to 22:5n-3
ratio lower in infants fed the formula with 0.4% 18:3n-3 rather than 1.5% 18:3n-3 (table 6).
The increase in 22:5n-6 is consistent with deficiency of 18:3n-3, and also provides evidence
that the final steps of elongation, Δ6-desaturation and peroxisomal partial chain shortening of
essential fatty acids are active, even in young infants. Stable isotope studies have also shown
that 22:6n-3 synthesis is higher in infants with higher intakes of 18:3n-3 (Sauerwald et al.,
1996). The findings, however, do not address whether a higher intake of 18:3n-3 can support
sufficient endogenous synthesis of 22:6n-3 to meet the needs of the developing brain.

8. LONG-CHAIN FATTY ACIDS IN HUMAN INFANT NUTRITION


Several studies have reported that groups of infants who are breast-fed perform better on tests
of neurodevelopment than bottle-fed infants (Anderson et al., 1999). This advantage appears
to remain even when many factors, such as socio-demographic variables, maternal eduction
and birth order, are controlled, although further work still needs to be done with respect to
clearly identifying and controlling extent and duration of breast-feeding in these comparative
studies (Drane and Logemann, 2000). Although milk contains numerous biologically active
components not present in infant formulas, the presence of 22:6n-3 in human milk and the
critical role of this fatty acid in normal retina and brain development has led to intense inves-
tigation of the essential fatty acid needs of the human infant for growth and development. The
central question is whether the rate of conversion of 18:3n-3 to 22:6n-3 in human infants is
sufficient to provide enough 22:6n-3 for optimal brain and retinal function. These studies are
complicated because the relationship between plasma or red blood cell and brain 22:6n-3 is
curvilinear, rather than linear (Arbuckle et al., 1991; Rioux et al., 1997; Ward et al., 1998).
At intakes above requirement, progressive increases in plasma and red blood cell 22:6n-3
with increasng dietary intake are not accompanied by similar increases in 22:6n-3 in the brain
Essential fatty acid metabolism during early development 255

and retina. As might be expected, increasing the dietary intake of 20:4n-6 and 22:6n-3 in young
infants results in an increase in these fatty acids in blood lipids, with the increase dependent
on the amount of 20:4n-6 and 22:6n-3 fed, and also the concurrent intake of 18:2n-6 (Carlson
et al., 1993b; Makrides et al., 1995a; Innis et al., 1996). Because of this, the assessment of
dietary n-3 fatty acid requirements requires an approach combining measurement of growth,
functions dependent on neural or other tissue 22:6n-3, and blood lipid 22:6n-3.
Premature infants are considered particularly vulnerable to nutritional deficiency because
of their limited adipose tissue mass and immaturity in many metabolic and physiological
pathways at birth. Following birth, there is a rapid decline in the high 20:4n-6 and 22:6n-3
characterisitic of foetal plasma, and a large increase in 18:2n-6 (Innis, 1991). The increase in
brain weight relative to body weight is also more rapid during the last trimester of gestation
than at later ages. Numerous studies have documented an increased incidence of develop-
mental problems in prematurely born children at school age (Bhutta et al., 2002), which is
likely to reflect in part failure to provide optimal essential nutrients to support third trimester
brain growth and development ex utero.
Several options are available for including 20:4n-6 and 22:6n-3 in infant formulas. For
20:4n-6, these include egg total lipids, egg-derived triglycerides and egg phospholipids, or
triglycerides from the single-cell fungus Mortierella alpina (which are highly enriched in
20:4n-6). Sources of 22:6n-3 include fish oils, either with high 20:5n-3 and 22:6n-3, fish oils
such as tuna oils specifically selected to be low in 20:5n-3 and high in 22:6n-3, egg lipids, or
oils from the single-cell micro-algae Crypthecodinium cohnii which contains >40% 22:6n-3
and virtually no 20:5n-3. Early studies by Birch et al. reported a higher rod threshold and
lower maximal amplitude values in the B wave in ERG recordings of infants fed formula
containing corn oil with only 0.5% 18:3n-3 (about 0.25% dietary energy) (Birch et al., 1992;
Uauy et al., 1992). The higher threshold and lower maximum amplitude suggest that greater
luminescence was needed to elicit a response, and that signal transduction was reduced,
respectively. Subsequent studies found that supplementation of preterm infant formula con-
taining 1.2% energy as 18:3n-3 (about 2.4% fatty acids) with 0.12% or more energy as
22:6n-3 and 0.23% or more energy as 20:4n-6 resulted in higher visual acuity when measured
either by VEP techniques or with behavioural measures based on the ability of the infant to
demonstrate a looking response to black and white gratings (stripes) of varying size (Carlson
et al., 1993b, 1996a; Faldella et al., 1996; San Giovanni et al., 2000; O’Connor et al., 2001).
Recently, a large multicenter trial found an advantage in Bayley mental developmental
inventories and the MacArthur communicative inventories in preterm infants of birth weight
less than 1250 g fed formula supplemented with 20:4n-6 and 22:6n-3 (O’Connor et al., 2001).
The long-term significance of these early changes in visual and neural development is not yet
known. The findings, however, show that 1.2% dietary energy as 18:3n-3 does not meet the
n-3 fatty acid requirements of preterm infants, and that visual and some aspects of neural
development are increased by small dietary intakes of 22:6n-3. Several studies, however, have
found evidence of reduced growth in preterm infants fed formulas containing 22:6n-3 from
fish oils (Carlson et al., 1992, 1996a; Ryan et al., 1999), and a positive relation between
20:4n-6 and growth has been described (Carlson et al., 1993a). Recent clinical studies have
also provided evidence of higher growth in preterm infants fed formulas supplemented with
0.27% energy 20:4n-6 and 0.14% 22:6n-3 from single-cell triglycerides (Innis et al., 2002).
While it is not clear if this is related to the role of 20:4n-6 in regulating early aspects of adipo-
genesis (Reginato et al., 1998), bone growth (Weiler and Fitzpatrick-Wong, 2002) or other
mechanisms, it is evident that both the n-6 and the n-3 long-chain polyunsaturated fatty acids
play an important role in the growth and development of preterm infants. Following these
256 S. M. Innis

studies, preterm infant formulas containing both 20:4n-6 and 22:6n-3 are now available in many
countries. Some evidence also suggests that inclusion of 20:4n-6 and 22:6n-3 in formula fed to
preterm infants may increase immune system development, as assessed by ex vivo mitogen
stimulation of peripheral blood lymphocytes (Field et al., 2001). Much more work on the role
of essential fatty acids and their role in the developing immune and other systems is needed.
Although many studies have addressed the role of dietary 22:6n-3 and 20:4n-6 in the
growth and development of term infants, current findings are not consistent. Several early
studies reported evidence of higher looking acuity and higher VEP acuity in term gestation
infants fed formula supplemented with 22:6n-3 in amounts similar to that in human milk
(Makrides et al., 1995c; Carlson et al., 1996b). Recent large multicenter trials, however, have
not found a benefit of adding 0.12% to up to 0.36% 22:6n-3 (about 0.06–0.18% energy) of
the formula fat on looking acuity, VEP acuity, or scores on tests of mental and motor skill
development (Auestad et al., 1997, 2001; Lucas et al., 1999; Makrides et al., 2000). Birch et al.
(1998, 2000), however, found higher VEP acuity during the the first year of life, and higher
scores on the Bayley scales of infant development at 18 months of age among term infants
fed formula containing a fat blend with 0.36% 22:6n-3 compared with no 22:6n-3 and 1.49%
18:3n-3. VEP acuity was also higher in 12-month-old infants who were weaned from breast-
feeding to formula with 22:6n-3 rather than unsupplemented formula at about 5 months of
age (Birch et al., 2002). Higher scores on a three-step problem-solving task have also been
reported for term infants fed a formula containing fat with 0.15–0.25% 22:6n-3 and 0.7%
18:3n-3 (Willatts et al., 1998). No evidence of altered growth was found in any of the latter
studies with term infants fed formulas containing 22:6n-3 and 20:4n-6. The discrepancy
among studies with term infants could involve the low 18:3n-3 content and thus dependence
on 22:6n-3 in some of the formulas tested, addition of inadequate amounts of 22:6n-3 to see
a positive effect, differences in the formula balance of 20:4n-6 to 22:6n-3, variability in the
22:6n-3 status of the infants at birth, age of the infants at testing, and lack of test sensititivity
to detect biologically important differences in development.
Recent studies have also addressed whether the variability in 22:6n-3 in human milk, and
as a result in the diet and blood lipids of the breast-fed infant, is of physiological significance
to infant development. Infants in the lowest tertile of RBC PE 22:6n-3 who received milk with
0.17% fat as 22:6n-3 had significantly lower visual acuity at 2 and 12 months of age than
infants in the highest tertile of RBC PE 22:6n-3 who received milk with 0.31% 22:6n-3 (Innis
et al., 2001; fig. 5). No relation was found between the infants’ 20:4n-6 or 22:6n-3 status and
scores on the Bayley II mental or motor developmental indices, novelty preference assessed
using the Fagan test, or on a standardized object search task (Piaget’s A not B). However, the
infants’ 22:6n-3 status at 2 months of age was significantly related to the ability to discrimi-
nate a non-native (Hindi) retroflex and dental phonetic contrast at 9 months of age, and to
language production and comprehension assessed with the CDI at 14 and 17 months of age,
after adjusting for confounding variables (Innis et al., 2001; Innis, 2002). A significant asso-
ciation between sweep VEP acuity and human milk 22:6n-3 was also recently reported in a
cross-sectional study of breast-fed infants (Jorgensen et al., 2001). These associations
between 22:6n-3 and visual and neural development in breast-fed infants, while consistent
with the essential role of 22:6n-3 in retina and brain function, cannot be interpreted as a
demonstration of causality. This requires dietary intervention that modifies the maternal
intake of 22:6n-3 but not other nutrients. However, newer research to show dependence of the
fetal and infant 22:6n-3 on the maternal intake of 22:6n-3 raises important questions about
the n-3 fatty acid requirements of pregnant and lactating women with respect to supporting
optimal visual and neural development in the infant.
Essential fatty acid metabolism during early development 257

Fig. 5. Relation of milk and infant red blood cell phosphatidylethanolamine (PE) 22:6n-3 at 2 months of
age to visual acuity at 2 and 12 months of age in healthy term gestation infants. The bars represent mean visual
acuity ± SD; * significantly different from infants in the lowest tertile of 22:6n-3. Adapted from Innis et al. (2001).

9. DIETARY ESSENTIAL FATTY ACID REQUIREMENTS


The physiological effects of essential fatty acid deficiency have been well described, partic-
ularly in rats fed a fat-free diet and replenished with varying amounts of n-6 fatty acids
(Mohrhauer and Holman, 1963; Holman, 1991; Innis, 1992; table 1). Studies in rodents have
shown that a dietary intake of as low as 1–2% energy 18:2n-6 is sufficient to avoid any
adverse effects on reproduction, gestation, perinatal mortality or growth. Careful dose-
response studies have shown that the minimum intake of 18:2n-6 required to sustain a
constant level of 20:4n-6 in all organs (including brain, liver, lung, heart, kidney, muscle and
adipose) is 2.4% of dietary energy (Bourre et al., 1990). A similar effect, with maintenance
of normal growth, is achieved with a much smaller amount of 20:4n-6, of about 0.2% dietary
energy (Mohrhauer and Holman, 1963). Similar studies on the requirement for n-3 fatty acids
have shown that a dietary intake of 0.26% energy 18:3n-3 in adults and 0.4% energy 18:3n-3
in developing animals maintains 22:6n-3 in brain, liver, heart and other organs (Bourre et al.,
1989a, 1993). There is no evidence that rodents require a dietary source of 22:6n-3 if fed a
diet containing sufficient 18:3n-3. The requirement for n-3 fatty acids, however, can be met
by small amounts of 22:6n-3 rather than conversion from 18:3n-3.
Studies in young piglets have shown that an intake of <0.7% energy from 18:3n-3 is inad-
equate to support accretion of 22:6n-3 in neural tissues (Hrboticky et al., 1989, 1990, 1991).
The accretion of 22:6n-3 was reduced when the diet of the neonatal piglet contained a ratio
of 18:2n-6 to 18:3n-3 of 16:1 or higher, rather than 8:1 or lower (Arbuckle et al., 1992, 1994).
Further understanding of the importance of the concurrent intake of 18:2n-6 to tissue accre-
tion of 20:4n-6 and 22:6n-3 is needed. Higher intakes of 2% energy 18:3n-3 resulted in higher
22:6n-3 in the developing piglet brain (Arbuckle et al., 1992, 1994), compatible with studies
showing postnatal desaturation and elongation of 18:3n-3 to 22:6n-3 in this species (Li et al.,
2000). However, studies in young piglets also provide ample evidence that 22:6n-3, either in
sows’ milk or in a milk replacer diet, is much more efficacious than 18:3n-3 in increasing
22:6n-3 in blood, liver, and brain and retina 22:6n-3 (Arbuckle et al., 1991; Arbuckle and
Innis 1992, 1993; de la Presa Owens and Innis, 1999a,b). A dietary intake of as little as 0.15%
energy as 22:6n-3 supports similar or higher brain and retina 22:6n-3 than a diet with 2%
energy as 18:3n-3.
258 S. M. Innis

Stable isotope tracer studies in pregnant and fetal baboons have also provided convincing
evidence that conversion of 18:3n-3 to 22:6n-3 is much less efficient in supplying 22:6n-3 for
incorporation into developing brain and other organs than preformed 22:6n-3 (Greiner et al.,
1997; Su et al., 1999, 2001). The available data suggest that 90% and perhaps more of dietary
18:3n-3 undergoes β-oxidation, with the acetyl-CoA either further oxidized to generate ATP,
or recycled for de novo synthesis of other lipids. Although the increase in brain and retina
22:6n-3 with increasing intakes of 22:6n-3 is much lower than the increase in 22:6n-3 in liver
or plasma, high intakes of 20:5n-3 and 22:6n-3 have been shown to decrease visual function
in guinea pigs, and to decrease auditory brainstem evoked potentials in neonatal rats
(Weisinger et al., 1995; Stockard et al., 2000). Although the mechanism of these effects is not
yet known, these findings suggest the need for dose-response studies to determine the safe
upper limit of 22:6n-3 intake in developing animals.
Studies in the 1960s identified skin lesions and growth failure among human infants fed
skimmed cows’ milk, which provides minimal amounts of n-6 fatty acids (Hansen et al., 1958).
Clinical and biochemical signs of n-6 fatty acid deficiency, many of which are related to the role
of n-6 fatty acids in the skin, are not seen in human infants fed diets providing 18:2n-6 in amounts
above recommended intakes of 3–4% energy as 18:2n-6 (FAO/WHO, 1993). Expert groups have
suggested an acceptable range of intake of 18:2n-6 for preterm and term human infants of
3.2–12.8% energy (LSRO, 1998, 2001). The absence of skin signs, normal growth, and a
triene/tetraene ratio of <0.2 in plasma lipids, however, ensures neither optimal 20:4n-6 nor
balance between n-6 and n-3 fatty acids in developing tissues. Definitive information on n-6 fatty
acid requirements that are based on functional endpoint indicators related to 20:4n-6 metabolism
are largely lacking. Similarly, the n-3 fatty acid requirements of human infants are still unclear.
The suggested acceptable range of 18:3n-3 intake for preterm and term infants is 0.7–2.1% of
dietary energy (LSRO, 1998, 2001). The requirement for 18:3n-3, however, depends on the
amount of 22:6n-3 provided to the infant. Recent clinical studies in premature infants suggest that
in the absence of a dietary intake of 22:6n-3, an intake of 1.2% energy as 18:3n-3 does not meet
the needs of the developing brain and retina (O’Connor et al., 2001).
Another approach is to estimate a dietary intake likely to meet the needs for accretion,
based on knowledge gained from autopsy tissue analyses. These analyses suggest that about
67 mg n-3 fatty acids (mostly 22:6n-3) and 552 mg n-6 fatty acids are accumulated per day
in fetal tissue during the last trimester of gestation (Clandinin et al., 1981). Thus, an infant
fed human milk or a milk substitute that provides the only dietary source of polyunsaturated
fatty acids would need to recieve about 0.23% fat as 22:6n-3 and 2% fat as n-6 fatty acids,
assuming an intake of 780 ml/day. Definitive data on the amount of dietary 18:3n-3 converted
to 22:6n-3 during development are not available. Assuming this to be 10%, then the infant diet
would need to contain at least 2.3% 18:3n-3 in order to meet the needs for n-3 fatty acids.
Some expert groups have suggested a dietary intake of essential polyunsaturated fatty acids
based on the amounts present in human milk. This is problematic because the essential fatty
acid content of milk is greatly influenced by the amounts of n-6 and n-3 fatty acids in the
maternal diet. The competition between 18:2n-6 and 18:3n-3 for desaturation has also led to
recommendations for the n-6 to n-3 fatty acid in infant diets that are based on the composition of
human milk. The total n-6 to total n-3 fatty acid ratio of human milk is generally in the range of
4:1 to 10:1 (Neuringer and Connor, 1986), but this ratio does not consider the differences in tissue
handling and biological activity of 18:2n-6 and 20:4n-6, and 18:3n-3, 20:5n-3 and 22:6n-3.
Dietary 22:6n-3 results in increased blood lipid 22:6n-3 in human infants, while the increasing
dietary intake of 18:3n-3 has little effect. For example, the plasma phospholipids of infants fed
formula with about 0.12% or 0% energy 22:6n-3 had 5.2 ± 0.2% and 2.0 ± 0.1% 22:2n-3,
Essential fatty acid metabolism during early development 259

respectively, but infants fed formula with 0.4% or 2.4% 18:3n-3 and 0% 22:6n-3 had 2.3 ± 0.2%
and 2.2 ± 0.3% 22:6n-3, respectively (Ponder et al., 1992; Innis et al., 1996). The increase in
red blood cell and plasma phospholipid 22:6n-3 was lower in infants fed formula with 32%
rather than 20% 18:2n-6 (Innis et al., 1996). The Food and Agriculture/World Health
Organization recommend that term infants receive, per kg body weight, 600 mg 18:2n-6, 50 mg
18:3n-3, 40 mg 20:4n-6 and 20 mg 22:6n-3 per day, based on the amounts in human milk
(FAO/WHO, 1993). Similarly, the Institute of Medicine of the National Academies of Medicine
provides an adequate intake (AI) for infants of 0–12 months of 4.4–4.6 g/day of n-6 fatty acids
and 0.5 g/day n-3 fatty acids, based on the intake of breast-fed infants (IOM, 2002). However,
there is no evidence from clinical studies to indicate that these amounts exceed or meet the
needs to maximize potential neural development in young infants (FAO/WHO, 1993).

10. CONCLUSIONS AND FUTURE PERSPECTIVES


It is clear that essential fatty acid deficiency can lead to profound problems in growth, and
functional disturbances in many organs including the developing central nervous system.
Large gaps still exist in understanding the requirements, metabolism and functions of essential
fatty acids, particularly the long-chain polyunsaturated fatty acids 20:4n-6 and 22:6n-3,
during development. In particular, pathways and regulatory mechanisms involved in the trans-
fer of 20:4n-6 and 22:6n-3 across the placenta and in milk, and the effect of the concurrent
intake of 18:2n-6 in modulating fatty acid accretion in developing tissues, is incompletely
understood. Likewise, much is yet to be learned regarding the role of n-6 and n-3 fatty acids
in the molecular, biochemical and histological development of the central nervous system,
and their relation to cognitive and behavioural impairments in developing animals and
infants. Little information is available on the role of polyunsaturated fatty acids in the molec-
ular and functional development of the immune system, intestine, bone, adipose tissue
and many other organs; these areas afford considerable oportunities for new research. Future
expert groups seeking to establish polyunsaturated fatty acid requirements in development
will benefit from a greater understanding of species differences in essential fatty acid metabo-
lism, and dose-response studies to elucidate the safe and adequate range and tolerable
upper limit of intake of individual n-6 and n-3 fatty acids, and their appropriate balance in
the diet.

REFERENCES
Abayasekara, D.R., Wathes, D.C., 1999. Effects of altering dietary fatty acid composition on prostaglandin
synthesis and fertility. Prostagland. Leuk. Essent. Fatty Acids 61, 275–287.
Adams, P.B., Lawson, S., Sanigorski, A., Sinclair, A.J., 1995. Arachidonic to eicosapentaenoic acid ratio
in blood correlates positively with clinical symptoms of depression. Lipids 31, S157–S161.
Ahmad, A., Moriguchi, T., Salem, N., 2002a. Decrease in neuron size in docosahexaenoic acid-deficient
brain. Pediatr. Neurol. 26, 210–218.
Ahmad, A., Murthy, M., Greiner, R.S., Moriguchi, T., Salem, N. Jr., 2002b. A decrease in cell size
accompanies loss of docosahexaenoate in rat hippocampus. Nutr. Neurosci. 5, 103–113.
Albert, C.M., Hennekens, C.H., O’Donnell, C.J., Ajani, U.A., Carey, V.J., Willett, W.C., Ruskin, J.N.,
Manson, J.E., 1998. Fish consumption and risk of sudden cardiac death. J. Amer. Med. Assoc. 279, 23–28.
Alvarez, R.A., Aguire, G.D., Acland, G.M., Anderson, R.E., 1994. Docosapentaenoic acid is converted
to docosahexaenoic acid in the retinas of normal and prcd-affected miniature poodle dogs. Invest.
Ophthalmol. Vis. Sci. 35, 402–408.
Anderson, G.J., Connor, W.E., 1988. Uptake of fatty acids by the developing rat brain. Lipids 23,
286–290.
260 S. M. Innis

Anderson, J.W., Johnstone, B.M., Remley, D.T., 1999. Breast-feeding and cognitive development:
a meta-analysis. Amer. J. Clin. Nutr. 70, 525–535.
Anderson, R.E., Landis, D.J., Dudley, P.A., 1976. Essential fatty acid deficiency and renewal of rod outer
segments in albino rat. Invest. Ophthalmol. 15, 232–236.
Antal, A., Keri, S., Bodis-Wollner, J., 1997. Dopamine D2 receptor blockade alters the primary and cog-
nitive components of visual and evoked potentials in the monkey, Macaca fasciliaris. Neurosci. Lett.
232, 179–181.
Arbuckle, L.D., Innis, S.M., 1992. Docosahexaenoic acid in developing brain and retina of piglets fed
high or low α-linolenate formula with and without fish oil. Lipids 27, 89–93.
Arbuckle, L.D., Innis, S.M., 1993. Docosahexaenoic acid is transferred through maternal diet to milk and
to tissues of natural milk-fed piglets. J. Nutr. 123, 1668–1675.
Arbuckle, L.D., MacKinnon, M.J., Innis, S.M., 1994. Formula 18:2(n-6) and 18:3(n-3) content and ratio
influence long-chain polyunsaturated fatty acids in the developing piglet liver and central nervous
system. J. Nutr. 124, 289–298.
Arbuckle, L.D., Rioux, F.M., Mackinnon, M.J., Hrboticky, N., Innis, S.M., 1991. Response of (n-3) and
(n-6) fatty acids in brain, liver and plasma of piglets fed formula to increasing, but low, levels of fish
supplementation. J. Nutr. 121, 1536–1547.
Arbuckle, L.D., Rioux, F.M., MacKinnon, M.J., Innis, S.M., 1992. Formula α-linoleic (18:3(n-6) and
linoleic (18:2(n-6)) acid influence neonatal piglet liver and brain saturated fatty acids, as well as
docosahexaenoic acid (22:6(n-3)). Biochim. Biophys. Acta 1125, 262–267.
Auestad, N., Halter, R., Hall, R., Blatter, M., Bogle, M.L., Burks, W., Erickson, J.R., Fitzgerald, K.M.,
Dobson, V., Innis, S.M., Singer, L.T., Montalto, M.G., Jacobs, J.R., Qiu, W., Bornstein, M.H., 2001.
Growth and development in term infants fed long-chain polyunsaturated fatty acids, a double-masked,
randomized, parallel, prospective, multivariate study. J. Pediatr. 108, 372–381.
Auestad, N., Montalto, M.B., Hall, R.T., Fitzgerald, K.M., Wheeler, R.E., Connor, W.E., Neuringer, M.,
Connor, S.L., Taylor, J.A., Hartmann, E.E., 1997. Visual acuity, erythrocyte fatty acid composition,
and growth in term infants fed formulas with long chain polyunsaturated fatty acids for one year.
Pediatr. Res. 41, 1–10.
Aveldano, M.I., 1987. A novel group of very long chain polyenoic fatty acids in dipolyunsaturated phos-
phatidylcholines from vertebrate retina. J. Biol. Chem. 262, 1172–1179.
Aveldano, M.I., Sprecher, H., 1987. Very long chain (C24–C36) polyenoic fatty acids of the n-3 and
n-6 series in dipolyunsaturated phosphatidylcholines from bovine retina. J. Biol. Chem. 262,
1180–1186.
Aveldano, M.I., Pasquare de Garcia, S.J., Bazan, N.G., 1983. Biosynthesis of molecular species of inos-
itol, choline, serine, and ethanolamine glycerophospholipids in the bovine retina. J. Lipid Res. 24,
628–639.
Basmak, H., Yildirim, N., Erdinc, O., Ywdakul, S., Ozdemir, G., 1999. Effect of levodopa therapy on
visual evoked potentials and visual acuity in amblyopia. Ophthalmologica 213, 110–113.
Bayourthe, C., Enjalbert, F., Moncoulon, R., 2000. Effects of differing forms of canola oil fatty acids plus
canola meal on milk composition and physical properties of butter. J. Dairy Sci. 83, 690–695.
Bell, R.M., 1986. Protein kinase C activation by diacylglycerol second messengers. Cell 45, 631–632.
Bell, R.M., Burns, D.J., 1991. Lipid activation of protein kinase C. J. Biol. Chem. 266, 4661–4664.
Benolken, R.M., Anderson, R.E., Wheeler, I.G., 1973. Membrane fatty acids associated with the electri-
cal response in visual excitation. Science 182, 1253.
Bernoud, N., Fenart, L., Moliere, P., Dehouk, M.P., Lagarde, M., Cecchelli, R., Lecerf, J., 1999.
Preferential transfer of 2-docosahexaenoyl-1-lysophosphatidylcholine through an in vitro blood-brain
barrier over unesterified docosahexaenoic acid. J. Neurochem. 72, 338–345.
Bernsohn, J., Spitz, F.J., 1974. Linoleic and linolenic acid dependency of some brain membrane-bound
enzymes after lipid deprivation in rats. Biochem. Biophys. Res. Commun. 57, 293–298.
Bhutta, A.T., Cleves, M.A., Casey, P.H., Cradock, M.M., Anand, K.J.S., 2002. Cognitive and behavioural
outcomes of school-aged children who were born preterm. J. Amer. Med. Assoc. 288, 728–737.
Birch, D.G., Birch, E.E., Hoffman, D.R., Uauy, R.D., 1992. Retinal development in very-low-birth-weight
infants fed diets differing in omega-3 fatty acids. Invest. Ophthalmol. Vis. Sci. 33, 2365–2376.
Birch, E.E., Garfield, S., Hoffman, D.R., Uauy, R., Birch, D.G., 2000. A randomized controlled trial of
early dietary supply of long-chain polyunsaturated fatty acids and mental development in term
infants. Dev. Med. Child. Neurol. 42, 174–181.
Essential fatty acid metabolism during early development 261

Birch, E.E., Hoffman, D.R., Castaneda, Y.S., Fawcett, S.L., Birch, D., Uauy, R.D., 2002. A randomized
controlled trial of long-chain polyunsaturated fatty acid supplementation of formula in term infants
after weaning at 6 wk of age. Amer. J. Clin. Nutr. 75, 570–580.
Birch, E.E., Hoffman, D.R., Uauy, R., Birch, D.G., Prestidge, C., 1998. Visual acuity and the essential-
ity of docosahexaenoic acid and arachidonic acid in the diet of term infants. Pediatr. Res. 44,
201–209.
Bonin, A., Khan, N.A., 2000. Regulation of calcium signaling by docosahexaenoic acid in human T-cells:
implication of CRAC channels. J. Lipid. Res. 41, 277–284.
Borgeois, J.P., 1997. Synaptogenesis, heterochrony and epigenesis in the mammalian cerebral cortex.
Acta Paediatr. Suppl. 422, 27–33.
Borghese, C.M., Gomez, R.A., Ramirez, O.A., 1993. Phosphatidylserine increases hippocampal synap-
tic efficacy. Brain. Res. Bull. 31, 697–700.
Bourre, J.M., Dumont, O., Pascal, G., Durand, G., 1993. Dietary α-linolenic acid at 1.3 g/kg maintains
maximal docosahexaenoic acid concentration in brain, heart and liver of adult rats. J. Nutr. 123,
1313–1319.
Bourre, J.M., Durand, G., Pascal, G., Youyou, A., 1989a. Brain cell and tissue recovery in rats made defi-
cient in n-3 fatty acids by alteration of dietary fat. J. Nutr. 119, 15–22.
Bourre, J.M., François, M., Youyou, A., Dumont, O., Piciotti, M., Pascal, G., Durand, G., 1989b. The
effects of dietary α-linolenic acid on the composition of nerve membranes, enzymatic activity, ampli-
tude of electrophysiological parameters, resistance to poisons and performance of learning tasks in
rats. J. Nutr. 119, 1880–1892.
Bourre, J.M., Pascal, G., Masson, M., Dumont, O., Piciotti, M., 1984. Alterations in the fatty acid
composition of rat brain cells (neurons, astrocytes and oligodendrocytes) and of subcellular
fractions (myelin and synaptosomes) induced by a diet devoid of n-3 fatty acids. J. Neurochem. 43,
342–348.
Bourre, J.M., Piciotti, M., 1992. Δ-6 desaturation of α-linolenic acid in brain and liver during develop-
ment and aging in the mouse. J. Neurosci. 6, 65–68.
Bourre, J.M., Piciotti, M., Dumont, O., Pascal, G., Durand, G., 1990. Dietary linoleic acid and polyun-
saturated fatty acid acids in rat brain and other organs: minimal requirements of linoleic acid. Lipids
25, 465–472.
Brenner, R.R., Peluffo, R.O., 1966. Effect of saturated and unsaturated fatty acids on the desaturation in
vitro of palmitic, stearic, oleic, linoleic, and linolenic acids. J. Biol. Chem. 241, 5213–5219.
Brenner, R.R., Peluffo, R.O., 1969. Regulation of unsaturated fatty acids biosynthesis. 1. Effect of unsat-
urated fatty acids of 18 carbons on the microsomal desaturation of linoleic acid into γ-linolenic acid.
Biochim. Biophys. Acta 176, 471–479.
Brenner, R.R., Peluffo, R.O., Nervi, A.M., De Tomas, M.E., 1969. Competitive effect of α- and
γ-linolenyl-CoA in linolenyl-CoA desaturation to γ-linolenyl-CoA. Biochim. Biophys. Acta 176,
420–422.
Broughton, K.S., Morgan, L.J., 1994. Frequency of (n-3) polyunsaturated fatty acid consumption induces
alterations in tissue lipid composition and eicosanoid synthesis in CD-1 mice. J. Nutr. 124,
1104–1111.
Broughton, K.S., Wade, J.W., 2002. Total fat and (n-3, n-6) fat ratios influence eicosanoid production in
mice. J. Nutr. 132, 88–94.
Brozoski, T.J., Brown, R.M., Rosvold, H.E., Goldman, P.S. 1979. Cognitive deficit caused by regional
depletion of dopamine in prefrontal cortex of rhesus monkey. Science 205, 929–932.
Calder, P.C., 1998. Dietary fatty acids and the immune system. Nutr. Rev. 56, 570–583.
Calder, P.C., 2001. Polyunsaturated fatty acids, inflammation, and immunity. Lipids 36, 1007–1024.
Campbell, F.M., Bush, P.G., Veerkamp, J.H., Dutta-Roy, A.K., 1998a. Detection and cellular localization
of plasma membrane-associated and cytoplasmic fatty acid-binding proteins in human placenta.
Placenta 19, 409–415.
Campbell, F.M., Gordon, M.J., Dutta-Roy, A.K., 1996. Preferential uptake of long chain polyunsaturated
fatty acids by isolated placental membranes. Mol. Cell. Biochem. 155, 77–83.
Campbell, F.M., Gordon, M.J., Dutta-Roy, A.K., 1998b. Placental membrane fatty acid-binding protein
preferentially binds arachidonic and docosahexaenoic acids. Life Sci. 63, 235–240.
Carlson, S.E., Cooke, R.J., Werkman, S.H., Tolley, E.A., 1992. First year growth of preterm infants fed
standard compared to marine-oil n-3 supplemented formula. Lipids 27, 901–907.
262 S. M. Innis

Carlson, S.E., Ford, A.J., Werkman, S.H., Peeples, J.M., Koo, W.W.K., 1996b. Visual acuity and fatty acid
status of term infants fed human milk and formulas with and without docosahexaenoate and arachi-
donate from egg yolk lecithin. Pediatr. Res. 39, 882–888.
Carlson, S.E., Werkman, S.H., Peeples, J.M., Cooke, R.J., Tolley, E.A., 1993b. Arachidonic acid status
correlates with first year growth in preterm infants. Proc. Natl. Acad. Sci. USA 90, 1073–1077.
Carlson, S.E., Werkman, S.H., Rhodes, P.G., Tolley, E.A., 1993a. Visual-acuity development in healthy
preterm infants, effect of marine-oil supplementation. Amer. J. Clin. Nutr. 58, 35–42.
Carlson, S.E., Werkman, S.H., Tolley, E.A., 1996a. Effect of long-chain n-3 fatty acid supplementation
on visual acuity and growth of preterm infants with and without bronchopulmonary dysplasia. Amer.
J. Clin. Nutr. 63, 687–697.
Carnielli, V.P., Wattimena, D.J., Luijendijk, I.H., Boerlage, A., Degenhart, H.J., Sauer, P.J., 1996. The
very low birth weight premature infant is capable of synthesizing arachidonic and docosahexaenoic
acids from linoleic and linolenic acids. Pediatr. Res. 40, 169–174.
Casamenti, F., Scali, C., Pepeu, G., 1991. Phosphatidylserine reverses the age-dependent decrease in
cortical acetylcholine release, a microdialysis study. Eur. J. Pharmacol. 194, 11–16.
Ceruku, S.R., Montegomery-Downs, H.E., Farkas, S.L., Thoman, E.B., Lammi-Keefe, C.J., 2002. Higher
maternal plasma docosahexaenoic acid during pregnancy is associated with more mature neonatal
sleep-state patterning. Amer. J. Clin. Nutr. 76, 608–613.
Chambaz, J., Ravel, D., Manier, M.C., Pepin, D., Mulliez, N., Bereziat, G., 1985. Essential fatty acids
interconversion in the human fetal liver. Biol. Neonate 47, 136–140.
Chang, M.C.J., Arai, T., Freed, L.M., Wakabayashi, S., Channing, M.A., Dunn, B.B., Der, M.G.,
Bell, J.M., Sasaki, T., Herscovitch, P., Eckelman, W.C., Rapoport, S.I., 1997. Brain incorporation of
[1-11C]-arachidonate in normocapnic and hypercapnic monkeys, measured with positron emission
tomography. Brain Res. 755, 74–83.
Charnock, J.S., Dryden, W.F., McMurchie, E.J., Abeywardena, M.Y., Russell, G.R., 1983. Differences in
the fatty acid composition of atrial and ventricular phospholipids of rat heart following standard and
lipid supplemented diets. Comp. Biochem. Physiol. 75, 47–52.
Chow, C.K. (Ed.), 2000. Fatty Acids in Foods and Their Health Implications, 2nd Edition. Marcel
Dekker, New York.
Choy, P.C., Tran, K., Hatch, G.M., Kroeger, E.A., 1997. Phospholipid metabolism in the mammalian
heart. Prog. Lipid Res. 36, 85–101.
Christensen, J.H., Gustenhoff, P., Korup, E., Aaroe, J., Toft, E., Moller, J., Rasmussen, K., Dyerberg, J.,
Schmidt, E.B., 1996. Effect of fish oil on heart rate variability in survivors of myocardial infarction,
a double blind randomised controlled trial. Brit. Med. J. 312, 677–678.
Christensen, J.H., Korup, E., Aaroe, J., Toft, E., Moller, J., Rasmussen, K., Dyerberg, J., Schmidt, E.B.,
1997. Fish consumption, n-3 fatty acids in cell membranes, and heart rate variability in survivors of
myocardial infarction with left ventricular dysfunction. Amer. J. Cardiol. 79, 1670–1673.
Chulei, R., Xiafang, I., Hongseng, M., Xiulan, M., Guizheng, I., Gianhong, D., DeFrancesco, C.A.,
Connor, W.E., 1995. Milk composition in women from five different regions of China, the great diver-
sity of milk fatty acids. J. Nutr. 125, 2993–2998.
Clandinin, M.T., Chappell, J.E., Heim, T., Swyer, P.R., Chance, G.W., 1981. Fatty acid utilization in peri-
natal de novo synthesis of tissues. Early Human Dev. 5, 355–366.
Clandinin, M.T., Chappell, J.E., Leong, S., Heim, T., Sawyer, P.R., Chance, G.W., 1980a. Intrauterine
fatty acid accretion rates in human brain, implication for fatty acid requirements. Early Human Dev.
4, 121–129.
Clandinin, M.T., Chappell, J.E., Leung, S., Heim, T., Swyer, P.R., Chance, G.W., 1980b. Extrauterine
fatty acid accretion in infant brain, implications for fatty acid requirements. Early Human Dev. 4,
131–138.
Clandinin, M.T., Wong, K., Hacker, R.R., 1985a. Synthesis of chain elongation-desaturation products
of linoleic acid by liver and brain microsomes during development of the pig. Biochem. J. 226,
305–309.
Clandinin, M.T., Wong, K., Hacker, R.R., 1985b. Δ5-desaturase activity in liver and brain microsomes
during development of the pig. Biochem. J. 227, 1021–1023.
Clarke, S.D., 2001. Polyunsaturated fatty acid regulation of gene transcription, a molecular mechanism
to improve the metabolic syndrome. J. Nutr. 131, 1129–1132.
Clarke, S.D., Armstrong, M.K., Jump, D.B., 1990. Dietary polyunsaturated fats uniquely suppress rat
liver fatty acid synthase and s14 mRNA content. J. Nutr. 120, 225–231.
Essential fatty acid metabolism during early development 263

Cohen, S.R., Bernsohn, J., 1978. The in vivo incorporation of linolenic acid into neuronal and glial cells
and myelin. J. Neurochem. 30, 661–669.
Connor, W.E., Lowensohn, R., Hacher, L., 1996. Increased docosahexaenoic acid levels in newborn
infants by administration of sardines and fish oil during pregnancy. Lipids 31, S183–S187.
Connor, W.E., Neuringer, M., Lin, D.S., 1990. Dietary effects on brain fatty acid composition, the
reversibility of n-3 fatty acid deficiency and turnover of docosahexaenoic acid in brain, erythrocytes,
and plasma of rhesus monkeys. J. Lipid Res. 31, 237–247.
Contreras, M.S., Chang, M.C., Rosenberger, T.A., Greiner, R.S., Myers, C.S., Salem, N. Jr., Rapoport, S.I.,
2001. Chronic nutritional deprivation of n-3 linolenic acid does not affect n-6 arachidonic acid recycling
with brain phospholipids in awake rats. J. Neurochem. 79, 1090–1099.
Crawford, M.A., Casperd, N.M., Sinclair, A.J., 1976. The long chain metabolites of linoleic and linolenic
acids in liver and brain in herbivores and carnivores. Comp. Biochem. Physiol. 54B, 395–401.
Crawford, M.A., Hassam, A.G., Stevens, P.A., 1981. Essential fatty acid requirements in pregnancy and
lactation with special reference to brain development. Prog. Lipid Res. 20, 31–40.
Cunnane, S.C., Menard, C.R., Likhodii, S.S., Brenna, J.T., Crawford, M.C., 1999. Carbon recycling into
de novo lipogenesis is a major pathway in neonatal metabolism of linoleate and α-linoleate.
Prostagland. Leuk. Essent. Fatty Acids 60, 387–392.
Cunnane, S.C., Williams, S.C., Bell, J.D., Brookes, S., Craig, K., Iles, R.A., Crawford, M.A., 1994.
Utilization of uniformly labeled 13C-polyunsaturated fatty acids in the synthesis of long-chain fatty
acids and cholesterol accumulating in the neonatal rat brain. J. Neurochem. 62, 2429–2436.
de la Presa Owens, S., Innis, S.M., 1999a. Docosahexaenoic and arachidonic acid reverse changes in
dopaminergic and serotoninergic neurotransmitters in piglets frontal cortex caused by a linoleic and
α-linolenic acid deficient diet. J. Nutr. 129, 2088–2093.
de la Presa Owens, S., Innis, S.M., 1999b. Diverse, region-specific effects of the addition of arachidonic
and docosahexaenoic acid to formula with low or adequate linoleic and α-linolenic acid on piglet
brain monoaminergic neurotransmitters. Pediatr. Res. 48, 125–130.
Delion, S., Chalon, S., Guilloteau, D., Besnard, J.-C., Dwand, G., 1996. α-Linolenic acid dietary defi-
ciency alters age-related changes of dopaminergic and serotoninergic neurotransmission in the rat
frontal cortex. J. Neurochem. 66, 1582–1591.
Delion, S., Chalon, S., Guilloteau, D., Lejeune, B., Besnard, J.C., Durand, G., 1997. Age-related
changes in phospholipid fatty acid composition and monoaminergic neurotransmission in the hip-
pocampus of rats fed a balanced or an n-3 polyunsaturated fatty acid-deficient diet. J. Lipid Res. 38,
680–689.
Delion, S., Chalon, S., Herault, J., Guilloteau, D., Besnard, J.C., Durand, G., 1994. Chronic dietary
α-linoleic acid deficiency alters dopaminergic and serotoninergic neurotransmitters in rats. J. Nutr.
24, 2466–2476.
Del Prado, M., Villalpando, S., Lance, A., Alfonso, E., Demmelmair, H., Koletzko, B., 2000. Contribution
of dietary and newly formed arachidonic acid to milk secretion in women on low fat diets. Adv. Exp.
Med. Biol. 478, 407–408.
Delton-Vandenbroucke, I., Gammas, P., Anderson, R.E., 1997. Polyunsaturated fatty acid metabolism in
retinal and cerebral microvascular endothelial cells. J. Lipid. Res. 38, 147–159.
Dhopeshwarkar, G.A., Subramanian, C., 1975a. Metabolism of linoleic acid in developing brain:
I. Incorporation of radioactivity from 1-14C linoleic acid into brain fatty acids. Lipids 10, 238–241.
Dhopeshwarkar, G.A., Subramanian, C., 1975b. Metabolism of 1-14C linolenic acid in developing brain:
II. Incorporation of radioactivity from 1-14C linolenate into brain lipids. Lipids 10, 242–245.
Dhopeshwarkar, G.A., Subramanian, C., 1976. Biosynthesis of polyunsaturated fatty acids in the
developing brain: I. Metabolic transformation of intracranially administered 1-14C linolenic acid.
Lipids 11, 67–71.
Dinh, T.K., Bourre, J.M., Durand, G., 1993. Effect of age and α-linolenic acid deficiency in Δ6 desaturase
activity and liver lipids in rats. Lipids 28, 517–523.
Dobbing, J., 1972. Vulnerable periods of brain development. In Eliott, K., Kights, J. (Eds.), Lipids,
Malnutrition and Developing Brain. A Ciba Foundation Symposium. Excerpta Medica, Amsterdam,
pp. 9–20.
Dobbing, J., Sands, J., 1979. Comparative aspects of the brain growth spurt. Early Human Dev. 3, 79–83.
Dobbing, J., Hopewell, J.W., Lynch, A., 1971. Vulnerability of developing brain. VII. Permanent deficit
of neurons in cerebral and cerebellar cortex following early mild undernutrition. Exp. Neurol. 32,
439–447.
264 S. M. Innis

Drane, D.L., Logemann, J.A., 2000. A critical evaluation of the evidence on the assocaition between type
of infant feeding and cognitive development. Paediatr. Perinat. Epidemiol. 14, 349–356.
Duplus, E., Glorian, M., Forest, C., 2000. Fatty acid regulation of gene transcription. J. Biol. Chem. 275,
30749–30752.
Dutta-Roy, A.K., 2000. Transport mechanisms for long-chain polyunsaturated fatty acids in the humana
placenta. Amer. J. Clin. Nutr. 71, 315S–332S.
Edwards, R., Peet, M., Shay, J., Horrobin, D., 1998. Depletion of docosahexaenoic acid in red blood cell
membranes of depressive patients. Biochem. Soc. Trans. 26, S142.
Elias, S.L., Innis, S.M., 2001. Newborn infant plasma trans conjugated linoleic, n-6 and n-3 fatty acids
are related to maternal plasma fatty acids, length of gestation and birth weight and length. Amer.
J. Clin. Nutr. 73, 807–814.
Emken, E.A., Adolf, R.O., Gulley, R.M., 1994. Dietary linoleic acid influences desaturation and
acylation of deuterium-labeled linoleic and linolenic acids in young adult males. Biochim. Biophys.
Acta 1213, 277–288.
Enslen, M., Milon, H., Malnoe, A., 1991. Effect of low intake of n-3 fatty acids during development
on brain phospholipid fatty acid composition and exploratory behaviour in rats. Lipids 26,
203–207.
Faldella, G., Govoni, M., Alessandroni, R., Marchiani, E., Salvioloi, G.P., Biagi, P.L., Spano, C., 1996.
Visual evoked potentials and dietary long chain polyunsaturated fatty acids in preterm infats. Arch.
Dis. Child. 75, F108–F112.
FAO/WHO (Food and Agriculture Organization, World Health Organization), 1993. Fats and Oils
in Human Nutrition: Report of a Joint Expert Consultation. Food and Nutrition Paper 57,
FAO, Rome.
Farquharson, J., Cockburn, F., Patrick, W.A., Jamieson, E.C., Logan, R.W., 1995. Effect of diet on the
fatty acid composition of the major phospholipids of infant cerebral cortex. Arch. Dis. Child. 72,
198–203.
Favrelière, S., Barrier, L., Darand, G., Chalon, S., Tallrieau, C., 1998. Chronic dietary n-3 polyunsatu-
rated fatty acids deficiency affects the fatty acid composition of plasmenylethanolamine and
phosphatidylethanolamine differently in rat frontal cortex, striatum, and cerebellum. Lipids 33,
401–407.
Ferdinandusse, S., Denis, S., Mooijer, P.A.W., Zhang, Z., Reddy, J.K., Spector, A.A., Wanders, R.J.A.,
2001. Identification of the peroxisomal β-oxidation enzymes involved in the biosynthesis of docosa-
hexaenoic acid. J. Lipid Res. 42, 1987–1995.
Ferretti, A., Nelson, G.J., Schmidt, P.S., Bartolini, G.L., Kelly, D.S., Flanagan, V.P., 1998. Dietary
docosahexaenoic acid reduces the thromboxane/prostacyclin synthesis ratio in humans. J. Nutr.
Biochem. 32, 79–82.
Ferrier, G.R., Redondo, I., Zhu, J., Murphy, M.G., 2002. Differential effects of docosaehexaenoic
acid on contractions and L type Ca2+ current in adult cardiomyocytes. Cardiovasc. Res. 54,
601–610.
Fidler, N., Sauerwald, T., Pohl, A., Demmelmair, H., Koletzko, B., 2000. Docosahexaenoic acid transfer
into human milk after dietary supplementation: a randomized clinical trial. J. Lipid Res. 41,
1376–1383.
Fiehn, W., Pewter, J.B., 1971. Lipids and fatty acids of sarcolemma, sarcoplasmic reticulum, and mito-
chondria from rat skeletal muscle. J. Biol. Chem. 246, 5617.
Field, C., Clandinin, M.T., van Aerde, J.E., 2001. Polyunsaturated fatty acids and T-cell function: impli-
cations for the neonate. Lipids 36, 1025–1032.
Fischer, S., 1989. Dietary polyunsaturated fatty acids and eicosanoid formation in humans. Adv. Lipid
Res. 23, 169–198.
Fliesler, S.J., Anderson, R.E., 1983. Chemistry and metabolism of lipids in the vertebrate retina. Prog.
Lipid Res. 22, 79–131.
Foote, K.D., Hrboticky, N., MacKinnon, M.J., Innis, S.M., 1990. Brain synaptosomal, liver, plasma, and
red blood cell lipids in piglets fed exclusively on a vegetable-oil-containing formula with and with-
out fish-oil supplements. Amer. J. Clin. Nutr. 51, 1001–1006.
Foreman-van Drangelen, M., van Houwelingen, A.-C., Kester, A.D.M., Hassart, T.H.M., Blanco, C.E.,
Hornstra, G., 1995. Long-chain polyunsaturated fatty acids in preterm infants, status at birth and its
fatty acids in preterm infants, status at birth and its influence on postnatal levels. J. Pediatr. 126,
611–618.
Essential fatty acid metabolism during early development 265

Frances, H., Coudereau, J.P., Sandouk, P., Clement, M., Monier, C., Bourre, J.M., 1996a. Influence of a
dietary α-linolenic acid deficiency on learning in the Morris water maze and on the effects of mor-
phine. Eur. J. Pharmacol. 298, 217–225.
Frances, H., Monier, C., Clement, M., Lecorsier, A., Debray, M., Bourre, J.M., 1996b. Effect of dietary
α-linolenic acid acid deficiency on habituation. Life Sci. 58, 1805–1816.
Funk, C.D., 2001. Prostaglandins and leukotrienes, advances in eicosanoid biology. Science 294,
1871–1875.
Galli, C., Trzeciak, H.I., Paoletti, R., 1971. Effects of dietary fatty acids on the fatty acid composition of
brain ethanolamine phosphoglyceride, reciprocal replacement of ω6 and ω3 polyunsaturated fatty
acids. Biochim. Biophys. Acta 248, 449–454.
Gamoh, S., Hashimoto, M., Sugioka, K., Hossain, M.S., Hata, N., Misawa, Y., Matsumura, S., 1999.
Chronic administration of docosahexaenoic acid improves reference memory-related learning ability
in young rats. Neuroscience 93, 237–241
Garcia, M.C., Ward, G., Ma, Y.C., Salem, N. Jr., Kim, H.Y., 1998. Effect of docosahexaenoic acid on
the synthesis of phosphatidylserine in rat brain microsomes and C6 glioma cells. J. Neurochem.
70, 24–30.
Gava, V.K., McKean, C.M., 1977. Role of 5-hydroxytryptamine in the modulation of acoustic brainstem
(far-field) potentials. Neuropharmacology 16, 447–449.
Gazzah, N., Gharib, A., Craset, M., Bobillier, P., Lagarde, M., Sarda, N., 1995. Decrease in brain
phospholipid synthesis in free-moving n-3 fatty acid deficient rats. J. Neurochem. 64, 908–918.
Gerbi, A., Zerouga, M., Debray, A., Durand, A., Chanez, C., Bourre, J.M., 1994. Effect of fish oil on fatty
acid compositon of phospholipids of brain membranes and on kinetic properties of Na+,K+-ATPase
isoenzymes of weaned and adult rats. J. Neurochem. 62, 1560–1569.
Gibson, R.A., Neuman, M.A., Makrides, M., 1997. Effect of increasing breast milk docosahexaenoic acid
on plasma and erythrocyte phospholipid fatty acids and neural indices of exclusively breast-fed
infants. Eur. J. Clin. Nutr. 51, 578–584.
Giusto, N.M., Pasquare, S.J., Salvador, G.A., Castagnet, P.I., Roque, M.E., Ilincheta de Boschero, M.G.,
2000. Lipid metabolism in vertebrate retinal rod outer segments. Prog. Lipid Res. 39, 315–391.
Green, P., Yavin, E., 1995. Modulation of fetal rat brain and liver phospholipid content by intraamniotic
ethyl docosahexaenoate administration. J. Neurochem. 65, 2555–2560.
Greiner, R.S., Moriguchi, T., Slotnick, B.M., Hutton, A., Salem, N., 2001. Olfactory discrimination
deficits in n-3 fatty acid-deficient rats. Physiol. Behav. 72, 379–385.
Greiner, R.S., Winter, J., Nathanielsz, P.W., Brenna, J.T., 1997. Brain docosahexaenoate accretion in fetal
baboons, bioequivalence of dietary α-linolenic and docosahexaenoate acids. Pediatr. Res. 42,
826–834.
Gudbjarnason, S., Doell, B., Oskarsdottir, G., 1978. Docosahexaenoic acid in cardiac metabolism and
function. Acta Biol. Med. Germ. 37, 777–784.
Guesnet, P., Pugo-Gunsam, P., Maurage, C., Pinault, M., Giraudeau, B., Alessandri, J.-M., Durand, G.,
Antoine, J.-M., Couet, C., 1999. Blood lipid concentrations of docosahexaenoic and arachidonic acids
at birth determine their relative postnatal changes in term infants fed breast milk or formula. Amer.
J. Clin. Nutr. 70, 292–298.
Haggarty, P., Page, K., Abramovich, D.R., Ashton, J., Brown, D., 1997. Long chain polyunsaturated fatty
acid transport across the human placenta. Placenta 18, 635–642.
Hamano, H., Nabekura, J., Nishikawa, M., Ogawa, T., 1996. Docosahexaenoic acid reduces GABA
response in substantia nigra neuron of rat. J. Neurophysiol. 75, 1264–1270.
Hamazaki, T., Sawazaki, S., Itomura, M., Asaoka, E., Nagao, Y., Nishimura, N., 1996. The effect of
docosahexaenoic acid on aggression in young adults: a placebo-controlled double blind study. J. Clin.
Invest. 97, 1129–1133.
Hamilton, L., Greiner, R., Salem, N. Jr., Kim, H.Y., 2000. n-3 fatty acid deficiency decreases phos-
phatidylserine accumulation selectively in neuronal tissues. Lipids 35, 863–869.
Hansen, A.E., Haggard, M.E., Boelsche, A.N., Adam, D.J.D., Wiese, H.F., 1958. Essential fatty acids in
infant nutrition. III. Clinical manifestations of linoleic acid deficiency. J. Nutr. 66, 565–576.
Hargreaves, K.M., Clandinin, M.T., 1987. Phosphatidlylethanolamine methyltransferase, evidence for
influence of diet fat on selectivity of substrate for methylation in rat brain synaptic plasma mem-
branes. Biochim. Biophys. Acta 918, 97–105.
Harris, W.S., Connor, W.E., Lindsay, S., 1984. Will dietary omega-3 fatty acids change the composition
of human milk? Amer. J. Clin. Nutr. 40, 780–785.
266 S. M. Innis

Helland, I.B., Saugstad, O.D., Smith, L., Saarem, K., Soluoll, R., Ganes, T., Drevon, C.A., 2001. Similar
effects on infants of n-3 and n-6 fatty acids supplementation to pregnant and lactating women.
Pediatrics 108, E82–E92.
Henderson, R.A., Jenson, R.G., Lammi-Keefe, C.J., Ferris, A.M., Dardick, K.R., 1992. Effect of fish oil
on the fatty acid composition of human milk and maternal and infant erythrocytes. Lipids 27,
863–869.
Hibbeln, J.R. 1998. Fish consumption and major depression. Lancet 351, 1213.
Hibbeln, J.R., Umhau, J.C., Linnoila, M., George, D.T., Ragan, P.W., Shoaf, S.E., Vaughan, M.R.,
Rawlings, R., Salem, N. Jr., 1998. A replication study of violent and non-violent subjects: cere-
brospinal fluid metabolites of serotonin and dopamine are predicted by plasma essential fatty acids.
Biol. Psychiatry 44, 243–249.
Holman, R.T., Johnson, S.B., Ogburn, P.L., 1991. Deficiency of essential fatty acids and membrane flu-
idity during pregnancy and lactation. Proc. Natl. Acad. Sci. USA 88, 4835–4839.
Hrboticky, N., MacKinnon, M.J., Innis, S.M., 1990. Effect of vegetable oil formula rich in linoleic acid
on tissue fatty acid accretion in brain, liver, plasma and erythrocytes of infant piglets. Amer. J. Clin.
Nutr. 51, 173–182.
Hrboticky, N., MacKinnon, M.J., Innis, S.M., 1991. Retina fatty acid composition of piglets fed from
birth with a linoleic acid-rich vegetable-oil formula for infants. Amer. J. Clin. Nutr. 53, 483–490.
Hrboticky, N., MacKinnon, M.J., Puterman, M.L., Innis, S.M., 1989. Effect of a linoleic acid infant
formula feeding on brain synaptosomal lipid accretion and enzyme thermotropic behavior in the
piglet. J. Lipid Res. 30, 1173–1184.
Huster, D., Arnold, K., Gawrisch, K., 1998. Influence of docosahexaenoic acid and cholesterol on lateral
lipid organization in phospholipid mixtures. Biochemistry 49, 17299–17308.
Huttenlocher, P.R., Dabholkar, A.S., 1997. Regional differences in synpatogenesis in human cerebral
cortex. J. Comp. Neurol. 20, 167–178.
Huttenlocher, P.R., de Courten, C., Garey, L.J., Van der Loos, H., 1982. Synaptogenesis in human
visual cortex: evidence for synapse elimination during normal development. Neurosci. Lett. 33,
247–253.
Ikemoto, A., Nitta, A., Furukawa, S., Ohishi, M., Nakamura, A., Fuji, Y., Okuyama, H., 2000. Dietary
n-3 fatty acid deficiency decreases nerve growth factor content in rat hippocampus. Neurosci. Lett.
285, 99–102.
Innis, S.M., 1991. Essential fatty acids in growth and development. Prog. Lipid. Res. 30, 39–103.
Innis, S.M., 1992. Human milk and formula fatty acids. J. Pediatr. 120, S56–S61.
Innis, S.M., 1996. Essential dietary lipids. In: Ziegler, E.E., Filer, L.J. (Eds.), Present Knowledge in
Nutrtion, 7th Edition. ILSI Press, Washington, DC, pp. 58–66.
Innis, S.M., 2003. Perinatal biochemistry and physiology of long chain polyunsaturated fatty acids.
J. Pediatr. 143, 4 Suppl., S1–S8.
Innis, S.M., Adamkim, D.H., Hall, R.T., Kalhan, S.C., Lair, C., Lim, M., Stevens, D.C., Twist, P.F.,
Diersen-Schade, D.A., Harris, C.L., Merkel, K.L., Hansen, J.W., 2002. Docosahexaenoic acid and
arachidonic acid from single cell triglycerides enhance growth with no adverse effects in preterm
infants fed formula. J. Pediatr. 140, 547–554.
Innis, S.M., Auestad, N., Siegman, J.S., 1996. Blood lipid docosahexaenoic acid in term gestation infants
fed formulas with high docosahexaenoic acid, low eicosapentaenoic acid fish oil. Lipids 31,
617–625.
Innis, S.M., de La Presa Owens, S., 2001. Dietary fatty acid composition in pregnancy alters neurite
membrane fatty acids and dopamine in newborn rat brain. J. Nutr. 131, 118–122.
Innis, S.M., Dyer, R.A., 2002. Brain astrocyte synthesis of docosahexaenoic acid from n-3 fatty acids is
limited at the elongation of docosapentaenoic acid. J. Lipid Res. 43, 1529–1536.
Innis, S.M., Elias, S.L., 2003. Intakes of n-6 and n-3 fatty acids among pregnant Canadian women. Amer.
J. Clin. Nutr. 77, 473–478.
Innis, S.M., Gilley, J., Werker, J., 2001. Are human-milk long-chain polyunsaturated fatty acids related
to visual and neural development in breast-fed infants? J. Pediatr. 139, 532–538.
Innis, S.M., King, D.J., 1999. Trans fatty acids in human milk are inversely associated with levels of
essential all-cis n-6 and n-3 fatty acids, and determine trans, but not n-6 and n-3 fatty acids in plasma
of breast-fed infants. Amer. J. Clin. Nutr. 70, 383–390.
Innis, S.M., Kuhnlein, H.V., 1988. Long chain n-3 fatty acids in breast milk of Inuit consuming tradi-
tional foods. Early Human Dev. 18, 185–189.
Essential fatty acid metabolism during early development 267

Insull, W., Ahrens, E.H., 1959. The fatty acids of human milk. I. Ad libitum maternal diets. J. Biochem.
72, 27–33.
Insull, W., Hirsch, J., James, J., Ahrens, E.H., 1959. The fatty acids of human milk. II. Alterations
produced by manipluation of calorie balance and exchange of dietary fats. J. Clin. Invest. 38,
443–450.
IOM (Institute of Medicine of the National Academies), 2002. Dietary Reference Intakes for Energy,
Carbohydrate, Fibre, Fat, Fatty Acids, Cholesterol, Protein and Amino Acids. The National
Academies Press, Washington, DC.
Itokazu, N., Ikegaya, Y., Nishikawa, M., Matsuki, N., 2000. Bidirectional actions of docosahexaenoic
acid on hippocampal neurotransmissions in vivo. Brain Res. 862, 211–216.
Jamieson, E.C., Farquharson, J., Logan, R.W., Howatson, A.G., Patrick, W.J.A., Weaver, L.T., Cockburn, F.,
1999. Infant cerebellar gray and white matter fatty acids in relation to age and diet. Lipids 34,
1065–1071.
Jensen, C.L., Maude, M., Anderson, R.E., Heird, W.C., 2000. Effect of docosahexaenoic acid supple-
mentation of lactating women on the fatty acid composition of breast milk lipids and maternal and
infant plasma phospholipids. Amer. J. Clin. Nutr. 71, 292S–299S.
Jensen, R.G., 1989. The Lipids of Human Milk. CRC Press, Boca Raton, FL.
Jensen, R.G., 1999. Lipids in human milk. Lipids 34, 1243–1247.
Jensen, R.G., 2002. The composition of bovine milk lipids, January 1995 to December 2000. J. Dairy Sci.
85, 295–350.
Jeppensen, P.B., Hoy, C.-E., Mortensen, P.B., 1998. Essential fatty acid deficiency in patients receiving
home parenteral nutrition. Amer. J. Clin. Nutr. 68, 126–133.
Jones, C.R., Arai, T., Rapoport, S.I., 1997. Evidence for the involvement of docosahexaenoic acid in
cholinergic stimulated signal transduction at the synapse. Neurochem. Res. 22, 663–670.
Jorgensen, M.H., Hernell, O., Hughes, E.L., Michaelsen, K.F., 2001. Is there a relation between docosa-
hexaenoic acid concentration in mothers’ milk and visual development in term infants? J. Pediatr.
Gastroenterol. Nutr. 32, 293–296.
Jump, D.B., 2002. Dietary polyunsaturated fatty acids and regulation of gene transcription. Curr. Lipidol.
13, 155–164.
Kaminski, W.E., Jendraschak, E., Kieft, R., von Schacky, C., 1993. Dietary omega-3 fatty acids
lower levels of platelet derived growth factor mRNA in human mononuclear cells. Blood 81,
1871–1879.
Kang, J.X., Leaf, A., 1996. Evidence that free polyunsaturated fatty acids modify Na+ channels by
directly binding to the channel proteins. Proc. Natl. Acad. Sci. USA 93, 3542–3546.
Khair-El-Din, T., Sicher, S.C., Vasquez, M.A., Chung, G.W., Stallworth, K.A., Kitamura, K., Miller, R.T.,
Lu, C.Y., 1996. Transcription of the murine iNOS gene is inhibited by docosahexaenoic acid, a major
constituent of fetal and neonatal sera as well as fish oil. J. Exp. Med. 183, 1241–1246.
Kinsella, J.E., Lokesh, B., 1990. Dietary lipids, eicosanoids and the immune system. Crit. Care Med. 18,
S95–S113.
Kitajka, K., Puskas, L.G., Zvara, A., Hackler, L. Jr., Barcelo-Coblijn, G., Yeo, Y.K., Farkas, T., 2002. The
role of n-3 polyunsaturated fatty acids in brain, modulation of rat brain gene expression by dietary
n-3 fatty acids. Proc. Natl. Acad. Sci. USA 99, 2619–2624.
Kneebone, G.M., Kneebone, R., Gibson, R.A., 1985. Fatty acid composition of breast milk from three
racial groups from Penang, Malaysia. Amer. J. Clin. Nutr. 41, 765–769.
Koenig, B.W., Strey, H.H., Gawrisch, K., 1997. Membrane lateral compressibility determined by NMR
and x-ray diffraction, effect of acyl chain polyunsaturation. Biophys. J. 73, 1954–1966
Koletzko, B., Braun, M., 1991. Arachidonic acid and early human growth, is there a relation? Ann. Nutr.
Metab. 35, 128–131.
Kristensen, S.D., Schmidt, E.B., Dyerberg, J., 1989 Dietary supplementation with n-3 polyunsaturated
fatty acids and human platelet function with particular emphasis on implications for cardiovascular
disease. J. Intern. Med., Suppl. 225, 141–150.
Lagarde, M., Bernoud, N., Lamaitre-Delaunay, D., Thies, F., Croset, M., Lecerf, J., 2001.
Lysophosphatidylcholine as a preferred carrier form of docosahexaenoic acid to the brain. J. Mol.
Neurosci. 16, 201–204.
Lampen, A., Meyer, S., Nau, H., 2001. Phytanic acid and docosahexaenoic acid increase the metabolism
of all-trans-retinoic acid and CYP 26 gene expression in intestinal cells. Biochim. Biophys. Acta
1521, 97–106.
268 S. M. Innis

Lamptey, M.S., Walker, B.L., 1976. A possible essential role for dietary linolenic acid in the development
of the young rat. J. Nutr. 106, 86–93.
Leaf, A., Khang, J.X., Xiao, Y.-F., Billman, G.E., Voskuyl, R.A., 1999. The antiarrhythmic and anticon-
vulsant effects of dietary n-3 fatty acids. J. Membrane Biol. 172, 1–11.
Le Doux, M., Rouzeau, A., Bas, P., Sauvant, D., 2002. Occurrence of trans-C18, 1 fatty acid isomers in
goat milk, effect of two dietary regimens. J. Dairy Sci. 85, 190–197.
Leifert, W.R., McMurchie, E.J., Saint, D.A., 1999. Inhibition of cardiac sodium currents in adult rat
myocytes by n-3 polyunsaturated fatty acids. J. Physiol. 520, 571–679.
Le Moal, M., Simon, H., 1991. Mesocorticolimbic dopaminergic networks. Physiol. Rev. 71, 155–234.
Levitsky, D.A., Strupp, B.J., 1995. Malnutrtion and the developing brain, changing concepts, changing
concerns. J. Nutr. 125, 2212S–2220S.
Levitt, P., 2003. Structural and functional maturation of the developing primate brain. J. Pediatr. 143,
S35–S45.
Leysen, J.E., Pauwels, P.J., 1990. 5-HT2 receptors, roles and regulation. Ann. N.Y. Acad. Sci. 600,
183–193.
Li, Z., Kaplan, M.L., Hachey, D., 2000. Hepatic microsomal and peroxisomal docosahexaenoate biosyn-
thesis during piglet development. Lipids 35, 1325–1333.
Litman, B.J., Mitchell, D.C., 1996. A role for phospholipid polyunsaturation in modulating membrane
protein function. Lipids 31, S193–S197.
Litman, B.J., Niu, S.L., Polozova, A., Mitchell, D.C., 2001. The role of docosahexaenoic acid containing
phospholipids in modulating G protein-coupled signaling pathways, visual transduction. J. Mol.
Neurosci. 16, 237–242.
LSRO (Life Sciences Research Office), 1998. Fat. In: Raiten, D.J., Talbot, J.M., Waters, J.H. (Eds.),
Assessment of Nutrient Requirements for Infant Formulas. LSRO, Bethesda, MD.
LSRO (Life Sciences Research Office), 2001. Fat. In: Klein, C. (Ed.), Assessment of Nutrient
Requirements for Preterm Infant Formulas. LSRO, Bethesda, MD.
Lucas, A., Stafford, M., Morley, R., Abbott, R., Stephenson, T., MacFayden, U., Elias-Jones, A.,
Clements, H., 1999. Efficacy and safety of long-chain polyunsaturated fatty acid supplementation of
infant-formula milk, a randomized trial. Lancet 354, 1948–1954.
Luthria, D.L., Mohammed, B.S., Sprecher, H., 1996. Regulation of the biosynthesis of 4,7,10,13,16-
docosapentaenoic acid. J. Biol. Chem. 271, 16020–16025.
MacDonald, M.L., Rogers, Q.R., Moris, J.G., 1983. Role of linoleate as an essential fatty acid for cat
independent of arachidonate synthesis. J. Nutr. 113, 1422–1433.
Maes, M., Christophe, A., Delanghe, J., Altamura, C., Neels, H., Meltzer, H.Y., 1999. Lowered omega-3
polyunsaturated fatty acids in serum phospholipids and cholesteryl esters of depressed patients.
Psychiatry Res. 85, 275–291.
Magret, V., Elkhalil, L., Nazih-Sanderson, F., Martin, F., Bourre, J.M., Fruchart, J.C., Delbart, C., 1996.
Entry of polyunsaturated fatty acids into the brain: evidence that high-density lipoprotein-induced
methylation of phosphatidylethanolamine and phospholipase A2 are involved. Biochem. J. 316,
805–811.
Makrides, M., Neumann, M.A., Byard, R.W., Simmer, K., Gibson, R.A., 1994. Fatty acid composition of
brain, retina, erythrocytes in breast- and formula-fed infants. Amer. J. Clin. Nutr. 60, 189–194.
Makrides, M., Neumann, M.A., Gibson, R.A., 1996. Effect of maternal docosahexaenoic acid (DHA)
supplementation on breast milk composition. Eur. J. Clin. Nutr. 50, 352–357.
Makrides, M., Neumann, M.A., Simmer, K., Gibson, R.A., 1995a. Erythrocyte fatty acids of term infants
fed either breast milk, standard formula, or formula supplemented with long-chain polyunsaturates.
Lipids 30, 941–948.
Makrides, M., Neumann, M.A., Simmer, K., Gibson, R.A., 2000. A critical appraisal of the role of dietary
long-chain polyunsaturated fatty acids on neural indices of term infants, a randomized, controlled
trial. Pediatrics 105, 32–38.
Makrides, M., Neumann, M., Simmer, K., Pater, J., Gibson, R., 1995c. Are long-chain polyunsaturated
fatty acids essential nutrients in infancy? Lancet 345, 1463–1468.
Makrides, M., Simmer, K., Neumann, M., Gibson, R., 1995b. Changes in the polyunsaturated fatty acids
of breast milk from mothers of full-term infants over 30 wk of lactation. Amer. J. Clin. Nutr. 61,
1231–1233.
Martinez, M., 1992. Tissue levels of polyunsaturated fatty acids in early human development. J. Pediatr.
120, 129–138.
Essential fatty acid metabolism during early development 269

Martinez, M., Vazquez, E., 1998. MRI evidence that docosahexaenoic acid ethyl ester improves myeli-
nation in generalized peroxisomal disorders. Neurology 51, 26–32.
Martinez, M., Vazquez, E., Garcia de Silva, M.T., Manzanares, J., Bertran, J.M., Castello, F., 2000.
Therapeutic effects of docosahexaenoic acid ethyl ester in patients with generalized peroxisomal
disorders. Amer. J. Clin. Nutr. 71, 376S–385S.
Mata de Urquiza, A., Liu, S., Sjoberg, M., Zetterstrom, R.H., Griffiths, W., Sjovall, J., Perlmann, T.,
2000. Docosahexaenoic acid, a ligand for the retinoid X receptor in mouse brain. Science 290,
2140–2144.
McGahon, B.M., Martin, D.S., Horrobin, D.F., Lynch, M.A., 1999. Age-related changes in synaptic func-
tion, analysis of the effect of dietary supplementation with omega-3 fatty acids. Neuroscience 94,
305–314.
McLennan, P.L., 2001. Myocardial membrane fatty acids and the antiarrhythmic actions of dietary fish
oil in animal models. Lipids 36, S111–S114.
Menard, C.R., Goodman, K.J., Corso, T.N., Brenna, J.T., Cunnane, S.C., 1998. Recycling of carbon into
lipids synthesized de novo is a quantitatively important pathway of α-[U-13C] linolenate utilization in
developing rat brain. J. Neurochem. 71, 2151–2158.
Minami, M., Kimura, S., Endo, T., Hamaue, N., Hirafugi, M., Togashi, H., Matsumoto, M., Yoshioka, M.,
Saito, H., Watanabe, S., Kobayashi, T., Okuyama, H., 1997. Dietary docosahexaenoic acid increases
cerebral acetylcholine levels and improves passive avoidance performance in stroke-prone sponta-
neously hypertensive rats. Pharmacol. Biochem. Behav. 58, 1123–1129.
Mitchell, D.C., Litman, B.J., 1998. Molecular order and dynamics in highly polyunsaturated phospho-
lipid bilayers. Biophys. J. 75, 896–908.
Mitchell, D.C., Straume, M., Litman, B.J., 1992. Role of sn-1-saturated, sn-2-polyunsaturated phospho-
lipids in control of membrane receptor conformational equilibrium, effects of cholesterol and acyl
chain unsaturation on the metarhodopsin I in equilibrium with metarhodopsin II equilibrium.
Biochemistry 31, 662–670.
Mohammad, B.S., Luthria, D.L., Bakousheva, S.P., Sprecher, H., 1997. Regulation of the biosynthesis of
4,7,10,13,16-docosapentaenoic acid. J. Biol. Chem. 326, 425–430.
Mohammed, B.S., Sankarappa, S., Geiger, M., Sprecher, H., 1995. Re-evaluation of the pathway for the
metabolism of 7,10,13,16-docosatetraenoic acid to 4,7,10,13,16-docosapentaenoic acid in rat liver.
Acta Biochem. Biophys. 317, 179–184.
Mohrhauer, H., Holman, R.T., 1963. The effect of dose level of essential fatty acids upon fatty acid
composition of the rat liver. J. Lipid Res. 4, 153–159.
Moore, S.A., 1994. Local synthesis and targeting of essential fatty acids at the cellular interface between
blood and brain: a role for cerebral endothelium and astrocytes in the accretion of CNS docosa-
hexaenoic acid. World Rev. Nutr. Diet. 75, 128–133.
Moore, S.A., 2001. Polyunasaturated fatty acid synthesis and release by brain-derived cells in vitro.
J. Mol. Neurosci. 16, 195–200.
Moore, S.A., Yoder, E., Murphy, S., Dutton, G.R., Spector, A.A., 1991. Astrocytes, not neurons, produce
docosahexaenoic acid (22:6n-3) and arachidonic acid (20:4n-6). J. Neurochem. 56, 518–524.
Morales, M.S., Palmquist, D.L., Weiss, W.P., 2000. Effects of fat source and copper on unsatura-
tion of milk and blood triacylglycerol fatty acids in Holstein and Jersey cows. J. Dairy Sci. 83,
2105–2111.
Morgane, P.J., Autin-LaFrance, R.J., Bronzinio, J.D., Galler, J.R., 1992. Malnutrition and the developing
central nervous system. In: Issacson, R.L., Jensen, K.F. (Eds.), The Vulnerable Brain and
Environmental Risks, Vol. 1: Malnutrition and Hazard Assessment. Plenum Press, New York, pp. 3–44.
Moriguchi, T., Greiner, S.R., Salem, N. Jr., 2000. Behavioural deficits associated with dietary induction
of decreased brain docosahexaenoic acid concentration. J. Neurochem. 75, 2563–2573.
Moriguchi, T., Loewke, T., Garrison, M., Catalan, J., Salem, N. Jr., 2001. Reversal of docosahexaenoic
acid deficiency in the rat brain, retina, liver and serum. J. Lipid Res. 42, 419–427.
Mosier, M., Newton, A.C., 1998. Mechanisms of apparent cooperativity in the interaction of protein
kinase C with phosphatidylserine. Biochem. J. 37, 17271–17279.
Murthy, M., Hamilton, J., Greiner, R.S., Moriguchi, T., Salem, N. Jr., Kim, H.Y., 2002. Differential
effects of n-3 fatty acids on phospholipid molecular species composition in the rat hippocampus.
J. Lipid Res. 43, 611–617.
Nair, S.S., Leitch, J.W., Falconer, J., Garg, M.L., 1997. Prevention of cardiac arrhythmia by dietary
(n-3) polyunsaturated fatty acids and their mechanism of action. J. Nutr. 127, 383–393.
270 S. M. Innis

Neuringer, M., Connor, W.E., 1986. n-3 fatty acids in the brain and retina, evidence for their essentiality.
Nutr. Rev. 44, 285–294.
Neuringer, M., Connor, W.E., Lin, D.S., Barstad, L., Luck, S. 1986. Biochemical and functional effects
of prenatal and postnatal ω-3 fatty acid deficiency on retina and brain in rhesus monkeys. Proc.
Natl. Acad. Sci. USA 83, 4021–4025.
Neuringer, M., Connor, W.E., Van Petten, C., Barstad, L., 1984. Dietary ω-3 fatty acid deficiency and
visual loss in infant rhesus monkeys. J. Clin. Invest. 73, 272–276.
Nishikawa, M., Kimura, S., Akaike, N., 1994. Facilitatory effect of docosahexaenoic acid on
N-methyl-D-aspartate response in pyramidal neurons of rat cerebral cortex. J. Physiol. 475, 83–93.
Ntambi, J.M., Buhrow, S.A., Kaestner, K.H., Christy, R.J., Sibley, E., Kelley, T.J., Lane, M.D., 1988.
Differentiation and induced gene expression in 3T3-L1 preadipocytes: characterization of a differen-
tially expressed gene encoding stearoyl-CoA desaturase. J. Biol. Chem. 263, 17291–17300.
O’Brien, J.S., Sampson, E.L., 1965. Fatty acid and fatty aldehyde composition of the major brain lipids
in normal human gray matter, white matter, and myelin. J. Lipid Res. 6, 545–551.
O’Connor, D.L., Auestad, N., Jacobs, J., 2001. Growth and development in preterm infants fed
long-chain polyunsaturated fatty acids, a prospective randomized controlled trial. Pediatrics 108,
359–372.
Okada, M., Amamoto, T., Tomonaga, M., Kawachi, A., Yazawa, K., Mine, K., Fujiwara, M., 1996. The
chronic administration of docosahexaenoic acid reduces the spatial cognitive deficit following
transient forebrain ischemia in rats. Neuroscience 71, 17–25.
Ollero, M., Powers, R.D., Alvarez, J.G., 2000. Variation of docosahexaenoic acid content in subsets of
human spermatozoa at different stages of maturation: implications for sperm lipoperoxidative
damage. Mol. Reprod. Dev. 55, 326–334.
Olsen, S.F., Hansen, H.S., Secher, N.J., Jensen, B., Sandstrom, B., 1995. Gestation length and birth
weight in relation to the intake of marine n-3 fatty acids. Brit. J. Nutr. 73, 397–404.
Olsen, S.F., Hansen, H.S., Sommer, S., Jensen, B., Sorensen, T.I.A., Secher, N.J., Zachariassen, P., 1991.
Gestational age in relation to marine n-3 fatty acids in maternal erythocytes, a study of women in
Faroe Islands and Denmark. Amer. J. Obstet. Gynecol. 164, 1203–1209.
Olsen, S.F., Sorensen, J.D., Secher, N.J., Hedegaard, M., Henriksen, T.B., Hansen, H., Grant, A., 1992.
Randomized controlled trial of effect of fish oil supplementation on pregnancy duration. Lancet 339,
1003–1007.
Pakala, R., Sheng, W.L., Benedict, C.R., 2000. Vascular smooth muscle cells preloaded with eicosapen-
taenoic acid and docosahexaenoic acid fail to respond to serotonin stimulation. Atherosclerosis 153,
47–57.
Park, C.C., Ahmed, Z., 1992. Alterations of plasma membrane fatty acid composition modify the kinetics
of Na+ current in cultured rat diencephalic neurons. Brain Res. 570, 75–84.
Parodi, P.W., 1982. Positional distribution of fatty acids in triglycerides from milk of several species of
mammals. Lipids 17, 437–442.
Pawlosky, R.J., Denkins, Y., Ward, G., Salem, N. Jr., 1997. Retinal and brain accretion of long-chain
polyunsaturated fatty acids in developing felines, the effects of corn oil-based maternal diets. Amer.
J. Clin. Nutr. 65, 465–472.
Pawlosky, R.J., Hibbeln, J.R., Novotny, J.A., Salem, N. Jr., 2001. Physiological compartmental analysis
of α-linolenic acid metabolism in adult humans. J. Lipid Res. 42, 1257–1265.
Poisson, J.P., Dupuy, R.P., Sarda, P., Descomps, B., Narce, M., Rieu, D., Crastes de Paulet, A., 1993.
Evidence that liver microsomes of human neonates desaturate essential fatty acids. Biochim. Biophys.
Acta 1167, 109–113.
Poling, J.S., Karanian, J.W., Salem, N. Jr., Vincini, S., 1995. Time- and voltage-dependent block of potas-
sium channels by docosahexaenoic acid. Mol. Pharmacol. 47, 381–390.
Pond, W.G., Boleman, S.L., Firotto, M.L., Ho, H., Knabe, D.A., Mersmann, H.J., Savell, J.W., Su, D.R.,
2000. Perinatal ontogeny of brain growth in the domestic pig. Proc. Soc. Exp. Biol. Med. 223,
102–108.
Ponder, D.L., Innis, S.M., Benson, J.D., Siegman, J., 1992. Docosahexaenoic acid status of term infants
fed breast milk or infant formula containing soy oil or corn oil. Pediatr. Res. 32, 683–688.
Pooviah, B.P., Tinoco, J., Lyman, R.L., 1976. Influence of diet on the conversion of 14C-linolenic acid to
docosahexaenoic acid in the rat. Lipids 1, 194–202.
Poulos, A., Darin-Bennett, A., White, I.G., 1973. The phospholipid bound fatty acids and aldehydes of
mammalian spermatozoa. Comp. Biochem. Physiol. 46, 541.
Essential fatty acid metabolism during early development 271

Putnam, J.C., Carlson, S.E., Phillip, W., DeVoe, M.D., Barness, L.A., 1982. The effect of variations
in dietary fatty acids on the fatty acid composition of erythrocyte phosphatidylcholine and
phosphatidylethanolamine in human infants. Amer. J. Clin. Nutr. 36, 106–114.
Reginato, M.J., Krakow, S.L., Bailey, S.T., Lazar, M.A., 1998. Prostaglandins promote and block adipo-
genesis through opposing effects on peroxisome prolierator-activated receptor gamma. J. Biol. Chem.
273, 1855–1858.
Reisbick, S., Neuringer, M., Hasnain, R., Connor, W.E., 1990. Polydipsia in rhesus monkeys deficient in
omega-3 fatty acids. Physiol. Behav. 47, 315–323.
Reisbick, S., Neuringer, M., Hasnain, R., Connor, W.E., 1994. Home cage behavior of rhesus monkeys
with long-term deficiency of omega-3 fatty acids. Physiol. Behav. 55, 231–239.
Reseland, J.E., Haugen, F., Hollung, K., Solvoll, K., Halvorsen, B., Brude, I.R., Nenseter, M.S.,
Christiansen, E.N., Drevon, C.A., 2001. Reduction of leptin gene expression by dietary polyunsatu-
rated fatty acids. J. Lipid Res. 42, 743–750.
Rioux, F.M., Innis, S.M., Dyer, R., MacKinnon, M., 1997. Diet-induced changes in liver and bile but not
brain fatty acids can be predicted from differences in plasma phospholipid fatty acids in milk and for-
mula fed piglets. J. Nutr. 127, 370–377.
Rivers, J.P.W., Mausan, A.G., Crawford, M.A., Brantell, M.R., 1976. The inability of the lion, Panthero
leo L., to desaturate linoleic acid. FEBS Lett. 67, 269–270.
Rivers, J.P.W., Sinclair, A.J., Crawford, M.A., 1975. Inability of the cat to desaturate essential fatty acids.
Nature 258, 171–173.
Rodriguez de Turco, E.B., Parkins, N., Ershov, A.V., Bazan, N.G., 1999. Selective retinal pigment epithe-
lial cell lipid metabolism and remodelling conserves photoreceptor docosahexaenoic acid following
phagocytosis. J. Neurosci. Res. 57, 479–486.
Rogers, S., James, K.S., Butland, B.K., 1987. Effects of a fish oil supplement on serum lipids, blood
pressure, bleeding time, hemostatic and rheological variables: a double blind randomized controlled
trial in healthy volunteers. Atherosclerosis 63, 137–143.
Ryan, A.S., Montalto, M.B., Groh-Wargo, S., Mimouni, F., Sentipal-Walerius, J., Doyle, J., Siegman, J.S.,
Thomas, A.J., 1999. Effect of DHA-containing formula on growth of preterm infants to 59 weeks
postmenstrual age. Amer. J. Human Biol. 11, 457–467.
Salem Jr., N., Vegher, B., Mena, P., Uauy, P., 1996. Arachidonic and docosahexaenoic acids are biosyn-
thesized from their 18-carbon precursors in human infants. Proc. Natl. Acad. Sci. USA 93, 45–54.
Sanders, T.A.B., Ellis, F.R., Dickerson, J.W.T., 1978. Studies of vegans, the fatty acid composition of
plasma choline phosphoglycerides, erythrocytes, adipose tissue, and breast milk, and some indicators
of susceptibility to ischemic heart disease in vegans and omnivore controls. Amer. J. Clin. Nutr. 31,
805–813.
Sanders, T.A.B., Rana, S.K., 1987. Comparison of the metabolism of linoleic and linolenic acids in the
fetal rat. Ann. Nutr. Metab. 31, 349–353.
San Giovanni, J.P., Parra-Cabrera, S., Colditz, G.A., Berkey, C.S., Dwyer, J.T., 2000. Meta-analysis of
dietary essential fatty acids and long-chain polyunsaturated fatty acids as they relate to visual resolu-
tion acuity in healthy preterm infants. Pediatrics 105, 1292–1298.
Sastry, P.S., 1985. Lipids of nervous tissue, composition and metabolism. Prog. Lipid Res. 24, 169–176.
Sauerwald, T.U., Hachey, D.L., Jensen, C.L., Chen, H., Anderson, R.E., Heird, W.C., 1996. Effect of
dietary α-linolenic acid intake on incorporation of docosahexaenoic and arachidonic acids into
plasma phospholipids of term infants. Lipids 31, S131–S135.
Sauerwald, T.U., Hachey, D.L., Jensen, C.L., Chen, H., Anderson, R.E., Heird, W.C., 1997. Intermediates
in endogenous synthesis of C22:6ω3 and C20:4ω6 by term and preterm infants. Pediatr. Res.
41,183–187.
Shand, J.H., Noble, R.C., 1981. The metabolism of 18:0 and 18:2(n-6) by the ovine placenta at 120 and
150 days of gestation. Lipids 16, 68–71.
Sheaff-Greiner, R.C., Zhang, Q., Goodman, K.J., Giussani, D.A., Nathanielsz, P.W., Brenna, J.T., 1996.
Linoleate, α-linoleate, and docosahexaenoate recycling into saturated and monounsaturated fatty
acids is a major pathway in pregnant or lactating adults and fetal or infant rhesus monkeys. J. Lipid
Res. 37, 2675–2686.
Simopoulos, A.P., 1999. Evolutionary aspects of omega-3 fatty acids in the food supply. Prostagland.
Leuk. Essent. Fatty Acids. 60, 421–429.
Sinclair, A.J., Crawford, M.A., 1972. The incorporation of linolenic acid and docosahexaenoic acid into
liver and brain lipids of developing rats. FEBS Lett. 26, 127–129.
272 S. M. Innis

Sinclair, A.J., Crawford, M.A., 1972. The accumulation of arachidonate and docosahexaenoate in the
developing rat brain. J. Neurochem. 99, 1753–1758.
Sprecher, H., Chen, Q., Yin, F.Q., 1999. Regulation of the biosynthesis of 22:5n-6 and 22:6n-3, a
complex intracellular process. Lipids 34, S153–S156.
Sprecher, H., Luthria, D.L., Mohammed, B.S., Baykousheva, S.P., 1995. Reevaluation of the pathways
for biosynthesis of polyunsaturated fatty acids. J. Lipid Res. 36, 2471–2477.
Stockard, J.E., Saste, M.D., Benford, V.J., Barness, L., Auestad, J.D., 2000. Effect of docosahexaenoic
acid content of maternal diet on auditory brainstem conduction times in rat pups. Dev. Neurosci. 22,
494–499.
Stoll, A.L., Severus, W.E., Freeman, M.P., Rueter, S., Zboyan, H.A., Diamond, E.A., 1999. Omega-3 fatty
acids in bipolar disorder, a preliminary double-blind, placebo-controlled trial. Arch. Gen. Psychiatry
56, 407–412.
Su, H.M., Bernardo, M., Mirmiran, X.H., Ma, T.N., Corso, P.W., Nathanielsz, J.T., Brenna, J.T.,
1999. Bioequivalence of dietary α-linolenate and docosahexaenoate acids as possible sources
of docosahexaenoate accretion in brain and associated organs of neonatal baboons. Pediatr. Res.
45, 87–93.
Su, H.-M., Huan, M.-C., Saad, N.M.R., Nathanielsz, P.W., Brenna, J.T., 2001. Fetal baboons convert
18:3n-3 to 22:6n-3 in vivo, a stable isotope tracer study. J. Lipid. Res. 42, 581–586.
Suh, M., Wierzbicki, A.A., Lien, E., Clandinin, M.T., 1996. Relationship between dietary supply of
long-chain fatty acids and membrane composition of long- and very long chain essential fatty acids
in developing rat photoreceptors. Lipids 313, 61–64.
Suh, M., Wierzbicki, A.A., Lien, E., Clandinin, M.T., 2000. Dietary 20:4n-6 and 22:6n-3 modulates the
profile of long- and very-long-chain essential fatty acids, rhodopsin content, and kinetics in develop-
ing photoreceptor cells. Pediatr. Res. 48, 524–530.
Thies, F., Delachambre, M.C., Bentejac, M., Lagarde, M., Lecerf, J., 1992. Unsaturated fatty acids
esterified in 2-acyl-1-lysophosphatidylcholine bound to albumin are more efficiently taken up by the
young rat brain than the unesterified from. J. Neurochem. 59, 1110–1116.
Thies, F., Pillon, C., Moliere, P., Lagarde, M., Lecerf, J., 1994. Preferential incorporation of sn-2 lysoPC
DHA over unesterified DHA in young rat brain. Amer. J. Physiol. 267, R1273–1279.
Tinoco, J., 1982. Dietary requirements and functions of α-linolenic acid in animals. Prog. Lipid Res.
21, 1–45.
Tinoco, J., Babcock, R., Hincenbergs, I., Medwadowski, B., Jiljanich, P., Williams, M.A., 1979.
Linolenic acid deficiency. Lipids 14, 166–173.
Tremoli, E., Maderna, P., Marangoni, F., Colli, S., Eligini, S., Catalano, I., Angeli, M.T., Pazzucconi, F.,
Gianfranceschi, G., Davi, G., 1995. Prolonged inhibition of platelet aggregation after n-3 fatty acid
ethyl ester ingestion by healthy volunteers. Amer. J. Clin. Nutr. 61, 607–613.
Uauy, R., Birch, D.G., Birch, E.E., 1992. Retinal development in very low birthweight infants fed diets
differing in omega-3 fatty acids. Pediatr. Res. 28, 485–492.
Uauy, R., Mena, P., Wegher, B., Nieto, S., Salem, N. Jr., 2000. Long chain polyunsaturated fatty acid
formation in neonates, effect of gestational age and intrauterine growth. Pediatr. Res. 47, 127–135.
Vance, D.E., 1990. Phosphatidylcholine metabolism, masochistic enzymology, metabolic regulation, and
lipoprotein assembly. Biochem. Cell. Biol. 68, 1151–1165.
van Houwelingen, A.C., Sorensen, J.D., Hornstra, G., Simonis, M.M., Boris, J., Olsen, S.F., Secher, N.J.,
1995. Essential fatty acid status in neonates after fish-oil supplementation during late pregnancy.
Brit. J. Nutr. 74, 723–731.
von Schacky, C., Weber, P.C., 1985. Metabolism and effects on platelet function of the purified
eicosapentaenoic and docosahexaenoic acids in humans. J. Clin. Invest. 76, 2446–2450.
Voss, A., Reinhart, M., Sankarappa, S., Sprecher, H., 1991. The metabolism of 7,10,13,16,
19-docosapentaenoic acid to 4,7,10,13,16,19-docosahexaenoic acid in rat liver is independent of a
4-desaturase. J. Biol. Chem. 226, 19995–20000.
Wainwright, P.E., Xing, H.C., Girard, T., Parker, L., Ward, G.R., 1998. Effects of dietary n-3 fatty acid
deficiency on Morris water-maze performance and amphetamine conditioned place preference in rats.
Nutr. Neurosci. 1, 281–293.
Wanders, R.J., Vreken, P., Ferdinandusse, S., Jansen, G.A., Waterham, C.W., Van Grunsven, E.G., 2001.
Peroxisomal fatty acid α- and β-oxidation in human enzymology, peroxisomal metabolite trans-
porters and peroxisomal diseases. Biochem. Soc. Trans. 29, 250–267.
Essential fatty acid metabolism during early development 273

Wang, N., Anderson, R.E., 1993. Synthesis of docosahexaenoic acid by retina and retina pigment epithe-
lium. Biochemistry 32, 13703–13709.
Ward, G.R., Huang, Y.S., Bobik, E., Xing, H.C, Mutsaers, L., Auestad, N., Montalto, M., Wainwright, P.,
1998. Long-chain polyunsaturated fatty acid levels in formulae influence deposition of docosa-
hexaenoic acid and arachidonic acid in brain and red blood cells of artificially reared neonatal rats.
J. Nutr. 128, 2473–2487.
Washizaki, K., Smith, Q.R., Rapoport, S.I., Purdon, A.D., 1994. Brain arachidonic acid incorporation and
precursor pool specific activity during intravenous infusion of unesterified [3H]arachidonate in the
anesthetized rat. J. Neurochem. 63, 727–736.
Weiler, H.A., Fitzpatrick-Wong, S., 2002. Dietary long-chain polyunsaturated fatty acids minimize
dexamethasone-induced reductions in arachidonic acid status but not bone mineral content in piglets.
Pediatr. Res. 51, 282–289.
Weisinger, H.S., Vinigrys, A.J., Bui, B.V., Sinclair, A.J., 1999. Effects of dietary n-3 fatty acid deficiency
and repletion in the guinea pig retina. Invest. Ophthalmol. Vis. Sci. 40, 327–328.
Weisinger, H.S., Vingrys, A.J., Sinclair, J.J., 1995. The effect of docosahexaenoic acid on the
electroretinogram of the guinea pig. Lipids 31, 65–70.
Wertz, P.W., Swartzendruber, D.C., Abraham, W., Madison, K.C., Downing, D.T., 1987 Essential fatty
acids and epidermal integrity. Arch. Dermatol. 123, 1381–1384.
Wheeler, T.G., Benolken, R.M., Anderson, R.E., 1975. Visual membranes, specificity of fatty acid
precursors for the electrical response to illumination. Science 188, 1312–1314.
Wiggins, R.C., 1986. Myelination, a critical stage in development. Neurotoxicology 7, 103–120.
Willatts, P., Forsyth, P., Forsyth, J.S., DiModugno, M.K., Varna, S., Colvin, M., 1998. Effect of long-
chain polyunsaturated fatty acids in formula on infant problem solving at 10 months of age. Lancet
352, 688–691.
Williard, D., Harmon, S., Kaduce, T., Spector, A., 2002. Comparison of 20-, 22-, and 24-carbon n-3 and
n-6 polyunsaturated fatty acid utilization in differentiated rat brain astrocytes. Prostagland. Leuk.
Essent. Fatty Acids 67, 99–104.
Yamamoto, N., Hashimoto, A., Takemoto, Y., Okuyama, H., Nomura, M., Kitajima, R., Togashi, T.,
Tamai, Y., 1988 Effects of the dietary α-linoleate/linoleate balance on lipid compositions and learn-
ing ability of rats. II. Discrimination process, extinction process, and glycolipid compositions.
J. Lipid Res. 29, 1013–1021.
Yamazaki, K., Fujikawa, W., Hamazaki, T., Yano, S., Shono, T., 1992. Comparison of the conversion rates
of α-linolenic acid (18,3(n - 3)) and stearidonic acid (18,4(n - 3)) to longer polyunsaturated fatty acids
in rats. Biochim. Biophys. Acta 1123, 18–26.
Yonekubo, A., Honda, S., Okano, M., Yamamoto, Y., 1993. Effects of dietary safflower oil or soybean oil
on the milk composition of the maternal rat, and tissue fatty acid composition and learning ability of
postnatal rats. Biosci. Biotech. Biochem. 57, 253–259.
Yoshida, S., Yasuda, A., Kawazato, K., Sakai, K., Shimada, T., Takeshita, Y., Yuasa, S., Kobayashi, T.,
Watanabe, S., Okuyama, H., 1997. Synaptic vesicle ultrastructural changes in the rat hippocampus
induced by a combination of α-linolenate deficiency and a learning task. J. Neurochem. 68,
1261–1268.
Young, C., Gean, P., Wu, S., Lin, C., Shen, Y., 1998. Cancellation of low-frequency stimulation-
induced long-term depression by docosahexaenoic acid in the rat hippocampus. Neurosci. Lett. 247,
198–200.
Youyou, A., Durand, G., Pascal, G., Piciotti, M., Dumont, O., Bourre, J.M., 1986. Recovery of altered
fatty acid composition induced by a diet devoid of n-3 fatty acids in myelin, synaptosomes, mito-
chondria and microsomes of developing rat brain. J. Neurochem. 46, 224–228.
Ziboh, V.A., Chapkin, R.S., 1988. Metabolism and function of skin lipids. Prog. Lipid Res. 27,
81–105.
Zimmer, L., Breton, P., Durand, G., Guilloteau, D., Besnard, J.C., Chalon, S., 1999. Prominent role of
n-3 polyunsaturated fatty acids in cortical dopamine metabolism. Nutr. Neurosci. 2, 257–265.
Zimmer, L., Delion-Vancassel, S., Durand, G., Guilloteau, D., Bodard, S., Besnard, J.C., Chalon, S.,
2000a. Modification of dopamine neurotransmission in the nucleus accumbens of rats deficient in
n-3 polyunsaturated fatty acids. J. Lipid Res. 41, 32–40.
Zimmer, L., Delpal, S., Guilloteau, D., Aioun, J., Durand, G., Chalon, S., 2000b. Chronic n-3 polyunsat-
urated fatty acid deficiency alters dopamine vesicle density in the rat frontal cortex. Neurosci. Lett.
284, 25–28.
274 S. M. Innis

Zimmer, L., Hembert, S., Durand, G., Breton, P., Guilloteau, D., Besnard, J.-C., Chalon, S., 1998.
Chronic n-3 polyunsaturated fatty acid diet-deficiency acts on dopamine metabolism in the rat frontal
cortex, a microdialysis study. Neurosci. Lett. 240, 177–181.
Zimmer, L., Vancassel, S., Contagrel, S., Breton, P., Delmanche, S., Guilloteau, D., Durand, G., 2002.
The dopamine mesocorticolimbic pathway is affected by deficiency in n-3 polyunsaturated fatty
acids. Amer. J. Clin. Nutr. 75, 662–667.
11 Development of white adipose tissue
lipid metabolism1

H. J. Mersmanna and S. B. Smithb

a USDA/ARS Children’s Nutrition Research Center, Department of Pediatrics,


Baylor College of Medicine, Houston, TX 77030, USA
b Department of Animal Science, Texas A & M University, College Station,

TX 77843-2471, USA

Most mammals are born with little white adipose tissue; however, the guinea pig and human
are exceptions. Limited adipose tissue places the newborn at risk when confronted with
environmental challenges such as cold temperatures or limited milk supply. White adipose
tissue develops rapidly after birth in those species with limited depots. Adipose tissue growth
is a combination of cell proliferation coupled with differentiation and cell hypertrophy.
Proliferation is a property of the undifferentiated preadipocyte and is predominant in neonatal
development. It continues at a lower rate to accommodate growth and replace cells, but can
be activated, even in adults, when caloric intake is excessive and continuous. Preadipocytes
differentiate into adipocytes with subsequent growth of the adipocyte, increase in mass of the
tissue being the result of accumulation of triacylglycerol in a large central intracellular lipid
droplet. Glucose is the carbon precursor of fatty acid synthesis in adipocytes from nonrumi-
nant mammals, whereas acetate is the carbon precursor in ruminant species. The human has
little or no capacity for adipocyte de novo fatty acid synthesis. Because milk is a high-fat
food, there is little fatty acid synthesis during the suckling period. De novo fatty acid synthesis
is primarily a process of importance in the postweaning mammal. Triacylglycerol is carried
in lipoproteins and these are cleaved by lipoprotein lipase to yield fatty acids that are readily
absorbed by adipocytes. Fatty acids are esterified in the adipocyte to triacylglycerol. The
lipoprotein lipase and triacylglycerol biosynthetic activities increase rapidly after birth to
allow uptake and esterification of fatty acids. The newborn mammal also needs fatty acids

1 This work is a publication of the USDA/ARS Children’s Nutrition Research Center, Department of Pediatrics, Baylor

College of Medicine, Houston, Texas. This project has been funded in part with federal funds from the USDA/ARS
under Cooperative Agreement No. 58-6250-6001. The contents of this publication do not necessarily reflect the views
or policies of the USDA, nor does mention of trade names, commercial products, or organizations imply endorsement
by the U.S. Government.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
275 © 2005 Elsevier Limited. All rights reserved.
276 H. J. Mersmann and S. B. Smith

for an oxidative fuel. The source of these fatty acids is from the diet and immediately from
circulating lipoproteins coupled with fatty acids released from the adipocyte by lipolysis, the
breakdown of adipocyte triacylglycerol.

1. INTRODUCTION
There are two types of adipose tissue in mammals, white and brown. The brown adipocyte is
a specialized cell for generation of heat. It has multiple lipid droplets, i.e. stores of triacyl-
glycerol (TG), but also a large number of mitochondria to oxidize long-chain fatty acids (FAs)
mobilized from TG. Brown adipose tissue is present in most mammalian newborns. It is
strategically located in the body, i.e. in the thorax and near the kidneys, to provide heat to
essential organs, by FA mobilization and inefficient oxidation. Brown adipose tissue properties,
developmental patterns, and functions are discussed by Smith and Carsten (Chapter 12). The
newborn pig is different from most newborn mammals in that it has no brown adipose tissue.
The white adipocyte is characterized by a large central lipid droplet that is the repository
for storage of energy in the form of TG containing esterified FAs. Most mammalian species
are born with little white adipose tissue (<5% of the body weight at birth). Most mammalian
fetuses have a limited capacity to synthesize fatty acids de novo and the passage of FAs across
the placenta is highly restricted (Noble, 1981; Martin et al., 1985; Hamosh, 1998; Kimura,
1998; Herrera, 2002). However, the guinea pig and human are born with 10% and 15% body
fat (Bonnet, 1981; Widdowson and Lister, 1991); in these species, de novo FA synthesis
increases markedly during the last trimester of pregnancy. White adipose tissue has multiple
functions, but in the neonate, the two most important are for insulation, primarily provided by
the subcutaneous fat depots, and as a repository for energy storage to be mobilized when the
nutrient intake does not provide the required energy. The uterine environment provides shelter
in the way of temperature control and provision of nutrients, both of which are lost at partu-
rition. Mammalian newborns are generally in a precarious position if they are challenged by
the environment, e.g. cold exposure, or if they cannot obtain sufficient nutrients by way of the
diet, i.e. by suckling. The newborn pig is at a particular disadvantage because it has <2% body
fat limiting the insulation and energy supplies, the FA mobilization and oxidation capacity is
reduced, there is very little hair (for insulation), and there is no brown adipose tissue
(Mersmann, 1974). The neonatal period is characterized by a multitude of metabolic changes
as the organism adapts to the many challenges of the environment. The exact timing of a par-
ticular adaptation varies among species, but these changes occur during the window of time
between the last days of gestation and the first days or weeks postpartum. More extensive dis-
cussions of metabolism in the neonatal pig are found in Chapter 14 by Herpin et al. and
Chapter 9 by Odle et al.
The distribution of adipose tissue is different in different species, as are the growth rates
for the individual depots (Berg and Walters, 1983; Trenkle and Marple, 1983). For example,
the predominant adipose tissue depot in the pig is the subcutaneous depot, with lesser fat
deposition at the perirenal, mesenteric/omental, and intermuscular sites (Walstra, 1980; Kauffman
et al., 1986; Kouba et al., 1999; Mitchell et al., 2001). In sheep, the subcutaneous adipose tissue
depot is large, but the intermuscular depot is almost as large and the omental depot is about 50%
of the subcutaneous depot (Moloney et al., 2002). In cattle, the subcutaneous depot is large,
particularly in breeds that are selected for muscle production and that fatten readily, whereas in
breeds with accentuated lactation rates, the internal fat depots are more extensively developed
(Truscott et al., 1983). The metabolic activity and the differentiation of individual adipose tissue
depots may be markedly different (Adams et al., 1997; Wajchenberg, 2000).
Development of white adipose tissue lipid metabolism 277

We describe the development of white adipose tissue, and the key metabolic pathways that
provide the primary functions of adipose tissue, the storage and mobilization of energy. There
are several major reviews of the literature regarding these subjects in cattle, pigs, and sheep
(Allen et al., 1976; Vernon, 1980, 1981; Mersmann, 1986; Smith and Smith, 1995).

2. DEVELOPMENT OF WHITE ADIPOSE TISSUE


A single large central lipid droplet occupying much of the volume of the cell characterizes
white adipocytes. The functional cytoplasm and nucleoplasm are pushed to the periphery;
thus, the relative space occupied by the cytoplasm and nucleus shrinks as the cell stores more
lipids. Adipocytes can become extremely large at full development, reaching sizes greater
than almost any cell in the body. The diameter of completely expanded adipocytes is in excess
of 100 μm in most mammalian species and can be 200–500 μm in pigs and cattle.

2.1. Hyperplasia

The expansion of adipose tissue depots involves an increase in cell number coupled with an
increase in the mass of individual adipocytes. The increase in cell number is a major factor in
the increase of mammalian adipose tissue in young mammals (Allen, 1976). Ultimately, the
cell number sets limits on the absolute mass of adipose tissue because differentiated adipocytes
do not divide. The hyperplastic process is characteristic of the preadipocyte, the precursor cell
that has not differentiated and begun to fill with TG; it is controlled by numerous factors
including age, depot site, sex, and endocrine and growth factors (Hausman et al., 2001). In
most adult mammals, the direct contribution of hyperplasia to an increase in the mass of adipose
tissue is limited. A low rate of hyperplasia and differentiation is required for cell replacement
during the entire life of the organism. However, after the major increase in cell number early
in the life of a mammal, substantial hyperplastic rates are only measured, in most species,
when the organism is exposed to excessive caloric intake.
The increase in the size of an individual adipocyte by accumulation of lipid is limited.
When some percentage of the adipocytes reach the size limits for that species, hyperplasia is
increased to provide additional preadipocytes that can differentiate and fill with lipid to
expand the depot. This has clearly been demonstrated in rodents where several adipose tissue
depots (perirenal, perigonadal, inguinal) can be dissected in toto to enable determination of
the entire number of cells per depot. Also, DNA synthesis can be determined readily in these
small animals in individual depots. If the animals are fed excess energy for an extended
period, hyperplasia is increased and is readily demonstrated (DiGirolamo and Mendlinger,
1971; Greenwood and Hirsch, 1974; Miller et al., 1984; Shillabeer and Lau, 1994). Hyperplasia
and/or DNA synthesis are not readily quantified in larger mammals where the expense of
measuring DNA synthesis is considerable and the individual depots are extremely difficult to
remove quantitatively, limiting the capability for determination of cell number in the entire
depot (Gurr and Kirtland, 1978).
Attempts to measure total adipocyte number have been reported for depots in cattle and
sheep (e.g. Hood and Allen, 1973b; Robelin, 1981; Vernon, 1986) and pigs (e.g. Anderson et al.,
1972; Enser et al., 1976; Hood and Allen, 1977; Desnoyers et al., 1980; Hauser et al., 1997),
as well as subcutaneous adipose tissue from specific portions of thoracic rib sections in pigs
(Demaree et al., 2002) and cattle (Schiavetta et al., 1990). Also, DNA synthesis was demon-
strated in young pigs (Gurr et al., 1977; Hausman and Kauffman, 1986a), and in subcutaneous
adipose tissue explants from mature cattle (May et al., 1994). Measurement of cell number in
278 H. J. Mersmann and S. B. Smith

a portion of an adipose tissue depot may not reflect changes in cell number in the entire depot
because cell number in a small sample from a large depot, with expression of number per g tissue,
may reflect changes in cell size and not hyperplasia. Likewise, changes in DNA synthesis
in explants may not reflect hyperplasia in the entire depot.

2.2. Differentiation

Adipocytes develop from precursor cells, preadipocytes. Preadipocytes can be prepared from
digested adipose tissue and are obtainable from both young and older mammals. However,
the concentration of preadipocytes is greater in neonatal mammals because, at this time, there
is rapid cell division to provide preadipocytes to differentiate and fill with lipids to provide
expansion of the adipose tissue depots. The stromal–vascular cell fraction, isolated from
digested adipose tissue, contains numerous cell types, e.g. fibroblasts, reticulocytes, endothe-
lial cells, blood cells, and preadipocytes. When plated with the appropriate medium, most of
the attached cells appear as elongated, fibroblastic-like cells. Cell replication is rapid in the
proper culture medium with serum being an important component. There are no morphological
indications that any of these cells are preadipocytes. However, there are proteins characteristic
of the preadipocyte, e.g. Pref 1 (Gregoire et al., 1998; Gregoire, 2001), and antigenic materials
detectable in the preadipocyte that are specific to the adipocyte (Lee et al., 1986; Wright and
Hausman, 1990; Cryer et al., 1992; Yu et al., 1997). When the appropriate growth factors and
hormones are provided, many of the cells in the stromal–vascular fraction differentiate into
adipocytes, i.e. they begin to deposit large amounts of lipid. The amount of differentiation
seems to depend on the exact culture conditions. For example, if porcine stromal–vascular
cells are differentiated in serum, the extent of differentiation (the total number of differentiated
cells) is limited and differentiation occurs in clusters of cells. If the medium does not contain
serum, differentiation is more uniform across the culture plate, i.e. there are no clusters, and
the extent of differentiation is greater than in the presence of serum. In clonal lines of
preadipocytes, derived from rodents, the extent of differentiation is extensive (80% to >90%),
even when serum is present. Probably, there are species differences, as well as differences
between the clonal cells and the primary preadipocytes directly derived from the adipose
tissue stromal–vascular fraction.
Most of the concepts about adipocyte development come from study of clonal cells coupled
with a few studies of primary preadipocytes isolated from rodents. Primary preadipocytes that
differentiate in culture under the appropriate conditions have been prepared from many species
(Novakofski and Hu, 1987; Suryawan and Hu, 1995), including cattle (Plaas and Cryer, 1980;
Cryer et al., 1984; Aso et al., 1995; Ohyama et al., 1998; Torii et al., 1998; Peixing et al., 2000;
Wu et al., 2000), pigs (Hausman et al., 1984; Suryawan and Hu, 1993; Boone et al., 2000), rats
(Bjorntorp et al., 1980), and sheep (Broad and Ham, 1983; Vierck et al., 1996; Soret et al.,
1999; Arana et al., 2002), as well as humans (Hauner et al., 1989). The most extensive studies
using preadipocytes isolated from a domestic species are with porcine preadipocytes with an
overwhelming contribution by Hausman and coworkers.
The distinguishing morphological feature of a cell beginning to differentiate into an adipocyte
is the deposition of small lipid droplets. Seldom do normal cells deposit more than a very few
small lipid droplets. In areas where adipose tissue is developing, elongated, fibroblastic-like
cells can be observed with multiple small lipid droplets. As the cells accumulate more lipids,
the number of droplets increases and droplets fuse to form larger droplets. The adipocytes with
multiple lipid droplets are termed multilocular. Eventually the cell will contain a few large
lipid droplets and as lipid deposition continues, fusion of lipid droplets will lead to a cell with
Development of white adipose tissue lipid metabolism 279

a very large central lipid droplet, a unilocular adipocyte. Unilocular adipocytes also contain
many small lipid droplets confined to the peripheral cytoplasmic space. These stages of
adipocyte development are readily observed in the neonatal period, either before or after
birth, depending on the species (Napolitano, 1963; Slavin, 1985; Cinti, 2001); the pig has
been extensively studied (Mersmann et al., 1975; Hausman and Richardson, 1982; Hausman
and Kauffman, 1986b).
Coupling of cell culture systems and molecular biology techniques has led to a model for
differentiation of preadipocytes to adipocytes (fig. 1). The events that characterize the
differentiation from the totipotent stem cell to the multipotent mesenchymal precursor cell to
the committed preadipocyte remain largely unknown. However, given a cell that is committed
to become an adipocyte, i.e. a preadipocyte, a relatively clear series of sequential events seems
to occur. The exact timing of the events may vary with the species from which the preadipocyte
was derived and with the exact conditions used for cell culture. Preadipocytes grown in serum
without the appropriate adipogenic factors continue to multiply with essentially no differentia-
tion. When preadipocytes are presented with the proper stimuli, i.e. an appropriate combination
of growth factors and hormones, the cells begin to differentiate. The factors needed to initiate
differentiation vary with the cell type and/or the species, e.g. the clonal 3T3-L1 cells require
insulin, a glucocorticoid, and a cAMP-phosphodiesterase inhibitor, whereas the related clonal
cell, 3T3-F442A, requires only insulin. All cells require insulin, most require a glucocorti-
coid, many require a phosphodiesterase inhibitor, some require thyroid hormone, etc.
(Hausman et al, 1989, 1993; Ramsay et al., 1989; Cryer et al., 1992; Jump and MacDougald,
1993). Porcine primary preadipocytes do not require thyroid hormone or a phosphodiesterase
inhibitor; serum impedes the extent of differentiation but does not stop it (Hentges and
Hausman, 1989; Suryawan and Hu, 1993). Porcine preadipocyte differentiation is suppressed
by serum, but if oleic acid is added to the serum-containing medium, almost every cell in the
culture differentiates (McNeel and Mersmann, unpublished data). It is important to recognize
that the exact culture conditions dictate the extent of differentiation observed.
An initial differentiation event is the decreased expression of several genes characteristic
of the preadipocyte. This is followed by an increase in transcription factors that guide the
development of the adipocyte phenotype (fig. 1). There are multiple reviews of adipocyte

Fig. 1. Simplified model for adipocyte differentiation.


280 H. J. Mersmann and S. B. Smith

differentiation (e.g. Smas and Sul, 1995; Brun et al., 1996; Loftus and Lane, 1997; Fajas
et al., 1998; Morrison and Farmer, 2000; Ntambi and Young-Cheul, 2000; Rangwala and
Lazar, 2000; Gregoire, 2001). Early in differentiation, there is an increase in the transcription
factors CCAAT-enhancer binding proteins beta and delta (C/EBPβ and C/EBPδ). This is
followed by an increase in another transcription factor, peroxisome proliferator-activated
receptor gamma (PPARγ), and finally by an increase in C/EBPα. The active form of the
PPARγ transcription factor is as a heterodimer with retinoid x receptor alpha (RXRα). This
heterodimer must be activated by the binding of an appropriate ligand. Fatty acids are potential
ligands that stimulate differentiation (Schoonjans et al., 1996; Kliewer et al., 1997). The tran-
scription factor, adipocyte determination and differentiation-dependent factor 1 (ADD1),
plays a key role in inducing both PPARγ and fatty acid synthase. The provision of FAs by
synthase provides potential ligand for activation of PPARγ-RXRα (Kim and Spiegelman,
1996). The process of transformation to the adipocyte phenotype, a cell that is primarily
geared to synthesize and mobilize lipid, is guided primarily by PPARγ and C/EBPα (Castillo
et al., 1999; Lane et al., 1999; Lazar, 1999).
A number of genes that characterize the adipocyte phenotype have response elements in
their promoter region that bind either PPARγ or C/EBPα. Binding of the appropriate tran-
scription factor activates the transcription of that gene (fig. 1). In most cases, an increase in
the transcript for a gene, i.e. the mRNA, results in activation of the translation process with a
resultant increase in the protein. The proteins that characterize the adipocyte each have a
unique chronological pattern during differentiation, e.g. lipoprotein lipase (LPL) appears
early in differentiation, whereas adipocyte fatty acid binding protein (aP2) or glucose trans-
porter 4 (Glut 4) appear later (fig. 2). Clonal cells, which have been used to develop this
model, appear to differentiate more synchronously than primary cells so that the chronological
patterns may be more clearly observed with clonal cells. Also, there is evidence that primary
preadipocytes are somewhat further along in development at isolation than the clonal
preadipocytes. For example, porcine primary adipocytes have substantial concentrations of
both the mRNA (Ding et al., 1999) and the protein (Kim et al., 2000) for PPARγ, whereas this

Fig. 2. Development of adipocyte transcripts in dorsal subcutaneous adipose tissue obtained from the neck
region of milk-fed neonatal pigs. Data adapted from Ding et al. (1999) and McNeel et al. (2000).
Development of white adipose tissue lipid metabolism 281

transcription factor is undetectable or present at very low concentration in clonal preadipocytes


(Chawla et al., 1994; Tontonoz et al., 1994). The pattern of development of the mRNAs for
numerous adipocyte transcription factors and adipocyte-characteristic genes has been docu-
mented in porcine primary preadipocytes differentiating in vitro (Ding et al., 1999; McNeel
et al., 2000). Also, the pattern for development of the proteins for the key transcription
factors, PPARγ (Kim et al., 2000) and the C/EBPs (Lee et al., 1998; Yu and Hausman, 1998;
Chen et al., 1999), have been described for the differentiating porcine preadipocyte. It is
important to emphasize that the culture conditions may modify the chronological pattern
observed, even with clonal cells.
It is difficult to study adipocyte differentiation in vivo because a tissue sample contains
cells in many stages of differentiation. The pig offers a modestly reasonable model in vivo
because at birth, there are many preadipocytes present, essentially all adipocytes are multi-
locular, and there is rapid transformation of many cells to the unilocular stage over a very few
days. The tissue is undoubtedly more heterogeneous than clonal preadipocytes, or even primary
preadipocytes differentiating in culture. However, development of several transcripts that
characterize adipocyte differentiation or the adipocyte phenotype (Ding et al., 1999; McNeel
et al., 2000), follow a pattern that is approximately the same as differentiation of porcine
preadipocytes in culture (fig. 2).

2.3. Hypertrophy

After differentiation of the adipocyte, most of the growth of the cell and, consequently, the
tissue is by hypertrophy, which is the lipid filling process. Thus, the average adipocyte size in
a given depot increases as the mammal grows, as indicated for ruminants (Hood and Allen,
1973b; Allen, 1976; Hood and Thornton, 1979; Hood, 1982; Robelin, 1986; Vernon, 1986)
and pigs (Anderson and Kauffman, 1973; Mersmann et al., 1973c; Hood and Allen, 1977;
Desnoyers et al., 1980; Hausman, 1985; Hauser et al., 1997). Adipocytes have flattened sides
in vivo, but the shape approximates a sphere. Thus, the diameter is a reasonable expression of
the cell size. It must be recognized that as the diameter doubles (e.g. from 20 to 40 μm), the
volume increases 8 times (from 4200 μm3 to 33,500 μm3). Measurement of average cell size
with extrapolation to calculate the number of cells packaged in a defined depot, e.g. in a
rodent depot, is not unreasonable. However, extrapolation of the average cell size from a
small sample obtained from a large mammal is probably not very valid because it is difficult
to quantitatively dissect the depot. This is particularly applicable to the generally contiguous
subcutaneous depot and the mesenteric depot, scattered within the entire gut mesentery.
Furthermore, a single size determination on a small sample from a large depot is probably
invalid because the cell size is not expected to be uniform over the entire depot.
The composition of adipose tissue changes with growth (fig. 3). Most of the changes are
attributable to the extensive accumulation of intracellular TG resulting in an increase in
adipocyte size. Consequently, the TG expressed per g tissue increases with adipose tissue
growth. In species of mammals with a marked increase in adipocyte size, the TG concentration
easily reaches 700 mg per g tissue and may approach 900 mg per g tissue. Because the indi-
vidual cells are increasing in size, the number of cells per g tissue is concomitantly decreasing.
Consequently, the DNA and protein expressed per g tissue also decreases. The exact chrono-
logical pattern for these changes will depend on the species, the capacity of the adipocyte to fill
with lipid, the nutritional status of the animals, and the particular tissue components measured.
To a large extent, adipose tissue growth is governed by the caloric intake of the animal. Excess
energy is deposited as fat. Thus, limits on energy intake limit the growth of adipose tissue.
282 H. J. Mersmann and S. B. Smith

Fig. 3. Composition of porcine dorsal subcutaneous adipose tissue obtained from the neck region of suck-
ling pigs until day 20, and from pigs weaned at day 21 and fed a low-fat grain-based diet. The 100% values
are: triacylglycerol (TG) = 900 mg/g tissue; DNA = 1.9 ng DNA phosphorus/g tissue; protein = 71.6 mg/g
tissue; collagen = 35.8 mg/g tissue. Data adapted from Mersmann et al. (1973c).

This is primarily the result of limits on adipocyte hypertrophy, but as indicated earlier, when
caloric intake is elevated chronically, there is usually an increase in cell hyperplasia.
Recent study of adipocyte biology indicates that the adipocyte is an endocrine cell export-
ing numerous peptides that have the potential to modify the biological function of the
adipocyte and other tissues (Hwang et al., 1997; Kim and Moustaid-Moussa, 2000; Fruhbeck
et al., 2001; Trayhurn and Beattie, 2001). The first factor discovered was leptin, a cytokine-
like peptide that is made and secreted by the adipocyte. There are leptin receptors in the
hypothalamus that regulate feeding behavior. Thus, as adipocytes grow and the amount of adi-
pose tissue increases, the amount of leptin secreted to the blood increases. The increased
leptin signals the brain to decrease feed intake. There are leptin receptors in other tissues,
including adipocytes themselves. Leptin decreases adipocyte lipid anabolism and increases
lipid catabolism in porcine adipocytes (Ramsay, 2001), but not in ovine adipocytes (Newby
et al., 2001).

3. DEVELOPMENT OF LIPID SYNTHESIS


The overall pathways for adipocyte lipid synthesis and degradation were understood decades
ago (Bauman, 1976); a simplified scheme is indicated in fig. 4. Glucose or acetate are
obtained from the plasma, transformed to acetyl-CoA, then synthesized to FAs. The degrada-
tion of lipoproteins by lipoprotein lipase (LPL) provides an additional source of FAs. The FAs
are esterified to TG, the major storage lipid. Degradation of TG occurs by lipolysis, yielding
3 FAs and glycerol. The overall development of these pathways (fig. 5) indicates that LPL
activity and TG synthesis develop rapidly after birth in the pig during the suckling period
when fat deposition is rapid using milk fat as a source of FAs. The degradative activity also
develops rapidly after birth to provide energy during times of environmental stress or infre-
quent feeding. De novo synthesis of FAs does not substantially increase until after weaning.
Fig. 4. Simplified depiction of adipocyte lipid anabolic and catabolic lipid metabolism. Abbreviations used are: AcCC = acetyl-CoA carboxylase; β-oxidation = mitochondrial oxidation
of fatty acids; FA = long-chain fatty acid; FAS = fatty acid synthase; lipolysis = sequential degradation of TG to 3 FAs + glycerol; LPL = lipoprotein lipase; OAA = oxaloacetate; P = phosphate;
TCA cycle = mitochondrial tricarboxylic acid cycle; TG = triacylglycerol.
284 H. J. Mersmann and S. B. Smith

Fig. 5. Ontogeny of porcine dorsal subcutaneous adipose tissue lipid metabolism. Weaning was at 21 days
postpartum with pigs fed a low-fat diet after weaning. The 100% data are: LPL (lipoprotein lipase) = 6.6 μmol
fatty acid released/60 min/106 cells; FA (de novo fatty acid synthesis from glucose) = 661 nmol glucose incor-
porated into total lipid/60 min/106 cells; TG (triacylglycerol + diacylglycerol + phospholipid synthesis from
glycerol 3-phosphate) = 1.9 μmol glycerophosphate incorporated/60 min/106 cells; lipolysis = 285 μmol fatty
acid released/60 min/106cells. Data adapted from: LPL (Steffen et al., 1978); FA (Mersmann et al., 1973c);
TG (Steffen et al., 1979); lipolysis (Mersmann et al., 1976).

Lipid and glucose metabolism in various tissues undergo multiple adaptations during the peri-
natal, suckling, and weaning periods (Girard et al., 1992), as well as during postweaning growth
(Saggerson, 1985; Vernon, 1992). There are extensive reviews of the early literature about rumi-
nant (Vernon, 1980, 1981; Noble, 1981) and porcine (Mersmann, 1986; Farnsworth and Kramer,
1987) adipocyte lipid metabolism, including developmental aspects. Adipose tissue lipid metab-
olism, including influences such as genetics, temperature, or diet on its development, continues
to be of interest (Bass et al., 1990; Wood, 1990; Chilliard, 1993) with emphasis on pigs
(Le Dividich et al., 1994; Mourot et al., 1995, 1996; Camara et al., 1996; Boone et al., 1999;
Gerfault et al., 2000; McNeel and Mersmann, 2000; Robert et al., 2000), cattle (Mendizabal
et al., 1999), goats (Bas, 1992), and lambs (e.g. Vernon, 1982; Mendizabal et al., 1997; Purroy
et al., 1997; Soret et al., 1998; Payne, 1999; Greathead et al., 2001).

3.1. Sources of fatty acids

The supply of FAs to the organism is from the diet and from de novo FA synthesis. The direct
source of FAs for the adipocyte is lipoproteins circulating in the blood plasma, and in some
species, from de novo synthesis in the adipocyte. Developmental aspects of plasma lipoproteins
in cattle and pigs have been reviewed (Chapman and Forgez, 1985). The major lipoprotein
sources of FAs for the adipocyte are chylomicrons, the very large, heavily lipid-laden particles
produced in the intestine after ingestion of a fatty meal and the very low-density lipoproteins
(VLDL). Ruminant species do not produce chylomicrons per se due to their low level of fat
intake. Rather, they export VLDL (associated with apolipoprotein B48) from the intestinal
mucosal cells (see Chapter 13 by Drackley).
Development of white adipose tissue lipid metabolism 285

Chylomicrons and VLDL have a high concentration of TG, but the adipocyte does not
directly absorb TG. The TG is cleaved at the endothelial cell surface by the adipocyte-
produced enzyme, LPL, and the products of this reaction, the FAs and 2-monoacylglycerols,
can be moved to the adipocyte and absorbed. Fatty acids are quickly processed in the cell by
a variety of mechanisms because the nonesterified and unbound FA is a detergent and poten-
tially quite toxic to the cell. Formation of the coenzyme A thioester is one pathway for further
processing of FAs. Once the CoA derivative is formed (acyl-CoA), the FA can be oxidized or
used for various biosynthetic purposes, including synthesis of phospholipids, cholesterol
esters, and TG. Nonesterified FAs are bound to several proteins that have specific binding
sites for FAs. In the blood plasma, almost all of the nonesterified FA is bound to albumin,
whereas in the cell, there are specific FA binding proteins, among which is adipocyte fatty
acid binding protein (aP2). Fatty acid binding has been demonstrated in adipose tissues of
pigs, sheep, and cattle (St. John et al., 1987; Coleman et al., 1988; Miller et al., 1988). In
cattle and sheep, depression of de novo fatty acid synthesis is accompanied by depressions in
fatty acid binding protein activity (Coleman et al., 1988; Miller et al., 1988).

3.2. Synthesis of FA (fig. 4)

Synthesis of FAs from nonlipid sources occurs in the adipocyte of many, but not all, mam-
malian species (Vernon and Clegg, 1985; Vernon and Taylor, 1986). Glucose or closely
related sugars derived from carbohydrates are the usual precursors of FA carbon in nonrumi-
nant species. Lactate and the carbon skeletons derived from nonessential amino acids by
transamination may each be important lipogenic precursors, under some circumstances. The
ultimate donor molecule for FA synthesis is the two-carbon donor, acetyl-CoA. Glycolytic
metabolism of glucose to pyruvate is followed by pyruvate entry into the mitochondrion with
subsequent decarboxylation to acetyl-CoA. The mitochondrial acetyl-CoA is not used for FA
synthesis because the enzymatic machinery for FA synthesis is located in the cytosol. The
acetyl-CoA does not traverse the mitochondrial membrane, but is coupled with the four-
carbon product of the tricarboxylic acid cycle, oxaloacetate, to yield the six-carbon, citrate.
Citrate can traverse the mitochondrial membrane and in the cytosol is converted back to
acetyl-CoA plus oxaloacetate by an enzyme, ATP-citrate lyase (citrate cleavage enzyme in the
older literature).
Ruminant species generally use glucose sparingly as a FA precursor because metabolism
in the rumen limits the glucose supply to the animal. Ruminants use acetate as the primary
carbon precursor for FA synthesis. De novo FA synthesis is initiated by the carboxylation of
cytosolic acetyl-CoA by acetyl-CoA carboxylase, the tightly regulated and many times rate-
limiting enzyme for FA synthesis. The three-carbon product of this reaction, malonyl-CoA, is
then the substrate for fatty acid synthase, the enzyme complex that polymerizes two carbon
moieties into long-chain FAs. One carbon is removed from malonyl-CoA during the poly-
merization process. Acetyl-CoA carboxylase limits the rate of de novo fatty acid biosynthesis
from glucose in porcine adipose tissue and from acetate in bovine adipose tissue. However,
glucose incorporation into fatty acids in bovine adipose tissue apparently is limited by the
glycolytic pathway, probably at 6-phosphofructokinase (Smith, 1984). Ruminant adipocyte
substrate utilization has been reviewed (Bauman, 1976; Smith, 1995).
The major product of de novo FA synthesis is the 16-carbon saturated FA, palmitic acid
(C16:0). (Fatty acids are designated as the number of carbons in the FA chain; C16 = 16 carbons,
with the number of double bonds indicated after the colon. This simple nomenclature does
not indicate the location of the double bonds.) The adipocyte can extend C16:0 by adding
286 H. J. Mersmann and S. B. Smith

two-carbon moieties in a chain-elongation reaction, for example, to yield the 18-carbon


saturated FA, stearic acid (C18:0). Double bonds can be inserted into the saturated FAs to pro-
duce a series of unsaturated FAs. The positions for desaturation are limited in mammals so
that certain FAs are essential and must be obtained from the diet. The enzyme stearoyl-CoA
desaturase uses C18:0 as a substrate to produce the monounsaturated FA, oleic acid (C18:1),
with one double bond between the 9 and 10 carbon atoms. This enzyme can also transform
other saturated fatty acids to their 9 monounsaturated counterparts (Yang et al., 1999).
One of the essential FAs, obtained from the diet and ultimately from plant products, is the
18-carbon FA with two double bonds (cis-9, cis-12), linoleic acid (C18:2). The cis-12 double
bond is at the number 12 carbon counting from the carboxyl group, but at the 6 carbon count-
ing from the methyl group of the FA. Thus, C18:2 is the primary member of a series of FAs
called n-6 or omega-6 FAs. Chain elongation and further desaturation of C18:2 gives rise to
arachidonic acid (C20:4). The C20:4 is an important constituent of some membranes and a
precursor for many eicosanoid molecules, including prostaglandins, leukotrienes, and throm-
boxanes. The other essential FA is the 18-carbon FA with three double bonds, α-linolenic acid
(C18:3). This FA is the primary member of the n-3 series of FAs; it has double bonds at the
9, 12, and 15 carbons (counting from the carboxyl group) or n-3, 6, and 9 counting from the
omega or methyl carbon. Chain elongation and desaturation of this FA produces several
eicosanoid molecules plus eicosapentaenoic acid (C20:5) and docosahexaenoic acid (C22:6),
both of which are important FA constituents of some membranes, particularly in the mam-
malian central nervous system. Most of these fatty acids are constituents of animal, plant, and
microorganism lipids; thus the diet, regardless of its composition, is a major source of long-
chain FAs. Individual dietary fats are specifically enriched with individual FAs: C16:0 and
C18:0 in mammalian meat and milk products, C18:1 in olive oil or some canola oils, C18:2
in corn oil or safflower oil, C18:3 in flaxseed or linseed oil, and C20:5 and C22:6 in certain
oils from fatty fishes.

3.3. Regulation of FA synthesis

The development of adipocyte de novo FA synthesis has been documented in ruminant and
nonruminant mammals. De novo FA synthesis is reciprocally regulated to accommodate the
dietary supply of FAs. The measured rate of placental transport of FAs is limited (Leat and
Harrison, 1980; Thulin et al., 1989). However, because transport is a continuous process over
the extended period of gestation and fetal de novo FA synthesis rates are low, in utero, FAs
are probably supplied primarily by the dam. After birth, the milk supplied to the suckling
mammal has a relatively high fat concentration so that de novo FA synthesis is again not
needed. After weaning, the extent of de novo FA synthesis is dictated by the synthetic and
oxidative requirements for FAs coupled with the FA supply from the diet. Most mammals
raised for meat production or for biomedical research are fed a relatively low-fat postwean-
ing diet so that de novo FA synthesis is increased at this time. Decreased feed intake and
fasting markedly decrease de novo FA synthesis (Mersmann et al., 1981); the timing of the
decrease during fasting is probably dependent on the gut transit time.
Within days of transformation from the high-fat suckling diet to the low-fat postweaning
diet, porcine adipocytes develop increased capacity to synthesize FAs de novo (Mersmann
et al., 1973a,b). Enzyme activities associated with de novo FA synthesis, e.g. ATP citrate
lyase, acetyl-CoA carboxylase, fatty acid synthase, and cytosolic enzymes that produce
reducing equivalents for the biosynthetic process, e.g. glucose-6 phosphate dehydrogenase,
NADP-linked malic enzyme, and NADP-linked isocitrate dehydrogenase, are all increased in
Development of white adipose tissue lipid metabolism 287

Fig. 6. Expression of adipocyte enzyme or metabolic data. Lipoprotein lipase activity (LPL) and glyc-
erophosphate acyltransferase activity (GPA) expressed per g adipose tissue or per adipocyte. The maximal
activity is indicated as 100%. Data are adapted from: LPL (Steffen et al., 1978) and GPA (Steffen et al., 1979).

the adipocyte. In cattle, there can be a considerable delay between weaning and the expres-
sion of ATP-citrate lyase and acetyl-CoA carboxylase (Smith et al., 1984; Martin et al., 1999).
The manner in which the carbon flux data or the enzyme activities are expressed is impor-
tant. As the adipocyte increases in size, there is increased capacity for de novo synthesis of
FAs. Thus, depending on the exact chronology of the increase in adipocyte size, expression
of these metabolic activities on a per g tissue basis can be misleading because the number of
cells per g tissue is decreasing. At some point, the metabolic activity per g tissue will
decrease, when in actuality the activity per adipocyte is increasing (fig. 6). It is important to
express adipocyte metabolic activity per adipocyte to understand the biology of the adipocyte.
To understand the biology of the animal, it would be desirable to know the activity for the
entire adipose tissue depot and ultimately for the entire animal. Total depot activity is most
readily obtained from small laboratory mammals and for selected depots, e.g. the peri-
gonadal, the perirenal, or the inguinal fat depots. Each of these depots can be dissected in toto
and because the size is not great, measurement of metabolic activity in a sample of the depot
can be extrapolated to the entire depot. Extrapolation is difficult in larger mammals because
dissection of an entire depot is usually difficult, and because of the size of the depot,
adipocyte development is not uniform across the depot. This is particularly true for the sub-
cutaneous depot that is prevalent in most agricultural species, and for the mesenteric depot
that is prominent in ruminant species. Modes of expression of adipose tissue metabolic data
are discussed for development of porcine adipose tissue (Hood and Allen, 1973a; Mersmann
and Brown, 1973).
Many mammals have both a hepatic and an adipocyte capacity for de novo FA biosynthesis,
e.g. rats, cattle, and sheep. Because there is little depot fat in the newborn of most mammalian
species, the predominant site of FA synthesis is initially the liver. As the organism develops,
288 H. J. Mersmann and S. B. Smith

with a continuous increase in adipose tissue mass, the adipocytes become more important
for total body FA synthesis. In the pig, there is little or no capacity for hepatic de novo FA
synthesis at any age so that the adipocyte is the primary tissue source of FAs. In the human
(and the chicken) the reverse is true, i.e. there is little adipocyte capacity for de novo synthesis
of long-chain FAs.
The extent of de novo FA synthesis is inversely regulated by the dietary fat concentration,
so that as the fat is increased in the diet, there is a progressive decrease in de novo FA syn-
thesis. Thus, when nonruminant mammals are fed a particular fat source in the diet, the FA
composition of the fat depots will more or less reflect the composition of the diet. In rumi-
nants, the FAs emerging from the rumen are highly saturated and it is difficult to substantially
increase the unsaturated FA concentration of adipose tissue depots (Rule et al., 1995).

3.4. Lipoprotein lipase

The alternative source of adipocyte FAs is from hydrolysis of plasma lipoprotein TG.
Lipoprotein lipase is synthesized by the adipocyte, and migrates to the surface of the adjacent
capillary endothelial cell where it hydrolyzes lipoprotein to release FAs. Triacylglycerol
cannot be transported into the adipocyte, but the FAs can. Adipocyte LPL is regulated such
that it is active in the fed animal to supply the adipocyte with FAs; the LPL activity decreases
in the fasted animal. Insulin is the primary mediator of this regulation of adipocyte LPL. In
the newborn pig there is limited adipocyte LPL activity, but this increases rapidly so that after
several days there is adequate LPL to hydrolyze the lipoproteins synthesized from the lipids
provided by the high-fat milk diet (fig. 5). There is a tendency for the LPL activity to be
greater in adipose tissue from animals fed high-fat diets compared to low-fat diets, so that in
pigs fed grain diets after weaning, the LPL activity is lower than before weaning. Skeletal
muscle LPL generally is regulated in an opposite direction to adipocyte LPL; the muscle
enzyme is increased during fasting to increase the supply of FAs to the muscle for oxidation.
Measurement of LPL activity, LPL protein, and LPL mRNA in the same animal allows inter-
pretation of the regulatory events controlling LPL function (Tavangar et al., 1992). This type
of study has not been done in agricultural species, but enzyme activity or the mRNA have
been documented in cattle (de la Hoz and Vernon, 1996; Hocquette et al., 2001), pigs (Steffen
et al., 1978; McNeel and Mersmann, 2000), and sheep (Andersen et al., 1996; Bonnet et al.,
1998, 2000).

3.5. Triacylglycerol synthesis (fig. 4)

As previously indicated, cells cannot accumulate large amounts of nonesterified FAs. The
adipocyte may be considered a sink or reservoir for FA storage with the TG molecule being
the major storage molecule. Synthesis of TG is a sequential process leading to esterification
of long-chain FAs to each of the three hydroxyl groups of glycerol. The glycerol molecule,
serving as the backbone for triacylglycerol synthesis, as well as synthesis of phospholipid
molecules, is derived from glucose via glycolysis. The six-carbon sugar derivative is split into
two three-carbon moieties, with one of these, dihydroxyacetone phosphate, providing the
initial source of carbons for the glyceride–glycerol moiety. Dihydroxyacetone phosphate
is hydrogenated to glycerol-3 phosphate by the enzyme glycerophosphate dehydrogenase.
(The activity or mRNA concentration of this enzyme is often used as an indicator for
adipocyte differentiation.) Two acyl groups are sequentially added to the glycerophosphate by
Development of white adipose tissue lipid metabolism 289

the enzyme, α-glycerophosphate acyltransferase. The resulting product is phosphatidic acid.


The phosphatidic acid is dephosphorylated by the enzyme, phosphatidate phosphohydrolase,
to yield diacylglycerol. Finally, the diacylglycerol is acylated to yield the triacylglycerol
molecule representing the major storage molecule in the adipocyte. Both phosphatidic acid
and diacylglycerol are also precursors of phospholipids, so the regulation of TG synthesis is
partly regulated by the demand for the intermediates. Phosphatidate phosphohydrolase is
generally considered to be a key regulatory enzyme for these pathways, but diacylglycerol
acyltransferase has recently emerged as an enzymatic site for regulation.
If FAs are supplied to the adipocyte, there must be adequate TG synthesis capacity to esterify
the FAs. In a mammal like the pig, with an extremely rapid postnatal increase in adipocyte
hypertrophy, there is an early postnatal demand for esterification of FAs (fig. 5). Enzymes or
other aspects of this pathway have seldom been measured in agricultural species (Rule, 1995);
however, there are reports for cattle (Lin et al., 1992; Wilson et al., 1992; Smith et al., 1998),
pigs (see Mersmann, 1986; Rule et al., 1988a,b, 1989), and lambs (Andersen et al., 1996).

3.6. Endocrine regulation of anabolic metabolism

Regulation of mammalian adipocyte lipid anabolic processes is primarily via adrenergic and
insulin receptors (Etherton and Walton, 1986; Mersmann, 1991; table 1). There are two major
types of adrenergic receptors, the α- and β-adrenergic receptors (αAR and βAR, respec-
tively). In many situations, stimulation of αAR produces actions opposing stimulation of
βAR. The insulin receptor tends to work in opposition to the βAR to provide the major
adipocyte regulatory system. Thus, insulin stimulates the adipocyte anabolic lipid metabolism
pathways and βAR agonists inhibit these same pathways. Rodent adipocytes are particularly
sensitive to insulin stimulation of anabolism in vitro, with rates of de novo FA synthesis being
increased 10 or more times by insulin. Anabolic processes in adipocytes from many other
mammalian species, including pigs (Romsos et al., 1971; Etherton and Chung, 1981; Walton
and Etherton, 1986, 1987; Mersmann and Hu, 1987; Mersmann, 1989a; Budd et al., 1994; Mills,
1999), are stimulated by insulin, but the magnitude is much less than in rodent adipocytes.
The bovine adipocyte is relatively insensitive to insulin (Smith et al., 1983; Vasilatos et al.,
1983). After a meal, when insulin is elevated, it stimulates lipid synthesis, whereas after fasting
the insulin concentrations are lowered to cause decreased lipid synthesis. The βAR agonists
decrease anabolic adipocyte lipid metabolism (Rule et al., 1987; Coleman et al., 1988;
Mersmann, 1989b, 1995, 1998, 2002a; Miller et al., 1989; Mills et al., 1990; Etherton and Smith,
1991; Budd et al., 1994; Moody et al., 2000; Bergen, 2001).
Adipocyte anabolism may also be controlled by other hormones. Decreased thyroid func-
tion leads to increased fat accumulation, as does increased adrenocorticoid production. In
contrast, increased somatotropin leads to decreased adipocyte anabolic lipid metabolism
(Vernon, 1991; Vernon et al., 1991; Etherton and Louveau, 1992; Harris et al., 1993; Etherton
et al., 1995 Bergen, 2001). Estrogens and androgens have effects on adipocyte function as well,
leading to the increased fat deposition in females compared to males, and in some species, sex-
dependent differential sites of fat deposition. Most of these additional endocrine regulations do
not enter into the day-to-day control of adipocyte lipid metabolism. Rather they provide tonic
control or long-term modulation. It should be noted that exogenous sex steroids (Hancock
et al., 1991), somatotropin (Sejrsen et al., 1989; Beermann and DeVol, 1991; Beermann, 1994),
and selected βAR agonists (Moloney et al., 1991) have been used to modify animal growth and
carcass composition (NRC, 1994; Steele, 1991; Steele et al., 1994).
290 H. J. Mersmann and S. B. Smith

3.7. Conjugated linoleic acid

There is recent evidence that administration of exogenous conjugated linoleic acid can reduce
carcass fat deposition in mice, rats, and pigs. The mechanism(s) is not clear, but there is evidence
for increased energy expenditure, increased fat oxidation, decreased preadipocyte prolifera-
tion and differentiation, reduced FA synthesis, reduced LPL activity, increased lipolysis
(triacylglycerol degradation), and reduced monounsaturated FA production because of
reduced stearoyl-CoA reductase activity (Mersmann, 2002b).

4. DEVELOPMENT OF LIPID DEGRADATION


The process of adipocyte lipid degradation, or lipolysis, is initiated by stimulation of a mem-
brane-bound receptor (Belfrage, 1985). Typically this would be a β-adrenergic receptor
(βAR) that is coupled to a Gs protein that is coupled to adenylyl cyclase (table 1). Thus, the
sequential activation of the receptor, the Gs protein, and adenylyl cyclase yields increased
synthesis of cAMP from ATP. An increase in intracellular cAMP leads to activation of pro-
tein kinase A by attachment of cAMP to the regulatory subunit of the kinase with subsequent
cleavage of the catalytic subunit. The kinase catalytic subunit, freed from the regulatory sub-
unit, then phosphorylates hormone-sensitive lipase to activate it. Hormone-sensitive lipase is
the rate-limiting enzyme in the process of lipolysis and for the most part is in the nonphos-
phorylated state until activated by protein kinase A. Hormone-sensitive lipase cleaves the first
two FAs from the TG substrate, whereas the last FA is cleaved by another enzyme, mono-
acylglycerol lipase. The latter lipase is continuously active and does not participate in the
regulation of lipolysis. The products of lipolysis are the three FAs that were esterified to TG
plus glycerol. The FAs may be transported to the plasma or re-enter the adipocyte intra-
cellular FA pool where they may be utilized for oxidation or esterification to form complex
lipid esters, including the synthesis of TG (fig. 4). Glycerol is not recycled in the adipocyte
because glycerol kinase, the enzyme responsible for phosphorylation of glycerol and its
re-entry into the pathway for synthesis of TG, is present at extremely low concentration in
adipocytes.
Typically the lipolytic rate in nonstimulated adipocytes is low, but after maximal stimula-
tion with a βAR agonist, the rate is increased several-fold. The extent of stimulation depends
on the species, the age of the animal, and the nutritional status. Many of these aspects of
the regulation lipolysis have been reviewed (Mersmann, 1990, 1991; Chilliard et al., 2000).

Table 1
Major endocrine effects on adipocyte lipid metabolism a

Anabolicb Catabolic

Hormone FA LPL TG Lipolysis

Insulin    
β-adrenergic    
GH    
Glucocorticoids    
a These effects are more or less evident, depending on the species. Broken arrows indicate effects are marginal or

reports are mixed regarding the effect.


b Abbreviations: FA = de novo fatty acid synthesis; LPL = lipoprotein lipase; TG = triacyglycerol synthesis.
Development of white adipose tissue lipid metabolism 291

For example, in adipocytes from young rats, lipolysis is stimulated as much as 6 to 10 times,
whereas in adipocytes from older rats the increase may be only 3 to 6 times. The degree of
lipolytic stimulation by a βAR agonist tends to be greater in rat adipocytes than in adipocytes
from cattle, chickens, or pigs, wherein the maximal rate is usually only 2 to 5 times that of
the unstimulated rate. The lipolytic rate is increased in the fasted compared to the fed state in
most species to provide substrate, i.e. FAs for oxidative metabolism.
There are three subtypes of βAR: β1AR, β2AR, and β3AR. These are three distinct proteins,
coded by three different genes. Within a mammalian species the homology of the three sub-
types is approximately 50%, whereas across species the homology for a single subtype
is ≥75%. Species-specific differences in the protein structure of a βAR subtype can lead to
major differences in the function of the receptor and in the potency and efficacy of agonists
and antagonists for the receptor. Thus, an antagonist that is specific for the rat β2AR, ICI
118,551, is not specific for the porcine β2AR, and a specific agonist for the rat β3AR, BRL
37,344, is not specific for the porcine β3AR, but is specific for the porcine β2AR, where it acts
as an antagonist (Liang and Mills, 2002). Propranolol is an antagonist for the cloned mouse
β3AR, but a partial agonist for the cloned bovine and human β3AR (Piétri-Rouxel et al.,
1995). Thus, it cannot be assumed that the βAR subtype specificity of an agonist or antago-
nist in one species will be maintained in another species; the specificity must be tested using
cloned receptors from the individual species of interest (Mills and Mersmann, 1995;
Mersmann, 2002a).
The βAR subtypes are differentially distributed in the various tissues of a particular
species, e.g. rat heart has >90% β1AR, rat lung has >85% β2AR, and rat adipocytes
have >90% β3AR. Receptor distribution in a particular tissue also varies across species, e.g.
there are approximately 65% β1AR in porcine heart compared to the 90% in rat heart and
there are approximately 75% β1AR in porcine adipocytes compared to the 90% β3AR in rat
adipocytes (McNeel and Mersmann, 1999; Liang and Mills, 2002). In a few cases, there is
evidence for a change in adrenergic receptors in a tissue during development. For example,
in rat adipocytes the α2AR increased 4-fold between 6 and 20 weeks of age, whereas the βAR
decreased 25% in the same time span (Kobatake et al., 1991). In human adipocytes isolated
from infants <2 months old, there was more α2AR activity than in adipocytes isolated from
adults (Marcus et al., 1987). Also, the rodent preadipocyte has few or no β3AR with a shift
to ≥90% β3AR during differentiation to adipocytes (Feve et al., 1991). Ultimately the response
of the adipocyte to adrenergic stimulation, in a particular species, will depend on the βAR
subtypes present, the ratio of α2AR to βAR on the adipocyte (stimulation of α2ARs inhibits
lipolysis), the stage of development of the adipocyte (possibly leading to shifts in receptor
subtypes), and the concentration of norepinephrine and epinephrine, the physiological
agonists for βAR, to which the adipocyte is exposed (Mersmann, 1998, 2002a).
Depending on the species, there are other membrane-bound receptors that may regulate
hormone-sensitive lipase through the cAMP-protein kinase A system. Rat adipocytes have
receptors for adrenocorticotropin, glucagon, somatotropin, thyrotropin, etc. Stimulation of
any of these receptors increases the lipolytic rate. For the most part, these receptors appear to
be minor factors in the regulation of lipolysis in most nonrodent species (Mersmann, 1990,
1991; Lanna and Bauman, 1999).
Lipolysis is regulated, not only by receptors that stimulate the process, but also by
inhibitory receptors. Stimulation of the α2-adrenergic receptor (α2AR) inhibits lipolysis
because the α2AR is coupled to the Gi protein, an inhibitory factor. The physiological adrener-
gic hormone, epinephrine, has both βAR and αAR activity. Thus, the effect of epinephrine on
adipocyte lipolysis depends on the relative populations of α2AR and βAR on the adipocyte.
292 H. J. Mersmann and S. B. Smith

There are few α2ARs on adipocytes from some species, e.g. rats and pigs, so that this mech-
anism of inhibition is not operative (Mersmann, 1990). Perhaps the most important
physiological negative modulator for lipolysis is insulin (table 1). Increased insulin inhibits
lipolysis (Mersmann and Hu, 1987; Mersmann, 1990). Thus, after a meal when insulin con-
centration is elevated, lipolysis is decreased. At least part of the mechanism for insulin
inhibition of lipolysis is stimulation of the activity of cAMP-phosphodiesterase to decrease
the concentration of intracellular cAMP, and consequently decrease the activation of lipolysis
through protein kinase A and hormone-sensitive lipase. Another negative control of lipolysis
is provided by the adenosine receptor (A1R). This receptor also couples to Gi proteins to
inhibit lipolysis. Adenosine, the agonist, is produced extracellularly by adipocytes from
cAMP. The process is very active and to measure substantial rates of lipolysis with adipocytes
isolated from many species, it is necessary to either inhibit the A1R or to destroy the adenosine
by addition of adenosine deaminase. The extent of adenosine formation in vivo is not yet
clear, but for many adipocyte preparations in vitro, inhibition of adenosine formation leads to
a considerable enhancement of the lipolytic rate (Carey, 1995).
Although subject to several qualifications, the lipolytic activity may be approximated
in vivo by changes in the plasma FA and/or glycerol concentration. The assumption is that the
source of most of the nonesterified FA and glycerol in the plasma is adipocyte lipolysis. These
types of measurements can estimate the extent of stimulation or inhibition by an acutely
infused compound, but do not reflect the actual rate. They also can be used to estimate a
dose-response for the compound upon infusion in vivo. Likewise, the nonesterified FA con-
centration in the plasma can be used to assess the effects of chronic treatment with a particular
compound on lipolytic activity in vivo. Such approaches have been used to estimate the
response of adipose tissue to various βAR and αAR agonists and antagonists, and other meta-
bolic hormones in cattle, pigs, and sheep (Mersmann, 1987, 1989c, 1995, 1998, 2002a).
Function of the lipolytic process in adipocytes from newborn or young mammals is
impaired relative to function in mature adipocytes. Thus, the βAR-stimulated rate increases
as adipocytes increase in size. Lipolysis is impaired in adipocytes isolated from newborn pigs,
but increases several-fold within the first few days of postnatal life (fig. 5; Mersmann, 1986).
The ability to mobilize FAs immediately after birth is species-specific and depends on the
amount of stored adipose tissue, the chronological development of the enzymatic machinery
for lipolysis, the establishment of the appropriate receptor populations, and the development
of coupled receptor-driven metabolism.
The development of both anabolic and catabolic adipocyte lipid metabolism primarily
occurs after birth in many mammalian species, including cattle, pigs, and sheep. The minimal
capacity to mobilize FAs after birth is critical in individuals exposed to cold or that do not
have sufficient milk intake.

5. FUTURE PERSPECTIVES
Although the goal in modern meat animal production is to produce animal protein with min-
imal fat content, the animal raised under conditions where the environment is not optimal and
constant requires fat depots for insulation, to provide oxidative substrates, and to produce
selected endocrine materials and growth factors. Mechanisms to decrease adipose tissue lipid
metabolism anabolic processes and to increase catabolic processes have been sought.
Somatotropin and βAR agonists are two exogenous compounds that mechanistically operate
in this fashion. However, in the neonate, the goal cannot be to decrease the anabolic activity
or increase the catabolic activity because the neonate is in a precarious metabolic state with
Development of white adipose tissue lipid metabolism 293

limited capacity to mobilize fatty acids for oxidative fuels or to synthesize fatty acids for
esterification to triacylglycerol. In fact, it might be appropriate to seek to temporarily increase
the rate of fatty acid biosynthesis and lipolysis in neonates. Ideally, if selected genes can be
regulated, i.e., turned on and off, it might be possible to enhance selected gene expression at
particular stages of growth and to diminish expression at others.

REFERENCES
Adams, M., Montague, C.T., Prins, J.B., Holder, J.C., Smith, S.A., Sanders, L., Digby, J.E., Sewter, C.P.,
Lazar, M.A., Chatterjee, V.K.K., O’Rahilly, S., 1997. Activators of peroxisome proliferator-activated
receptor have depot-specific effects on human preadipocyte differentiation. J. Clin. Invest. 100,
3149–3153.
Allen, C.E., 1976. Cellularity of adipose tissue in meat animals. Fed. Proc. 35, 2302–2307.
Allen, C.E., Beitz, D.C., Cramer, D.A., Kauffman, R.G., 1976. Biology of fat in meat animals. North
Central Regional Research Publication No. 234, Research Division, College of Agricultural and Life
Sciences, University of Wisconsin-Madison.
Andersen, M.K., Bailey, J.W., Wilken, C., Rule, D.C., 1996. Lipoprotein lipase and glycerophosphate
acyltransferase in ovine tissues are influenced by growth and energy intake regimen. J. Nutr.
Biochem. 7, 610–616.
Anderson, D.B., Kauffman, R.G., 1973. Cellular and enzymatic changes in porcine adipose tissue during
growth. J. Lipid Res. 14, 160–168.
Anderson, D.B., Kauffman, R.G., Kastenschmidt, L.L., 1972. Lipogenic enzyme activities and cellular-
ity of porcine adipose tissue from various anatomical locations. J. Lipid. Res. 13, 593–599.
Arana, A., Vernon, R.G., Eguinoa, P., Soret, B., Mendizabal, A., Purroy, A., 2002. Differentiation in vitro
of omental and subcutaneous pre-adipocytes from Spanish Lacha and Rasa Aragonesa sheep. Anim.
Sci. 74, 469–476.
Aso, H., Abe, H., Nakajima, I., Ozutsumi, K., Yamaguchi, T., Takamori, Y., Kodama, A., Hoshino, F.B.,
Takano, S., 1995. A preadipocyte clonal line from bovine intramuscular adipose tissue: nonexpression
of glut-4 protein during adipocyte differentiation. Biochem. Biophys. Res. Commun. 213, 369–375.
Bauman, D.E., 1976. Intermediary metabolism of adipose tissue. Fed. Proc. 35, 2308–2313.
Bas, P., 1992. Changes in activities of lipogenic enzymes in adipose tissue and liver of growing goats.
J. Anim. Sci. 70, 3857–3866.
Bass, J.J., Butler-Hogg, B.W., Kirton, A.H., 1990. Practical methods of controlling fatness in farm
animals. In: Wood, J.D., Fisher, A.V. (Eds.), Reducing Fat in Meat Animals. Elsevier Applied Science,
London, pp. 145–200.
Beermann, D.H., 1994. Carcass composition of animals given partitioning agents. In: Hafs, H.D.,
Zimbelman, R.G. (Eds.), Low-Fat Meats Design Strategies and Human Implications. Academic Press,
New York, pp. 203–232.
Beermann, D.H., DeVol, D.L., 1991. Effects of somatotropin, somatotropin releasing factor and somato-
statin on growth. In: Pearson, A.M., Dutson, T.R. (Eds.), Growth Regulation in Farm Animals.
Elsevier Applied Science, London, pp. 373–426.
Belfrage, P., 1985. Hormonal control of lipid degradation. In: Cryer, A., Van, R.L.R. (Eds.),
New Perspectives in Adipose Tissue: Structure, Function and Development. Butterworths, London,
pp. 121–144.
Berg, R.T., Walters, L.E., 1983. The meat animal: changes and challenges. J. Anim. Sci. 57, Suppl. 2,
133–146.
Bergen, W.G., 2001. The role of cyclic AMP elevating agents and somatotropin in pre and posttransla-
tional regulation of lipogenesis and lipolysis in Bos taurus and Sus scrofa. Recent Res. Dev. Lipids,
5, 47–59.
Bjorntorp, P., Karlsson, M., Pettersson, P., Sypniewska, G., 1980. Differentiation and function of rat
adipocyte precursor cells in primary culture. J. Lipid Res. 21, 714–723.
Bonnet, F.P., 1981. Adipose Tissue in Childhood. CRC Press, Boca Raton, FL.
Bonnet, M., Faulconnier, Y., Flechet, J., Hocquette, J.F., Leroux, C., Langin, D., Martin, P., Chilliard, Y.,
1998. Messenger RNAs encoding lipoprotein lipase, fatty acid synthase and hormone-sensitive lipase
in the adipose tissue of underfed-refed ewes and cows. Reprod. Nutr. Dev. 38, 297–307.
294 H. J. Mersmann and S. B. Smith

Bonnet, M., Leroux, C., Faulconnier, Y., Hocquette, J.F., Bocquier, F., Martin, P., Chilliard, Y., 2000.
Lipoprotein lipase activity and mRNA are up-regulated by refeeding in adipose tissue and cardiac
muscle of sheep. J. Nutr. 130, 749–756.
Boone, C., Gregoire, F., Remacle, C., 1999. Regulation of porcine adipogenesis in vitro, as compared
with other species. Domest. Anim. Endocrinol. 17, 257–267.
Boone, C., Gregoire, F., Remacle, C., 2000. Culture of porcine stromal-vascular cells in serum-free
medium: differential action of various hormonal agents on adipose conversion. J. Anim. Sci. 78,
885–895.
Broad, T.E., Ham, R.G., 1983. Growth and adipose differentiation of sheep preadipocyte fibroblasts in
serum-free medium. Eur. J. Biochem. 135, 33–39.
Brun, R.P., Kim, J.B., Hu, E., Altiok, S., Spiegelman, B.M., 1996. Adipocyte differentiation: A tran-
scriptional regulatory cascade. Curr. Opin. Cell. Biol. 8, 826–832.
Budd, T.J., Atkinson, J.L., Buttery, P.J., Salter, A.M., Wiseman, J., 1994. Effect of insulin and isopro-
terenol on lipid metabolism in porcine adipose tissue from different depots. Comp. Biochem. Physiol.
108C, 137–143.
Camara, M., Mourot, J., Fevrier, C., 1996. Influence of two dairy fats on lipid syntheis in the pig: com-
parative study of liver, muscle and the two backfat layers. Ann. Nutr. Metab. 40, 287–295.
Carey, G.B., 1995. Adenosine regulation of adipose tissue metabolism. In: Smith, S.B., Smith, D.R.
(Eds.), The Biology of Fat in Meat Animals: Current Advances. Amer. Soc. Anim. Sci., Champaign, IL,
pp. 35–52.
Castillo, G., Hauser, S., Rosenfield, J.K., Spiegelman, B.M., 1999. Role and regulation of PPARγ during
adipogenesis. J. Anim. Sci. 77, Suppl. 3, 9–15.
Chapman, M.J., Forgez, P., 1985. Lipid transport systems: some recent aspects in swine, cattle and trout
during development. Reprod. Nutr. Dev. 25, 217–226.
Chawla, A., Schwarz, E.J., Dimaculangan, D.D., Lazar, M.A., 1994. Peroxisome proliferator-activated
receptor (PPAR): adipose-predominant expression and induction early in adipocyte differentiation.
Endocrinology 135, 798–800.
Chen, X.L., Dean, R.G., Hausman, G.J., 1999. Expression of leptin MRNA and CCAAT-enhancer
binding proteins in response to insulin deprivation during preadipocyte differentiation in primary
cultures of porcine stromal-vascular cells. Domest. Anim. Endocrinol. 17, 389–401.
Chilliard, Y., 1993. Dietary fat and adipose tissue metabolism in ruminants, pigs, and rodents: a review.
J. Dairy Sci. 76, 3897–3931.
Chilliard, Y., Ferlay, A., Faulconnier, Y., Bonnet, M., Rouel, J., Bocquier, F., 2000. Adipose tissue metab-
olism and its role in adaptations to undernutrition in ruminants. Proc. Nutr. Soc. 59, 127–134.
Cinti, S., 2001. The adipose organ: morphological perspectives of adipose tissues. Proc. Nutr. Soc. 60,
319–328.
Coleman, M.E., Ekeren, P.A., Smith, S.B., 1988. Lipid synthesis and adipocyte growth in adipose tissue
from sheep chronically fed a beta-adrenergic agent. J. Anim. Sci. 66, 372–378.
Cryer, A., Gray, B.R., Woodhead, J.S., 1984. Studies on the characterization of bovine adipocyte precursor
cells and their differentiation in vitro, using an indirect-labeled-second-antibody cellular immunoassay.
J. Dev. Physiol. 6, 159–176.
Cryer, A., Williams, S.E., Cryer, J., 1992. Dietary and other factors involved in the proliferation, deter-
mination and differentiation of adipocyte precursor cells. Proc. Nutr. Soc. 51, 379–385.
de la Hoz, L., Vernon, R.G., 1996. Regulation of lipoprotein lipase activity by growth hormone in differ-
entiated adipocytes from sheep. Horm. Metab. Res. 28, 202–204.
Demaree, S.R., Gilbert, C.D., Mersmann, H.J., Smith, S.B., 2002. Conjugated linoleic acid differentially
modifies fatty acid composition in subcellular fractions of muscle and adipose tissue but not adiposity
of postweanling pigs. J. Nutr. 132, 3272–3279.
Desnoyers, F., Pascal, G., Etienne, M., Vodovar, N., 1980. Cellularity of adipose tissue in fetal pig.
J. Lipid Res. 21, 301–308.
DiGirolamo, M., Mendlinger, S., 1971. Role of fat cell size and number in enlargement of epididymal fat
pads in three species. Amer. J. Physiol. 221, 859–864.
Ding, S.T., McNeel, R.L., Mersmann, H.J., 1999. Expression of porcine adipocyte transcripts: tissue
distribution and differentiation in vitro and in vivo. Comp. Biochem. Physiol. 123B, 307–318.
Enser, M.B., Wood, J.D., Restall, D.J., MacFie, H.J.H., 1976. The cellularity of adipose tissue from pigs
of different weights. J. Agr. Sci. Camb. 86, 633–638.
Development of white adipose tissue lipid metabolism 295

Etherton, T.D., Chung, C.S., 1981. Preparation, characterization, and insulin sensitivity of isolated swine
adipocytes: comparison with adipose tissue slices. J. Lipid Res. 22, 1053–1059.
Etherton, T.D., Louveau, I., 1992. Manipulation of adiposity by somatotropin and β-adrenergic agonists:
a comparison of their mechanisms of action. Proc. Nutr. Soc. 51, 419–431.
Etherton, T.D., Smith, S.B., 1991. Somatotropin and β-adrenergic agonists: their efficacy and mecha-
nisms of action. J. Anim. Sci. 69, 2–26.
Etherton, T.D., Walton, P.E., 1986. Hormonal and metabolic regulation of lipid metabolism in domestic
livestock. J. Anim. Sci. 63, Suppl. 2, 76–88.
Etherton, T.D., Donkin, S.S., Bauman, D.E., 1995. Mechanisms by which porcine somatotropin
(pST) decreases adipose tissue growth in growing pigs. In: Smith, S.B., Smith, D.R. (Eds.). The
Biology of Fat in Meat Animals: Current Advances, Amer. Soc. Anim. Sci., Champaign, IL,
pp. 53–69.
Fajas, L., Fruchart, J.C., Auwerx, J., 1998. Transcriptional control of adipogenesis. Curr. Opin. Cell. Biol.
10, 165–173.
Farnsworth, E.R., Kramer, J.K.G., 1987. Fat metabolism in growing swine: a review. Can. J. Anim. Sci.
67, 301–318.
Feve, B., Emorine, L.J., Lasnier, F., Blin, N., Baude, B., Nahmias, C., Strosberg, A.D., Pairault, J., 1991.
Atypical β-adrenergic receptor in 3T3-F442A adipocytes. J. Biol. Chem. 266, 20329–20336.
Fruhbeck, G., Gomez-Ambrosi, J., Muruzabal, F.J., Burrell, M.A., 2001. The adipocyte: a model for inte-
gration of endocrine and metabolic signaling in energy metabolism regulation. Amer. J. Physiol.
Endocrinol. Metab. 280, E827–E847.
Gerfault, V., Louveau, I., Mourot, J., Le Dividich, J., 2000. Lipogenic enzyme activities in subcutaneous
adipose tissue and skeletal muscle from neonatal pigs consuming maternal or formula milk. Reprod.
Nutr. Dev. 40, 103–112.
Girard, J., Ferre, P., Pegorier, J.P., Duee, P.H., 1992. Adaptations of glucose and fatty acid metabolism
during perinatal period and suckling-weaning transition. Physiol. Rev. 72, 507–562.
Greathead, H.M., Dawson, J.M., Scollan, N.D., Buttery, P.J., 2001. In vivo measurement of lipogenesis
in ruminants using [1-14C]acetate. Brit. J. Nutr. 86, 37–44.
Greenwood, M.R., Hirsch, J., 1974. Postnatal development of adipocyte cellularity in the normal rat.
J. Lipid Res. 15, 474–483.
Gregoire, F.M., 2001. Adipocyte differentiation: from fibroblast to endocrine cell. Exp. Biol. Med. 226,
997–1002.
Gregoire, F.M., Smas, C.M., Sul, H.S., 1998. Understanding adipocyte differentiation. Physiol. Rev.
78, 783–809.
Gurr, M.I., Kirtland, J., 1978. Adipose tissue cellularity: a review. I. Techniques for studying cellularity.
Int. J. Obes. 2, 401–427.
Gurr, M.I., Kirtland, J., Phillip, M., Robinson, M.P., 1977. The consequences of early overnutrition for
fat cell size and number: the pig as an experimental model for human obesity. Int. J. Obes. 1, 151–170.
Hamosh, M., 1998. Neonatal lipid metabolism. In: Cowett, R.M. (Ed.), Principles of Perinatal-Neonatal
Metabolism. Springer, New York, pp. 821–846.
Hancock, D.L., Wagner, J.F., Anderson, D.B., 1991. Effects of estrogens and androgens on animal
growth. In: Pearson, A.M., Dutson, T.R. (Eds.), Growth Regulation in Farm Animals. Elsevier
Applied Science, London, pp. 255–297.
Harris, D.M., Dunshea, F.R., Bauman, D.E., Boyd, R.D., Wang, S.Y., Johnson, P.A., Clarke, S.D., 1993.
Effect of in vivo somatotropin treatment of growing pigs on adipose tissue lipogenesis. J. Anim. Sci.
71, 3293–3300.
Hauner, H., Entenmann, G., Wabitsch, M., Gaillard, D., Ailhaud, G., Negrel, R., Pfeiffer, E.F., 1989.
Promoting effect of glucocorticoids on the differentiation of human adipocyte precursor cells cultured
in a chemically defined medium. J. Clin. Invest. 84, 1663–1670.
Hauser, N., Mourot, J., De Clercq, L., Genart, C., Remacle, C., 1997. The cellularity of developing
adipose tissues in Pietrain and Meishan pigs. Reprod. Nutr. Dev. 37, 617–625.
Hausman, D.B., DiGirolamo, M., Bartness, T.J., Hausman, G.J., Martin, R.J., 2001. The biology of white
adipocyte proliferation. Obes. Rev. 2, 239–254.
Hausman, G.J., 1985. The comparative anatomy of adipose tissue. In: Cryer, A., Van, R.L.R. (Eds.),
New Perspectives in Adipose Tissue: Structure, Function and Development. Butterworths, London,
pp. 1–21.
296 H. J. Mersmann and S. B. Smith

Hausman, G.J., Kauffman, R.G., 1986a. Mitotic activity in fetal and early postnatal porcine adipose
tissue. J. Anim. Sci. 63, 659–673.
Hausman, G.J., Kauffman, R.G., 1986b. The histology of developing porcine adipose tissue. J. Anim. Sci.
63, 642–658.
Hausman, G.J., Richardson, L.R., 1982. Histochemical and ultrastructural analysis of developing
adipocytes in the fetal pig. Acta Anat. 114, 228–247.
Hausman, G.J., Jewell, D.E., Hentges, E.J., 1989. Endocrine regulation of adipogenesis. In: Champion, D.R.,
Hausman, G.J., Martin, R.J. (Eds.), Animal Growth Regulation. Plenum Press, New York,
pp. 49–68.
Hausman, G.J., Novakofski, J.E., Martin, R.J., Thomas, G.B., 1984. The development of adipocytes
in primary stromal-vascular culture of fetal pig adipose tissue. Cell Tissue Res. 236, 459–464.
Hausman, G.J., Wright, J.T., Dean, R., Richardson, R.L., 1993. Cellular and molecular aspects of the
regulation of adipogenesis. J. Anim. Sci. 71, Suppl. 2, 33–55.
Hentges, E.J., Hausman, G.J., 1989. Primary cultures of stromal-vascular cells from pig adipose tissue:
the influence of glucocorticoids and insulin as inducers of adipocyte differentiation. Domest. Anim.
Endocrinol. 6, 275–285.
Herrera, E., 2002. Implications of dietary fatty acids during pregnancy on placental, fetal and postnatal
development: a review. Placenta 23, S9–S19.
Hocquette, J.F., Graulet, B., Vermorel, M., Bauchart, D., 2001. Weaning affects lipoprotein lipase
activity and gene expression in adipose tissues and in masseter but not in other muscles of the calf.
Brit. J. Nutr. 86, 433–441.
Hood, R.L., 1982. Relationships among growth, adipose cell size, and lipid metabolism in ruminant
adipose tissue. Fed. Proc. 41, 2555–2561.
Hood, R.L., Allen, C.E., 1973a. Comparative methods for the expression of enzyme data in porcine
adipose tissue. Comp. Biochem. Physiol. 44B, 677–686.
Hood, R.L., Allen, C.E., 1973b. Cellularity of bovine adipose tissue. J. Lipid Res. 14, 605–610.
Hood, R.L., Allen, C.E., 1977. Cellularity of porcine adipose tissue: effects of growth and adiposity.
J. Lipid Res. 18, 275–284.
Hood, R.L., Thornton, R.F., 1979. The cellularity of ovine adipose tissue. Aust. J. Agric. Res. 30,
153–161.
Hwang, C.S., Loftus, T.M., Mandrup, S., Lane, M.D., 1997. Adipocyte differentiation and leptin expression.
Annu. Rev. Cell. Dev. Biol. 13, 231–259.
Jump, D.B., MacDougald, O.A., 1993. Hormonal regulation of gene expression in cultured adipocytes.
J. Anim. Sci. 71, Suppl. 2, 56–64.
Kauffman, R.G., Crenshaw, T.D., Rutledge, J.J., Hull, D.H., Grisdale, B.S., Panalba, J., 1986.
Porcine growth: postnatal development of major body components in the boar. In: University
of Wisconsin-Madison, College of Agricultural and Life Sciences Research Report R3355,
pp. 1–25.
Kim, H.S., Hausman, G.J., Hausman, D.B., Martin, R.J., Dean, R.G., 2000. The expression of peroxi-
some proliferator-activated receptor gamma in pig fetal tissue and primary stromal-vascular cultures.
Obes. Res. 8, 83–88.
Kim, J.B., Spiegelman, B.M., 1996. ADD1/SREBP1 promotes adipocyte differentiation and gene expres-
sion linked to fatty acid metabolism. Genes Dev. 10, 1096–1107.
Kim, S., Moustaid-Moussa, N., 2000. Secretory, endocrine and autocrine / paracrine function of the
adipocyte. J. Nutr. 130, 3110S–3115S.
Kimura, R.E., 1998. Lipid metabolism in the fetal-placental unit. In: Cowett, R.M. (Ed.), Principles of
Perinatal-Neonatal Metabolism. Springer, New York, pp. 389–402.
Kliewer, S.A., Sundseth, S.S., Jones, S.A., Brown, P.J., Wisely, G.B., Koble, C.S., Devchand, P., Wahli, W.,
Willson, T.M., Lenhard, J.M., Lehmann, J.M., 1997. Fatty acids and eicosanoids regulate gene
expression through direct interactions with peroxisome proliferator-activated receptors. Proc. Natl.
Acad. Sci. USA 94, 4318–4323.
Kobatake, T., Watanabe, Y., Matsuzawa, Y., Tokunaga, K., Fujioka, S., Kawamoto, T., Keno, Y., Tarui, S.,
Yoshida, H., 1991. Age-related changes in adrenergic alpha 1, alpha 2, and beta receptors of rat white
fat cell membranes: an analysis using [3H]bunazosin as a novel ligand for the alpha 1 adrenoceptor.
J. Lipid Res. 32, 191–196.
Kouba, M., Bonneau, M., Noble, J., 1999. Relative development of subcutaneous, intermuscular, and
kidney fat in growing pigs with different body compositions. J. Anim. Sci. 77, 622–629.
Development of white adipose tissue lipid metabolism 297

Lane, M.D., Tang, Q.Q., Jiang, M.S., 1999. The adipocyte differentiation program: repression/depression
of the C/EBPα gene. J. Anim. Sci. 77, Suppl. 3, 23–32.
Lanna, D.P.D., Bauman, D.E., 1999. Effect of somatotropin, insulin, and glucocorticoid on lipolysis in
chronic cultures of adipose tissue from lactating cows. J. Dairy Sci. 82, 60–68.
Lazar, M.A., 1999. PPARγ in adipocyte differentiation. J. Anim. Sci. 77, Suppl. 3, 16–22.
Leat, W.M.F., Harrison, F.A., 1980. Transfer of long-chain fatty acids to the fetal and neonatal lamb.
J. Dev. Physiol. 2, 257–274.
Le Dividich, J., Herpin, P., Mourot, J., Colin, A.-P., 1994. Effect of low-fat colostrums on fat accretion and
lipogenic enzyme activities in adipose tissue in the 1-day-old pig. Comp. Biochem. Physiol. 108A,
663–671.
Lee, K., Hausman, G.J., Dean, R.G., 1998. Expression of C/EBP alpha, beta and delta in fetal and post-
natal subcutaneous adipose tissue. Mol. Cell. Biochem. 178, 269–274.
Lee, S.R., Tume, R.K., Cryer, J., Cryer, A., 1986. Studies on the expression of adipocyte-specific cell
surface antigens during the differentiation of adipocyte precursor cells in vitro. J. Dev. Physiol.
8, 207–226.
Liang, W., Mills, S.E., 2002. Quantitative analysis of beta-adrenergic receptor subtypes in pig tissues.
J. Anim. Sci. 80, 963–970.
Lin, K.C., Cross, H.R., Smith, S.B., 1992. Esterification of fatty acids by bovine intramuscular and
subcutaneous adipose tissues. Lipids 27, 111–116.
Loftus, T.M., Lane, M.D., 1997. Modulating the transcriptional control of adipogenesis. Curr. Opin.
Genet. Dev. 7, 603–608.
Marcus, C., Karpe, B., Bolme, P., Sonnenfeld, T., Arner, P., 1987. Changes in catecholamine induced
lipolysis in isolated human fat cells during the first year of life. J. Clin. Invest. 79, 1812–1818.
Martin, G.S., Lunt, D.K., Britain, K.G., Smith, S.B., 1999. Postnatal development of stearoyl coenzyme
A desaturase gene expression and adiposity in bovine subcutaneous adipose tissue. J. Anim. Sci. 77,
630–636.
Martin, R.J., Kasser, T.R., Ramsay, T.G., Hausman, G.J., 1985. Regulation of adipose tissue development
in utero. In: Cryer, A., Van, R.L.R. (Eds.), New Perspectives in Adipose Tissue: Structure, Function
and Development. Butterworths, London, pp. 303–317.
May, S.G., Savell, J.W., Lunt, D.K., Wilson, J.J., Laurenz, J.C., Smith, S.B., 1994. Evidence for
preadipocyte proliferation during culture of subcutaneous and intramuscular adipose tissues from
Angus and Wagyu crossbred steers. J. Anim. Sci. 72, 3110–3117.
McNeel, R.L., Mersmann, H.J., 1999. Distribution and quantification of beta1-, beta2-, and beta3-
adrenergic receptor subtype transcripts in porcine tissues. J. Anim. Sci. 77, 611–621.
McNeel, R.L., Mersmann, H.J., 2000. Nutritional deprivation reduces the transcripts for transcription
factors and adipocyte-characteristic proteins in porcine adipocytes. J. Nutr. Biochem. 11,
139–146.
McNeel, R.L., Ding, S., O’Brian, Smith, E., Mersmann, H.J., 2000. Expression of porcine adipocyte
transcripts during differentiation in vitro and in vivo. Comp. Biochem. Physiol. 126B, 291–302.
Mendizabal, J.A., Alberti, P., Eguinoa, P., Arana, A., Soret, B., Purroy, A., 1999. Adipocyte size and
enzyme lipogenic activities in different adipose tissue depots in steers of local Spanish breeds. Anim.
Sci. 69, 115–121.
Mendizabal, J.A., Soret, B., Purroy, A., Arana, A., Horcada, A., 1997. Infuence of sex on cellularity and
on lipogenic enzymes of Spanish lamb breeds (Lacha and Rasa Aragonesa). Anim. Sci. 64, 283–289.
Mersmann, H.J., 1974. Metabolic patterns in the neonatal swine. J. Anim. Sci. 38, 1022–1030.
Mersmann, H.J., 1986. Lipid metabolism in swine. In: Stanton, H.C., Mersmann, H.J. (Eds.), Swine in
Cardiovascular Research. CRC Press, Boca Raton, FL, Vol. 1, pp. 75–103.
Mersmann, H.J., 1987. Acute metabolic effects of adrenergic agents in swine. Amer. J. Physiol. 252,
E85–E95.
Mersmann, H.J., 1989a. The effect of insulin on porcine adipose tissue lipogenesis. Comp. Biochem.
Physiol. 94B, 709–713.
Mersmann, H.J., 1989b. Inhibition of porcine adipose tissue lipogenesis by beta-adrenergic agonists.
Comp. Biochem. Physiol. 94C, 619–623.
Mersmann, H.J., 1989c. Influence of infused β-adrenergic agonists on porcine blood metabolites and
catecholamines. J. Anim. Sci. 67, 2633–2645.
Mersmann, H.J., 1990. Metabolic and endocrine control of adipose tissue accretion. In: Wood, J.D.,
Fisher, A.V. (Eds.), Reducing Fat in Meat Animals. Elsevier Applied Science, London, pp. 101–144.
298 H. J. Mersmann and S. B. Smith

Mersmann, H.J., 1991. Regulation of adipose tissue metabolism and accretion in mammals raised for
meat production. In: Pearson, A.M., Dutson, T.R. (Eds.), Growth Regulation in Farm Animals.
Advances in Meat Research, Vol. 7. Elsevier Applied Science, London. pp. 135–168.
Mersmann, H.J., 1995. Species variation in mechanisms for modulation of growth by beta-adrenergic
receptors. J. Nutr. 125, 1777S–1782S.
Mersmann, H.J., 1998. Overview of the effects of beta-adrenergic receptor agonists on animal growth
including mechanisms of action. J. Anim. Sci. 76, 160–172.
Mersmann, H.J., 2002a. Beta-adrenergic receptor modulation of adipocyte metabolism and growth.
J. Anim. Sci. 80, E24–E29.
Mersmann, H.J., 2002b. Mechanisms for conjugated linoleic acid-mediated reduction in fat deposition.
J. Anim. Sci. 80, E126–E134.
Mersmann, H.J., Brown, L.J., 1973. Adaptation of swine (Sus domesticus) adipose tissue to increased
lipogenesis: expression of data. Int. J. Biochem. 4, 503–510.
Mersmann, H.J., Hu, C.Y., 1987. Factors affecting measurements of glucose metabolism and lipolytic
rates in porcine adipose tissue slices in vitro. J. Anim. Sci. 64, 148–164.
Mersmann, H.J., Allen, C.D., Chai, E.Y., Brown, L.J., Fogg, T.J., 1981. Factors influencing the lipogenic
rate in swine adipose tissue. J. Anim. Sci. 52, 1298–1305.
Mersmann, H.J., Brown, L.J., Beuving, R.M., Arakelian, M.C., 1976. Lipolytic activity of swine
adipocytes. Amer. J. Physiol. 230, 1439–1443.
Mersmann, H.J., Goodman, J.R., Brown, L.J., 1975. Development of swine adipose tissue: morphology
and chemical composition. J. Lipid. Res. 16, 269–279.
Mersmann, H.J., Houk, J.M., Phinney, G., Underwood, M.C., 1973a. Effect of diet and weaning age on
in vitro lipogenesis in young swine. J. Nutr. 103, 821–828.
Mersmann, H.J., Houk, J.M., Phinney, G., Underwood, M.C., Brown, L.J., 1973b. Lipogenesis by in vitro
liver and adipose tissue preparations from neonatal swine. Amer. J. Physiol. 224, 1123–1129.
Mersmann, H.J., Underwood, M.C., Brown, L.J., Houk, J.M., 1973c. Adipose tissue composition and
lipogenic capacity in developing swine. Amer. J. Physiol. 224, 1130–1135.
Miller, M.F., Cross, H.R., Wilson, J.J., Smith, S.B., 1989. Acute and long-term lipogenic response to
insulin and clenbuterol in bovine intramuscular and subcutaneous adipose tissues. J. Anim. Sci. 67,
928–933.
Miller, M.F., Garcia, D.K., Coleman, M.E., Ekeren, P.A., Lunt, D.K., Wagner, K.A., Procknor, M.,
Welsh, T.H. Jr., Smith, S.B., 1988. Adipose tissue, longissimus muscle and anterior pituitary growth
and function in clenbuterol-fed heifers. J. Anim. Sci. 66, 12–20.
Miller, W.H. Jr., Faust, I.M., Hirsch, J., 1984. Demonstration of de novo production of adipocytes in adult
rats by biochemical and radioautographic techniques. J. Lipid. Res. 25, 336–347.
Mills, S.E., 1999. Regulation of porcine adipocyte metabolism by insulin and adenosine. J. Anim. Sci.
77, 3201–3207.
Mills, S., Mersmann, H.J., 1995. Beta-adrenergic agonists, their receptors, and growth: special reference
to the peculiarities in pigs. In: Smith, S.B., Smith, D.R. (Eds.), The Biology of Fat in Meat Animals:
Current Advances. Amer. Soc. Anim. Sci., Champaign, IL. pp. 1–34.
Mills, S.E., Liu, C.Y., Gu, Y., Schinckel, A.P., 1990. Effects of ractopamine on adipose tissue metabolism
and insulin binding in finishing hogs: interaction with genotype and slaughter weight. Domest. Anim.
Endocrinol. 7, 251–263.
Mitchell, A.D., Scholz, A.M., Mersmann, H.J., 2001. Growth and body composition. In: Pond, W.G.,
Mersmann, H.J. (Eds.), Biology of the Domestic Pig. Cornell University Press, Ithaca, NY,
pp. 225–308.
Moloney, A.P., Allen, P., Enright, W.J., 2002. Body composition and adipose tissue accretion in lambs
passively immunised against adipose tissue. Livest. Prod. Sci, 74, 165–174.
Moloney, A., Allen, P., Joseph, R., Tarrant, V., 1991. Influence of beta-adrenergic agonists and similar
compounds in growth. In: Pearson, A.M., Dutson, T.R. (Eds.), Growth Regulation in Farm Animals.
Elsevier Applied Science, London, pp. 455–513.
Moody, D.E., Hancock, D.L., Anderson, D.B., 2000. Phenethanolamine repartitioning agents. In:
D’Mello, J.P.F. (Ed.), Farm Animal Metabolism and Nutrition. CABI Publishing, New York
pp. 65–96.
Morrison, R.F., Farmer, S.R., 2000. Hormonal signaling and transcriptional control of adipocyte differ-
entiation. J. Nutr. 130, 3116S–3121S.
Development of white adipose tissue lipid metabolism 299

Mourot, J., Kouba, M., Bonneau, M., 1996. Comparative study of in vitro lipogenesis in various adipose
tissues in the growing Meishan pig: comparison with the large white pig (Sus domesticus). Comp.
Biochem. Physiol. 115B, 383–388.
Mourot, J., Kouba, M., Peiniau, P., 1995. Comparative study of in vitro lipogenesis in various
adipose tissues in the growing domestic pig (Sus domesticus). Comp. Biochem. Physiol. 111B,
379–384.
Napolitano, L.M., 1963. The differentiation of white adipose cells: an electron microscope study. J. Cell
Biol. 18, 663–679.
Newby, D., Gertler, A., Vernon, R.G., 2001. Effects of recombinant ovine leptin on in vitro lipolysis and
lipogenesis in subcutaneous adipose tissue from lactating and nonlactating sheep. J. Anim. Sci. 79,
445–452.
Noble, R.C., 1981. Lipid metabolism in the neonatal ruminant. In: Christie, W.W. (Ed.), Lipid
Metabolism in Ruminant Animals. Pergamon Press, Oxford, pp. 411–448.
Novakofski, J., Hu, C.Y., 1987. Culture of isolated adipose tissue cells. J. Anim. Sci. 65, Suppl. 2, 12–24.
NRC, 1994. Metabolic Modifiers: Effects on the Nutrient Requirements of Food-Producing Animals.
National Research Council, National Academy Press, Washington, DC.
Ntambi, J.M., Young-Cheul, K., 2000. Adipocyte differentiation and gene expression. J. Nutr. 130,
3122S–3126S.
Ohyama, M., Matsuda, K., Torii, S., Matsui, T., Yano, H., Kawada, T., Ishihara, T., 1998. The interaction
between vitamin A and thiazolidinedione on bovine adipocyte differentiation in primary culture.
J. Anim. Sci. 76, 61–65.
Payne, E., 1999. Fatty acid synthesis in lambs in vivo as affected by nutrition and hormone administra-
tion. Aust. J. Agric. Res., 50, 163–174.
Peixing, W.U., Sato, K., Suzuta, F., Hikasa, Y., Kagota, K., 2000. Effects of lipid-related factors on
adipocyte differentiation of bovine stromal-vascular cells in primary culture. J. Vet. Med. Sci. 62,
933–939.
Piétri-Rouxel, F., Lenzen, G., Kapoor, A., Drumare, M.-F., Archimbault, P., Strosberg, A.D.,
Manning, B.S.J., 1995. Molecular cloning and pharmacological characterization of the bovine
β3-adrenergic receptor. Eur. J. Biochem. 230, 350–358.
Plaas, H.A., Cryer, A., 1980. The isolation and characterization of a proposed adipocyte precursor cell
type from bovine subcutaneous white adipose tissue. J. Dev. Physiol. 2, 275–289.
Purroy, A., Mendizabal, J.A., Soret, B., Arana, A., Mendizabal, F.J., 1997. Changes in cell number and
size and in lipogenic enzyme activity in adipose tissues during growth and fattening of Lacha
(Manech) lambs. Ann. Zootech. 46, 309–319.
Ramsay, T.G., 2001. Porcine leptin alters insulin inhibition of lipolysis in porcine adipocytes in vitro.
J. Anim. Sci. 79, 653–657.
Ramsay, T.G., White, M.E., Wolverton, C.K., 1989. Glucocorticoids and the differentiation of porcine
preadipocytes. J. Anim. Sci. 67, 2222–2229.
Rangwala, S.M., Lazar, M.A., 2000. Transcriptional control of adipogenesis. Annu. Rev. Nutr. 20,
535–559.
Robelin, J., 1981. Cellularity of bovine adipose tissues: developmental changes from 15 to 65 percent
mature weight. J. Lipid Res. 22, 452–457.
Robelin, J., 1986. Growth of adipose tissues in cattle: partitioning between depots, chemical composition
and cellularity. A review. Livest. Prod. Sci, 14, 349–364.
Robert, C., Palin, M.-F., Silversides, F.G., McKay, R.M., Pelletier, G., 2000. Messenger RNA levels of
growth factors, ligands, receptors, and proteins affecting lipid metabolism in pigs. Can. J. Anim. Sci.
80, 559–567.
Romsos, D.R., Leveille, G.A., Allee, G.L., 1971. Alloxan diabetes in the pig (Sus domesticus): respone
to glucose, tolbutamide and insulin administration. Comp. Biochem. Physiol. 40A, 557–568.
Rule, D.C., 1995. Adipose tissue glycerolipid biosynthesis. In: Smith, S.B., Smith, D.R. (Eds.),
The Biology of Fat in Meat Animals: Current Advances. Amer. Soc. Anim. Sci., Champaign, IL,
pp. 129–143.
Rule, D.C., Smith, S.B., Mersmann, H.J., 1987. Effects of adrenergic agonists and insulin on porcine
adipose tissue lipid metabolism in vitro. J. Anim. Sci. 65, 136–149.
Rule, D.C., Smith, S.B., Mersmann, H.J., 1988a. Glycerolipid biosynthesis in porcine adipose tissue
in vitro. II. Synthesis by various types of cellular preparations. J. Anim. Sci. 66, 1665–1675.
300 H. J. Mersmann and S. B. Smith

Rule, D.C., Smith, S.B., Mersmann, H.J., 1988b. Glycerolipid biosynthesis in porcine adipose tissue
in vitro. I. Assay conditions for homogenates. J. Anim. Sci. 66, 1656–1664.
Rule, D.C., Smith, S.B., Mersmann, H.J., 1989. Glycerolipid biosynthesis in porcine adipose tissue
in vitro: effect of adiposity and depot site. J. Anim. Sci. 67, 364–373.
Rule, D.C., Smith, S.B., Romans, J.R., 1995. Fatty acid composition of muscle and adipose tissue of meat
animals. In: Smith, S.B., Smith, D.R. (Eds.), The Biology of Fat in Meat Animals: Current Advances.
Amer. Soc. Anim. Sci., Champaign, IL, pp. 144–165.
Saggerson, E.D., 1985. Hormonal regulation of biosynthetic activities in white adipose tissue. In: Cryer, A.,
Van, R.L.R. (Eds.), New Perspectives in Adipose Tissue: Structure, Function and Development.
Butterworths, London, pp. 87–120.
Schiavetta, A.M., Miller, M.F., Lunt, D.K., Davis, S.K., Smith, S.B., 1990. Adipose tissue cellularity and
muscle growth in young steers fed the beta-adrenergic agonist clenbuterol for 50 days and after
78 days of withdrawal. J. Anim. Sci. 68, 3614–3623.
Schoonjans, K., Staels, B., Auwerx, J., 1996. Role of the peroxisome proliferator-activated receptor (PPAR)
in mediating the effect of fibrates and fatty acids on gene expression. J. Lipid Res. 37, 907–925.
Sejrsen, K., Vestergaard, M., Neimann-Sorensen, A. (Eds.), 1989. Use of Somatotropin in Livestock
Production. Elsevier Applied Science, London.
Shillabeer, G., Lau, D.C., 1994. Regulation of new fat cell formation in rats: the role of dietary fats.
J. Lipid Res. 35, 592–600.
Slavin, B.G., 1985. The morphology of adipose tissue. In: Cryer, A., Van, R.L.R. (Eds.), New
Perspectives in Adipose Tissue: Structure, Function and Development. Butterworths, London,
pp. 23–43.
Smas, C.M., Sul, H.S., 1995. Control of adipocyte differentiation. Biochem. J. 309, 697–710.
Smith, S.B., 1984. Evidence that phosphofructokinase limits glucose utilization in bovine adipose tissue.
J. Anim. Sci., 58, 1198–1204.
Smith, S.B., 1995. Substrate utilization in ruminant adipose tissues. In: Smith, S.B., Smith, D.R. (Eds.),
The Biology of Fat in Meat Animals: Current Advances. Amer. Soc. Anim. Sci., Champaign, IL,
pp. 166–188.
Smith, S.B., Smith, D.R. (Eds.), 1995. The Biology of Fat in Meat Animals: Current Advances. Amer.
Soc. Anim. Sci., Champaign, IL.
Smith, S.B., Lin, K.C., Wilson, J.J., Lunt, D.K., Cross, H.R., 1998. Starvation depresses acylglycerol
biosynthesis in bovine subcutaneous but not intramuscular adipose tissue homogenates. Comp.
Biochem. Physiol. 120B, 165–174.
Smith, S.B., Prior, R.L., Ferrell, C.L., Mersmann, H.J., 1984. Interrelationships among diet, age, fat
deposition and lipid metabolism in growing steers. J. Nutr. 114, 153–162.
Smith, S.B., Prior, R.L., Mersmann, H.J., 1983. Interrelationships between insulin and lipid metabolism
in normal and alloxan-diabetic cattle. J. Nutr. 113, 1002–1015.
Soret, B., Lee, H.J., Finley, E., Lee, S.C., Vernon, R.G., 1999. Regulation of differentiation of sheep
subcutaneous and abdominal preadipocytes in culture. J. Endocrinol. 161, 517–524.
Soret, B., Mendizabal, J.A., Arana, A., Purroy, A., Eguinoa, P., 1998. Breed effects on cellularity and
lipogenic enzymes in growing Spanish lambs. Small Ruminant Res. 29, 103–112.
Steele, N.C. (Ed.), 1991. Emerging technologies to alter fat:lean ratio in animal products. J. Anim. Sci.
69, Suppl. 2, pp. 1–124.
Steele, N.C., Boyd, R.D., Campbell, R.G., 1994. Growth, metabolic modifiers, and nutrient considera-
tions. In: Hafs, H.D., Zimbelman, R.G. (Eds.), Low-Fat Meats Design Strategies and Human
Implications. Academic Press, New York, pp. 167–189.
Steffen, D.G., Brown, L.J., Mersmann, H.J., 1978. Ontogenic development of swine (Sus domesticus)
adipose tissue lipases. Comp. Biochem. Physiol. 59, 195–198.
Steffen, D.G., Phinney, G., Brown, L.J., Mersmann, H.J., 1979. Ontogeny of glycerolipid biosynthetic
enzymes in swine liver and adipose tissue. J. Lipid Res. 20, 246–253.
St. John, L.C., Rule, D.C., Knabe, D.A., Mersmann, H.J., Smith, S.B., 1987. Fatty acid-binding protein
activity in tissues from pigs fed diets containing 0 and 20% high oleate oil. J. Nutr. 117, 2021–2026.
Suryawan, A., Hu, C.Y., 1993. Effect of serum on differentiation of porcine adipose stromal-vascular
cells in primary culture. Comp. Biochem. Physiol. 105A, 485–492.
Suryawan, A., Hu, C.Y., 1995. The primary cell culture system for preadipocytes. In: Smith, S.B.,
Smith, D.R. (Eds.), The Biology of Fat in Meat Animals: Current Advances. Amer. Soc. Anim. Sci.,
Champaign, IL, pp. 78–92.
Development of white adipose tissue lipid metabolism 301

Tavangar, K., Murata, Y., Patel, S., Kalinyak, J.E., Pedersen, M.E., Goers, J.F., Hoffman, A.R.,
Kraemer, F.B., 1992. Developmental regulation of lipoprotein lipase in rats. Amer. J. Physiol. 262,
E330–E337.
Thulin, A.J., Allee, G.L., Harmon, D.L., Davis, D.L., 1989. Utero-placental transfer of octanoic, palmitic
and linoleic acids during late gestation in gilts. J. Anim. Sci. 67, 738–745.
Tontonoz, P., Hu, E., Graves, R.A., Budavari, A.I., Spiegelman, B.M., 1994. mPPAR gamma 2: tissue-
specific regulator of an adipocyte enhancer. Genes Dev. 8, 1224–1234.
Torii, S.-I., Kawada, T., Matsuda, K., Matsui, T., Ishihara, T., Yano, H., 1998. Thiazolidinedione induces
the adipose differentiation of fibroblast-like cells resident within bovine skeletal muscle. Cell Biol.
Int. 22, 421–427.
Trayhurn, P., Beattie, J.H., 2001. Physiological role of adipose tissue: white adipose tissue as an
endocrine and secretory organ. Proc. Nutr. Soc. 60, 329–339.
Trenkle, A., Marple, D.N., 1983. Growth and development of meat animals. J. Anim. Sci. 57, Suppl. 2,
273–283.
Truscott, T.G., Wood, J.D., Macfie, H.J.H., 1983. Fat deposition in Hereford and Friesian steers. J. Agr.
Sci. Camb. 100, 257–270.
Vasilatos, R., Etherton, T.D., Wangsness, P.J., 1983. Preparation of isolated bovine adipocytes: validation
of use for studies characterizing insulin sensitivity and binding. Endocrinology 112, 1667–1673.
Vernon, R.G., 1980. Lipid metabolism in the adipose tissue of ruminant animals. Prog. Lipid Res. 19,
23–106.
Vernon, R.G., 1981. Lipid metabolism in the adipose tissue of ruminant animals. In: Christie, W.W. (Ed.),
Lipid Metabolism in Ruminant Animals. Pergamon Press, Oxford, pp. 279–362.
Vernon, R.G., 1982. Development in vitro of fatty acid synthesis in perirenal adipose tissue from
new-born lambs. Biol. Neonate 42, 275–278.
Vernon, R.G., 1986. The growth and metabolism of adipocytes. In: Buttery, P.J., Lindsay, D.B.,
Haynes, N.B. (Eds.), Control and Manipulation of Animal Growth. Butterworths, London, pp. 67–84.
Vernon, R.G., 1991. Depot specific endocrine control of fatty acid synthesis in adipose tissues of foetal
lambs. Domest. Anim. Endocrinol. 8, 161–164.
Vernon, R.G., 1992. Control of lipogenesis and lipolysis. In: Boorman, K.N., Buttery, P.J., Lindsay, D.B.
(Eds.), The Control of Fat and Lean Deposition. Butterworth-Heinemann, Oxford, pp. 59–81.
Vernon, R.G., Clegg, R.A., 1985. The metabolism of white adipose tissue in vivo and in vitro. In: Cryer, A.,
Van, R.L.R. (Eds.), New Perspectives in Adipose Tissue: Structure, Function and Development.
Buttersworths. London, pp. 65–86.
Vernon, R.G., Taylor, E., 1986. Acetyl-CoA carboxylase of sheep adipose tissue: problems of the assay
and adaptation during fetal development. J. Anim. Sci. 63, 1119–1125.
Vernon, R.G., Barber, M.C., Finley, E., 1991. Modulation of the activity of acetyl-CoA carboxylase and
other lipogenic enzymes by growth hormone, insulin and dexamethasone in sheep adipose tissue and
relationship to adaptations to lactation. Biochem. J. 274, 543–548.
Vierck, J.L., McNamara, J.P., Dodson, M.V., 1996. Proliferation and differentiation of progeny of ovine
unilocular fat cells (adipofibroblasts). In Vitro Cell Dev. Biol. Anim. 32, 564–572.
Wajchenberg, B.L., 2000. Subcutaneous and visceral adipose tissue: their relation to the metabolic
syndrome. Endocr. Rev. 21, 697–738.
Walstra, P., 1980. Growth and Carcass Composition from Birth to Maturity in Relation to Feeding
Level and Sex in Dutch Landrace Pigs. Mededelingen Landbouwhogeschool, Wageningen, The
Netherlands.
Walton, P.E., Etherton, T.D., 1986. Stimulation of lipogenesis by insulin in swine adipose tissue: antag-
onism by porcine growth hormone. J. Anim. Sci. 62, 1584–1595.
Walton, P.E., Etherton, T.D., 1987. The culture of adipose tissue explants in serum-free medium. J. Anim.
Sci. 65, Suppl. 2, 25–30.
Widdowson, E.M., Lister, D., 1991. Nutritional control on growth. In: Pearson, A.M., Dutson, T.R. (Eds.),
Growth Regulation in Farm Animals. Elsevier Applied Science, London, pp. 67–101.
Wilson, J.J., Young, C.R., Smith, S.B., 1992. Triacylglycerol biosynthesis in bovine liver and subcuta-
neous adipose tissue. Comp. Biochem. Physiol. 103B, 511–516.
Wood, J.D., 1990. Consequences for meat quality of reducing carcass fatness. In: Wood, J.D., Fisher, A.V.
(Eds.), Reducing Fat in Meat Animals. Elsevier Applied Science, London, pp. 344–397.
Wright, J.T., Hausman, G.J., 1990. Adipose tissue development in the fetal pig examined using mono-
clonal antibodies. J. Anim. Sci. 68, 1170–1175.
302 H. J. Mersmann and S. B. Smith

Wu, P., Sato, K., Suzuta, F., Hikasa, Y., Kagota, K., 2000. Effects of lipid-related factors on adipocyte
differentiation of bovine stromal-vascular cells in primary culture. J. Vet. Med. Sci. 62, 933–939.
Yang, A., Larsen, T.W., Smith, S.B., Tume, R.K., 1999. Delta 9 desaturase activity in bovine subcutaneous
adipose tissue of different fatty acid composition. Lipids 34, 971–978.
Yu, Z.K., Hausman, G.J., 1998. Expression of CCAAT/enhancer binding proteins during porcine
preadipocyte differentiation. Exp. Cell. Res. 245, 343–349.
Yu, Z.K., Wright, J.T., Hausman, G.J., 1997. Preadipocyte recruitment in stromal vascular cultures after
depletion of committed preadipocytes by immunocytotoxicity. Obes. Res. 5, 9–15.
12 Ontogeny and metabolism of brown
adipose tissue in livestock species

S. B. Smith and G. E. Carstens

Department of Animal Science, Texas A & M University, College Station,


TX 77483-2971, USA

We have reported several aspects of brown adipose tissue (BAT) development and response
to environmental stimuli. The metabolism of BAT resembles that from mature ruminant white
adipose tissue, in that acetate is the primary precursor for lipogenesis. There is a precipitous
decline in rates of de novo lipogenesis during the last trimester, indicating that the contribu-
tion of fatty acid biosynthesis to lipid filling is more important earlier in fetal development.
The mixture of brown and white adipocytes in subcutaneous adipose tissues from newborn
calves suggests that brown adipocytes may involute into white adipocytes in this species.
However, there is growing evidence in rodent species that brown and white adipocytes
differentiate and develop independently. Clearly, brown adipose tissue from sheep and cattle
rapidly dedifferentiates and/or is lost via apoptosis early postnatally. This phenomenon is
especially rapid in warm ambient temperatures, but may be delayed by feeding dams diets
high in polyunsaturated fatty acids. In Wagyu × Angus crossbred calves, brown adipose tissue
function is remarkably refractory to profound reductions in dietary protein intake in dams,
indicating the high priority animals place in ensuring adequate thermogenic capacity in their
newborn.

1. INTRODUCTION
Adverse climatic conditions during the early postnatal period can disrupt thermal balance in
newborn lambs and calves leading to hypothermia and/or death. Numerous reports have
demonstrated that calf mortality increases during inclement weather. In an epidemiological
study involving more than 87,000 Bos taurus calves, Azzam et al. (1993) found that calf mor-
tality increased progressively as ambient temperature decreased or as precipitation amount on
the day of birth increased. Moreover, the stress of maintaining homeothermy during severe
cold exposure for extended periods may interact with other etiological factors associated with
neonatal calf mortality and morbidity through depletion of energy reserves, induction of
physical weakness, and/or delay in absorption of immunoglobulins.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
303 © 2005 Elsevier Limited. All rights reserved.
304 S. B. Smith and G. E. Carstens

A major thermoregulatory mechanism for survival of neonatal ruminants during cold stress
is heat production by brown adipose tissue (BAT). Maintenance of homeothermy during the
early postnatal period necessitates an acute and sustained thermogenic response by the new-
born calf. Maximal thermogenic response to cold (i.e. summit metabolism), which includes
both shivering and nonshivering thermogenesis, is 4- to 5-fold higher than thermoneutral
metabolism in neonatal lambs and 3- to 4-fold higher in neonatal calves (Stott and Slee, 1985;
Okamoto et al., 1986; Robinson and Young, 1988a,b).
Approximately half of the cold-induced summit metabolism in newborn lambs is derived
from nonshivering thermogenesis (Stott and Slee, 1985). Therefore, newborn lambs and
calves must possess highly active BAT during the early postnatal period when the demand for
thermogenesis is greatest. Most of the adipose tissue in the newborn ruminant species is BAT,
although small amounts of white adipose tissue (WAT) are present (Alexander et al., 1975;
Alexander, 1978; Martin et al., 1997, 1999). Alexander et al. (1975) estimated that the quantity
of BAT in newborn calves was ~1.5–2.0% of body weight.
This review will focus primarily on BAT in lambs and calves. Little mention will be made
of piglets because distinctive brown adipocytes do not appear to be present in porcine adipose
tissue.

2. MORPHOLOGY OF BROWN ADIPOSE TISSUE IN NEWBORN


LIVESTOCK SPECIES
Brown adipocytes from newborn calves do not display the typical multilocular feature that is
characteristic of brown adipose tissue in other species. Instead, bovine brown adipocytes con-
tain a large central lipid vacuole with few peripheral lipid inclusions. This is consistent with
Alexander et al. (1975), who examined perirenal brown adipocytes in newborn calves at lower
magnification and described adipocytes as dominated by a large lipid vacuole with smaller
lipid inclusions in the marginal cytoplasm of some of the cells. Napolitano (1963) stated that
the major criterion used to characterize brown adipocytes morphologically should be the
appearance and differentiation of mitochondria rather than the occurrence of multilocular
lipid droplets.
Subcutaneous adipose tissue overlying the sternum from Wagyu × Angus crossbred calves
contains unilocular adipocytes with few cytoplasmic inclusions (Martin et al., 1997). Unlike
perirenal adipocytes from these same calves, only a small number of mitochondria with
poorly developed cristae were present. Thus, sternum adipose tissue represents a WAT depot
in Wagyu × Angus calves. These results are similar to those of Alexander et al. (1975) for
newborn calves, although the earlier work described adipocytes containing a few small lipid
droplets in addition to the large central vacuole.

3. SPECIFIC GENE EXPRESSION IN BROWN ADIPOSE TISSUE


3.1. Uncoupling protein-1

Brown adipose tissue’s thermogenic capacity is attributed to uncoupling protein-1 (UCP1)


located in the inner mitochondrial membrane. The principle function of UCP1 is to dissipate
the proton gradient created by mitochondrial respiration, which uncouples mitochondrial res-
piration from synthesis of ATP and allows energy to be dissipated as heat. The concentration
of UCP1 is a key biochemical marker of the thermogenic capacity of BAT (Himms-Hagen,
1986). Brander et al. (1993) demonstrated that UCP1 mRNA was present in intrascapular
Ontogeny and metabolism of brown adipose tissue 305

BAT, but was not present in rat omental, liver, kidney, skeletal muscle, or heart. Casteilla
et al. (1989) examined perirenal, pericardial, peritoneal, and subcutaneous adipose tissue
depots in newborn calves, and found UCP1 mRNA in all tissues except subcutaneous adipose
tissue. UCP1 mRNA was highest in perirenal adipose tissue, followed by pericardial tissue
(~50% of perirenal) and peritoneal tissue (~15% of perirenal). More recently, Landis et al.
(2002) reported the presence of UCP1 mRNA in s.c. adipose tissue of Brahman and Angus
fetal calves (last trimester), indicating that s.c. adipose tissue contained a population of brown
adipocytes during late fetal development.

3.2. β-Adrenergic receptors

Expression of the β3-adrenergic receptor (β3-AR) gene is especially high in bovine perirenal
BAT during fetal and early postnatal development (Casteilla et al., 1989, 1994). β3-Adrenergic
receptors have been documented in porcine (white) adipocytes (Mersmann, 1996), indicating that
this receptor type is not unique to BAT. However, β3-AR gene expression may be prerequisite
to BAT differentiation in fetal calves (Casteilla et al., 1994).

3.3. Brown unknown gene

Moulin et al. (2001a,b) reported the existence of a gene that is expressed early in the differ-
entiation of stromal–vascular cells from rat brown adipose tissue. The gene, which they
termed BUG (brown unknown gene), is expressed at high levels in rat interscapular adipose
tissue, cardiac muscle, brain, and kidney, but at low levels in white inguinal adipose tissue,
muscle, liver, and spleen. It also is expressed at higher levels in interscapular adipose tissue
from obese rats, which exhibits depressed UCP1 gene expression (Moulin et al., 2001a). The
authors concluded that BUG gene expression must be depressed to obtain high rates of UCP1
gene expression. Their results also indicate that stromal–vascular preadipocytes from inter-
scapular (brown) adipocytes represent a cell line that is distinct from preadipocytes from
inguinal (white) adipose tissue. To our knowledge, the expression of BUG in BAT from ruminant
species has not been described.

4. ONTOGENIC DEVELOPMENT OF BROWN ADIPOSE TISSUE


Detectable quantities of perirenal BAT in lambs first appear at 70 days of gestation
(Alexander, 1978). Alexander (1978) also reported that the mass of perirenal adipose tissue
in fetal Merino sheep increased by approximately 34% over the last 3 weeks of gestation.
Vernon et al. (1981) documented a 40% increase in adipocyte volume of fetal lambs in the
last 4 weeks of pregnancy, indicating that most, if not all, of the increase in perirenal BAT
mass in fetal lambs was due to adipocyte hypertrophy. In fetal calves, some portion of BAT
hypertrophy is due to increased adipocyte volume (Landis et al., 2002), at least during the
period between 96 and 48 days before parturition (fig. 1). During this period there is a large
numerical increase in brown adipocyte volume. However, subsequent to 48 days before birth,
total apparent perirenal brown adipocyte number more than doubles (Landis et al., 2002).
Thus, hyperplastic growth of adipocytes contributes substantially to the fetal growth of
bovine perirenal BAT.
Growth of BAT in lambs, relative to fetal weight, is allometric (100 mg/kg of fetal weight) from
70 to 120 days of gestation, and isometric (6 mg/kg) thereafter until term (ovine gestation
306 S. B. Smith and G. E. Carstens

Fig. 1. Adipocyte density and mean volume of perirenal brown adipose tissue. Each data point represents
the mean for 3 fetal calves per breed type. The overall SEM for each breed type is affixed to the symbols. The
SEM are not large enough to be visible for adipocytes/g. Adapted from Figure 4 of Landis et al. (2002).

length is 150 days) (Alexander, 1978). Lipid locules were first identified in perirenal adipocytes
on day 70 of gestation, and by day 80 to 90, mitochondria began to proliferate and take on the
morphological features of brown adipocytes. Rapid accumulation of lipid in perirenal
adipocytes occurs during the allometric growth phase, whereas increased mitochondrial
biogenesis and sympathetic innervation of adipocytes occur during the isometric growth phase.
Therefore, even though prenatal growth of BAT is most rapid during mid-gestation, functional
development of brown adipocytes does not occur until late gestation.

4.1. Morphological changes during ontogeny


Based on histological examinations of bovine BAT during late gestation, fetal brown adipocytes
contain fewer mitochondria that are also less developed compared to brown adipocytes from
newborn calves (Landis et al., 2002; fig. 2). Mitochondria from fetuses sampled as late as
midway through the last trimester are spherical, with cristae that do not traverse the entire
mitochondrion, whereas brown adipocytes from late term and newborn BAT are nearly
unilocular, with small lipid inclusions peripheral to larger central lipid vacuoles, and the
mitochondria are more abundant and more fully differentiated (fig. 3). These observations
corroborate Nedergaard et al. (1986), who described similar morphological changes in brown
adipocytes during fetal development of rats.
At 96 days before birth, some BAT cells have accumulated very little lipid, whereas others
already are nearly unilocular (figs. 2A and D). Brown adipocytes gradually accumulate lipid,
with the multilocular appearance persisting in Brahman BAT at 24 days before birth (fig. 2E).
Ontogeny and metabolism of brown adipose tissue 307

Fig. 2. Transmission electron micrographs of perirenal brown adipocytes. Perirenal brown adipose tissue
was obtained by cesarean section at 96 and 24 days before birth, and at birth, from Angus (left column) and
Brahman (right column) fetuses. Thick arrows indicate lipid vacuoles and thin arrows point to mitochondria.
A nucleus is indicated at the bottom left. Scale bar at the bottom right indicates magnification. Adapted from
Figure 2 of Landis et al. (2002).

At birth, brown adipocytes become essentially unilocular, with only a few small lipid
vacuoles set apart from the large, central vacuole (figs. 2C and F).
At the earliest sampling period, BAT mitochondria are smaller, spherical, and quite dense
(figs. 3A and D). By 24 days before birth, the mitochondria become quite large, and cristae,
although distinct, are extensive (figs. 3B and E). Within 7 days of parturition, the morphology
of the mitochondria changes from primarily spherical to markedly elongated, with extensively
differentiated cristae. This morphology is apparent at birth (figs. 3C and F).
308 S. B. Smith and G. E. Carstens

Fig. 3. Transmission electron micrographs of mitochondria from perirenal brown adipocytes. Perirenal brown
adipose tissue was obtained by cesarean section at 96 and 24 days before birth, and at birth, from Angus
(left column) and Brahman (right column) fetuses. Arrows point to cristae and mitochondria are labeled M in
some panels. Scale bar at lower right indicates magnification. Adapted from Figure 3 of Landis et al. (2002).

4.2. Gene expression during ontogeny

Fetal bovine perirenal UCP1 mRNA is not detectable until day 211 of gestation, and only in
small quantities thereafter until day 259, when levels increase markedly (Casteilla et al.,
1989). We recently documented a depression and then recovery in UCP1 gene expression
during the last 14 days of gestation (fig. 4; Landis et al., 2002). The decline in UCP1 gene
expression coincided with the change in mitochondrial morphology from spherical to elongated,
and therefore may have biological significance.
Ontogeny and metabolism of brown adipose tissue 309

Fig. 4. UCP1 gene expression in perirenal and tailhead s.c. adipose tissues. Total RNA was extracted
from the perirenal (brown adipose tissue, BAT) and tailhead s.c. adipose tissue (white adipose tissue, WAT)
samples for slot blot analysis of uncoupling protein-1 (UCP1). Each data point represents the mean for 3 fetal
calves per breed. The overall SEM for each breed type is affixed to the symbols. Relative concentrations of
UCP1 mRNA are expressed as the ratio of UCP1:18S ribosomal RNA × 100. Adapted from Figure 5 of
Landis et al. (2002).

The activity of 5′-deiodinase first appears in fetal bovine perirenal adipose tissue at
2 months of gestation, increases rapidly until activity peaks at 7 months, and declines there-
after until birth (Giralt et al., 1989). The activity of 5′-deiodinase is responsible for
deiodination of T4 to T3, which is thought to be required for optimal synthesis of UCP1 in
cold-adapted rats (Bianco and Silva, 1988). Giralt et al. (1989) noted that 5′-deiodinase peaked
at approximately the same time that Casteilla et al. (1989) first detected UCP1 mRNA, and
hypothesized that endogenous production of T3 may be involved in prenatal induction of
UCP1 expression in BAT.
Trayhurn et al. (1993) demonstrated small amounts of UCP1 protein in goat s.c. adipose
tissue at 2.5 days of age. Consistent with this, our laboratory (Martin et al., 1999) demon-
strated that s.c. adipose tissue from newborn Brahman and Angus calves contained adipocytes
with distinctive brown adipocyte morphology. More recent results (Landis et al., 2002)
demonstrated UCP1 mRNA in s.c. adipose tissue during the last trimester of gestation,
although the abundance of UCP1 mRNA dropped precipitously during the last 30 days prior
to parturition (Fig. 4).
Casteilla et al. (1994) examined the expression of adrenergic receptor (β1- and β3-AR)
genes in bovine perirenal BAT during fetal and early postnatal development. The β3-AR
mRNA was first measurable during mid-gestation, and its concentration increased dramati-
cally between 6 months of fetal life and birth. In contrast, β1-AR mRNA was expressed at low
levels throughout fetal life. The appearance of β3-AR mRNA preceded the expression of
UCP1 mRNA (Casteilla et al., 1989), suggesting that β3-AR gene expression may be prereq-
uisite to BAT differentiation (Casteilla et al., 1994). Postnatally, β3-AR gene expression
310 S. B. Smith and G. E. Carstens

declined after approximately 3 months of age; the decline was much slower than the more
rapid postnatal decline of UCP1 mRNA (Nougues et al., 1993; Trayhurn et al., 1993).
Response to β3-adrenergic stimulus apparently persists after birth. Acute treatment of rats
with a selective β3-agonist increased BAT GDP-binding of mitochondria (a measure of UCP1
activity) after only 60 min of treatment (Milner et al., 1988). In adult dogs, treatment with a
β3-agonist increased UCP1 concentration and the expression of UCP1 mRNA (Champigny
et al., 1991). Administration of the β3-agonist ICI-D7114 to newborn lambs during the first
several weeks of life delayed the apparent involution of BAT to WAT (Nougues et al., 1993).
The β3-agonist-treated lambs continued to express UCP1 mRNA in perirenal and pericardial
fat depots at 25 days of age, whereas control lambs of the same age did not.

4.3. Fatty acid metabolism during ontogeny

The incorporation of acetate, glucose, and palmitate into glycerolipids of perirenal adipose
tissue decreases markedly during late gestation, especially in Brahman BAT (Landis et al.,
2002) (fig. 5). Vernon et al. (1981) demonstrated reductions in fatty acid synthesis from
acetate and glucose in fetal ovine perirenal BAT that were similar in magnitude to those

Fig. 5. Lipogenesis from acetate, glucose, and palmitate in perirenal brown adipose tissue as a function of
average fetal age. The incorporation rate is expressed as nmol substrate incorporated/106 cells/h. Each data
point represents the mean for 3 fetal calves per breed type. The rate of palmitate esterification in Brahman
fetuses at 96 days before birth was 1,570 nmol/106 cells/h (off scale in the figure). The overall SEM for each
breed type is affixed to the symbols. The SEM for glucose incorporation are not large enough to be visible.
There was a significant age effect for acetate incorporation into glycerolipids. There was a significant age ×
breed type effect for palmitate and glucose incorporation into glycerolipids. Adapted from Figure 6 of
Landis et al. (2002).
Ontogeny and metabolism of brown adipose tissue 311

observed in fetal calves. Although de novo fatty acid biosynthesis is barely detectable by
parturition in bovine BAT, glycerolipid synthesis from palmitate remains elevated at that time.
Thus, palmitate esterification accounts for 98% of total glycerolipid synthesis in vitro in the
newborn calves, and is a primary contributor to lipid filling throughout gestation.
Acetate has been well documented as the principal source of carbon for de novo fatty acid
biosynthesis in adipose tissue of young and adult ruminant species (Ballard et al., 1972; Ingle
et al., 1972; Smith and Prior, 1986). Both acetate and glucose contribute significantly to
de novo lipogenesis in newborn calves (Martin et al., 1999). At the beginning of the third
trimester in fetal calves, the rate of lipogenesis from acetate is approximately 10-fold greater
than lipogenesis from glucose. Some portion of the glucose carbon would have been recov-
ered as glyceride–glycerol (consistent with the high rates of palmitate incorporation into
lipids), so that actual rates of de novo fatty acid biosynthesis from glucose would be very low
relative to fatty acid biosynthesis from acetate. Thus, fetal bovine adipose tissue preferentially
uses acetate as the carbon source for de novo fatty acid biosynthesis early in the third
trimester of gestation.

5. ENVIRONMENTAL EFFECTS ON BROWN ADIPOSE TISSUE


5.1. Metabolism and thermogenesis during cold exposure

Cold-induced secretion of norepinephrine (NE) increases the specific thermogenic activity of


BAT. In addition to its role in acute activation of BAT thermogenesis, NE is also involved in
long-term modulation of BAT growth and development during cold stress by enhancing dif-
ferentiation of BAT precursor cells, mitochondrial proliferation, and transcription of UCP1 via
β- and α1-AR pathways (Géloën et al., 1988). Type II thyroxine 5′-deiodinase, which is an
important enzyme regulating the thermogenic capacity of BAT, is increased in BAT during
cold exposure (Puig-Domingo et al., 1989).
Previous studies suggested that the improved survival of lambs born to cold-exposed ewes
was due to an enhanced rate of BAT thermogenesis. Stott and Slee (1985) found that lambs
born to cold-exposed ewes exhibited significantly higher NE-induced thermogenic rates (in vivo
assessment of BAT thermogenesis) than lambs from warm-exposed ewes. Likewise, Symonds
et al. (1992) found that lambs from cold-exposed ewes were 15% heavier at birth, and pos-
sessed 21% more perirenal BAT that was also 40% more active thermogenically compared to
lambs from control ewes. Newborn lambs from cold-exposed ewes were clearly more cold-
tolerant as thermogenic rates were 16% greater in a warm (28°C) environment and 40% greater
in a cold (14°C) environment, relative to lambs from control ewes.
We investigated the postnatal changes in perirenal BAT in neonatal calves, and found that
cytochrome c oxidase activity of perirenal BAT was highest at birth, and decreased substan-
tially after 7 days of warm exposure (Carstens, 1994). Mitochondrial protein concentrations
in warm-exposed calves were only 20% of that found in newborn calves (Carstens, 1994), and
bovine perirenal UCP1 was reduced markedly in warm-exposed calves (Martin et al., 1997).
Trayhurn et al. (1993) reported a similar postnatal decline in cytochrome c oxidase activity
in neonatal goats (325, 50, and 3 μmol oxidized/min/g of BAT in newborn, 7-day-old, and
21-day-old goats, respectively). Cytochrome c oxidase activity was also lower in cold-exposed
calves than newborn calves, but was still 2.7-fold higher than in warm-exposed calves at
7 days of age (Carstens, 1994). Thus, cold exposure during the early postnatal period delays
the apparent involution of BAT to WAT, resulting in higher BAT thermogenic rates during the
neonatal period.
312 S. B. Smith and G. E. Carstens

5.2. Involution during warm exposure

Postnatally, in neonatal ruminants BAT is involuted into WAT within 2 to 3 weeks if the
neonate is not exposed to cold (Alexander et al., 1975). In bovine perirenal BAT, there is no
apparent atrophy or degeneration of brown adipocytes postnatally (Alexander et al., 1975).
Instead, there appears to be a progressive accumulation of lipid, accompanied by the dedif-
ferentiation and loss of mitochondria, as perirenal adipocytes acquire the morphology of
white adipocytes. More recently, Trayhurn et al. (1993) demonstrated the presence of UCP
protein (measured by Western analysis) in s.c. (hindlimb and neck) and internal (perirenal,
pericardial, and omental) adipose tissues of goats. Trayhurn et al. (1993) detected UCP
protein in s.c. adipose tissue of 7-day-old goats, and in perirenal BAT of 21-day-old goats.
Uncoupling protein is more persistent than its mRNA; uncoupling protein mRNA was not
detectable by day 2.5 in goat perirenal BAT, and was undetectable in other adipose tissue
depots at birth.
We examined the effects of postnatal cold exposure on NE-induced BAT thermogenesis
in Holstein calves in vivo (table 1). Peak metabolic (PM) rates were 26% lower in warm-
exposed calves than in newborn calves, yet PM rates were similar between newborn and
cold-exposed calves, demonstrating that cold exposure delayed the apparent postnatal invo-
lution of BAT. This observation is supported by the fact that cytochrome c oxidase activity
and total mitochondrial protein in BAT were 2.7- and 2.5-fold higher in cold-exposed than in
warm-exposed calves.
We measured UCP1 mRNA in a small number of BAT samples from newborn calves that
had been exposed to 4°C for 7 days postnatally. As observed for cytochrome c oxidase activ-
ity and mitochondrial protein, there was a dramatic loss of UCP1 mRNA postnatally even
during cold exposure (fig. 6). We also recently examined the effects of postnatal cold expo-
sure on BAT thermogenesis in newborn lambs (table 2). Brown adipose tissue mass tended to
be greater in the lambs held for 48 h at 28°C than in lambs held at 6°C. Similarly, total and
mitochondrial BAT protein concentrations were increased significantly by cold exposure.
This observation is supported by the fact that cytochrome c oxidase activity and UCP1 mRNA
were approximately 3-fold and 5-fold higher, respectively, in BAT from cold-exposed lambs
than in warm-exposed lambs (table 2).

Table 1
Effects of cold exposure on postnatal changes in thermoneutral (TM) and NE-induced peak
metabolic rates (PM), and BAT cytochrome c oxidase activity and mitochondrial protein in
newborn Holstein, and 7-day-old calves exposed to 4°C (cold-exposed) or 22°C (warm-exposed)
temperatures from birth to 7 days of age
Ontogeny and metabolism of brown adipose tissue 313

Fig. 6. Uncoupling protein-1 (UCP1) mRNA in


newborn and 7-day cold-adapted calves. Lane 1,
bovine longissimus muscle RNA. Lanes 2 and 3,
RNA from BAT of newborn calves. Lanes 4 and 5,
RNA from BAT of calves that had been subjected to
4°C for 7 days postnatally. No bands are visible in
lane 5. Lane 6, DNA from a PCR reaction that used
the calf UCP1 cDNA as template. The upper band
is the UCP-PCR probe. The lower band corresponds
to unincorporated primers. Lane 7, Eco RI-excised
calf UCP partial cDNA (1.4 kb). (S.B. Smith and
G.E. Carstens, unpublished data.)

The reduction in BAT mass, and concomitant elevation in BAT protein content, were the
result of cold-induced mobilization of BAT lipid stores, which caused a remarkable delipida-
tion of BAT in response to cold exposure (fig. 7). Mitochondria from warm-exposed lambs
underwent extensive dedifferentiation, and by this stage were structurally similar to mito-
chondria from fetal calves (fig. 3), i.e. they were larger with less extensive cristae. There also
was some apparent dedifferentiation of mitochondria in the cold-exposed lambs. The loss of
mitochondrial integrity could have been the result of dedifferentiation (involution).
Alternatively, brown adipocytes in warm-exposed lambs may have been undergoing apoptosis,
to be replaced subsequently with white adipocytes.

6. APOPTOSIS OF BROWN ADIPOCYTES


It is not clear if BAT from ruminant species undergoes dedifferentiation and direct conversion
to WAT, or whether brown adipocytes experience apoptosis, to be replaced by newly differ-
entiated white adipocytes. In bovine perirenal BAT, there is no apparent atrophy or
degeneration of brown adipocytes postnatally (Alexander et al., 1975). Instead, there appears
to be a progressive accumulation of lipid, accompanied by the dedifferentiation and loss of

Table 2
Effects of cold exposure (28°C vs 6°C) on perirenal adipose tissue composition, cytochrome c
oxidase activity, mitochondrial protein, and UCP mRNA in 48-hour-old lambs

Warm-exposed Cold-exposed P<

Birth weight, kg 4.18 ± 0.17 4.06 ± 0.15 NS


BAT, g/kg body weight 4.40 ± 0.27 2.69 ± 0.14 0.10
BAT protein, mg/g 97.6 ± 2.8 131.2 ± 4.0 0.001
Cytochrome c oxidase activity, μmol/g/min 57.7 ± 5.0 153.6 ± 7.5 0.001
Mitochondrial protein, mg/g BAT 13.06 ± 1.4 26.65 ± 1.72 0.001
UCP mRNA:28S rRNA ratio 0.10 ± 0.05 0.64 ± 0.11 0.05
n = 20 lambs/treatment.
NS = not statistically different (P > 0.05).
UCP:28S ratio = (Laser densitometer area for UCP mRNA: Laser densitometer area for 28S rRNA) × 100. Areas
were calculated from slot blots of 5 μg total RNA, hybridized either to a 32P-labeled 300-bp PCR-generated UCP
probe or a 32P-labeled cDNA for the rat 28S ribosomal subunit. (S. B. Smith and G. E. Carstens, unpublished data.)
314 S. B. Smith and G. E. Carstens

Fig. 7. Brown adipose tissue from newborn lambs exposed to 28°C (left) or 6°C (right) for 48 h postnatally.
Thick arrows indicate lipid droplets and thin arrows indicate mitochondria. Adipocytes from warm-exposed
lambs are multilocular, whereas adipocytes from cold-exposed lambs are nearly devoid of lipid. Mitochondria
are less numerous in BAT from warm-exposed lambs. Cristae structure is becoming disrupted in BAT from
warm-exposed lambs, indicating dedifferentiation of brown adipocytes. Scale bars are indicated for the lowest
and highest magnifications. (S.B. Smith and G.E. Carstens, unpublished data.)

mitochondria, as perirenal adipocytes acquired the morphology of white adipocytes. These


data provided evidence to suggest that, prenatally, adipocytes from all depots may initially
differentiate as BAT, and that they subsequently involute to acquire WAT morphological char-
acteristics. Our histological examination of Angus and Brahman s.c. adipose tissue (Martin
et al., 1999; fig. 8) tends to support this postulate. Although most adipocytes have already
acquired WAT characteristics by parturition, we located several adipocytes with distinctly
Ontogeny and metabolism of brown adipose tissue 315

Fig. 8. Subcutaneous adipose tissue from Angus (left column) and Brahman (right column) newborn calves.
Cells with brown adipocyte morphology (arrows) and white adipocyte morphology (arrowheads) are apparent.
A capillary endothelial cell (CE) is centrally located in the Brahman sample. Mitochondrial size and cristae
density (bottom panels) are indicative of brown adipocytes, with no apparent difference between breed types.
Scale bars are indicated for the lowest and highest magnifications. Adapted from Figure 5 of Martin et al.
(1999).

brown adipocyte morphology, i.e. extensive, highly differentiated mitochondria surrounding


a smaller lipid vacuole. This could be interpreted to mean that s.c. adipose tissue initially dif-
ferentiated as BAT, and we are observing the involution of BAT to WAT initiated prenatally.
However, we cannot rule out the possibility that brown adipocytes were lost to the total
population of adipocytes via apoptosis, a process not detectable in transmission electron
photomicrographs.
Lindquist and Rehnmark (1998) demonstrated that transferring cold-adapted mice to 28°C
caused a rapid increase in the rate of apoptosis in interscapular BAT. Murine brown
316 S. B. Smith and G. E. Carstens

adipocytes exposed to NE in culture exhibited a 50% decline in DNA fragmentation (i.e.


apoptosis) (Lindquist and Rehnmark, 1998). The latter results demonstrate the involvement
of sympathetic innervation of BAT in maintaining the differentiated state. These data also
suggest that apoptosis in BAT may have occurred in our investigation of lambs held at warm
vs cold temperatures (Smith et al., 2004). The maintenance of the BAT differentiated state by
NE has been demonstrated previously. Norepinephrine induces the expression of UCP1 and
may increase of number of brown adipocytes (i.e. be mitogenic; Nedergaard et al., 1995).
Conversely, cold exposure maintains the BAT viability by reducing the rate of DNA and
protein degradation (Desautels and Heal, 1999).

7. NUTRITION AND BROWN ADIPOSE TISSUE THERMOGENESIS


7.1. Dietary protein restriction

Malnutrition of the dam during late gestation has been shown to reduce neonatal calf survival
(Hight, 1966; Corah et al., 1975). The inability of the neonate to maximize thermogenesis in
response to cold stress during the early postnatal period may be caused by prepartum protein
and/or energy malnutrition. Previous studies have demonstrated that prepartum protein
(Carstens et al., 1987) and energy (Ridder et al., 1991) restriction of nulliparous heifers
reduced thermoneutral metabolism in newborn calves. Carstens et al. (1987) reported that
prepartum protein restriction reduced thermoneutral metabolic rates by 11.4%, even though
birth weights were unaffected by prepartum protein treatment.
Unlike previous reports, the thermoneutral metabolic rate of Wagyu × Angus newborn
calves was not affected by prepartum protein restriction (Martin et al., 1997). Consistent with
the lack of a treatment effect on peak metabolic rates, prepartum protein restriction did not
affect perirenal adipose tissue mass or composition. Nor did prepartum protein restriction
alter UCP1 gene expression in BAT. Alexander (1978) fed high- and low-energy diets to
pregnant ewes, beginning on day 90 of gestation, and found that prepartum energy restriction
reduced the proportional weight of perirenal adipose tissue (the primary BAT depot in
newborn ruminants) by 17% in single and 24% in twin fetuses at 125 days of gestation.
Furthermore, Tyzbir (1984) demonstrated that prepartum protein restriction of rats reduced
BAT mass by 40–50% as well as BAT mitochondrial thermogenic capacity in newborn
rat pups, even though birth weights were not affected by prepartum protein restriction. In
contrast, the NE-induced peak metabolic rate was the same in calves born to adequate- and
restricted-protein heifers in the investigation of Martin et al. (1997).
The results of Martin et al. (1997) may have been confounded by the unusual breed type
of calves used in that study (Wagyu × Angus crossbred calves). Wagyu calves have lower
birth weights than Angus calves (Smith et al., 1992) and, unlike Angus or Brahman purebred
calves, subcutaneous adipose tissue of Wagyu crossbred calves contains no detectable brown
adipocytes at birth (Martin et al., 1997, 1999). Wagyu calves represent a distinct genetic line
resulting from crossbreeding of native Japanese cattle in the mid-nineteenth century (Smith
et al., 2001). Thus, the lack of effect of protein malnutrition on fetal calf thermogenesis
reported by Martin et al. (1997) should not be considered typical for this species.

7.2. Essential fatty acid supplementation

Several studies have shown that brown fat thermogenic rates were higher and sympathetic activ-
ity of brown fat increased (higher norepinephrine turnover rates) in rats fed diets containing
Ontogeny and metabolism of brown adipose tissue 317

safflower oil (high in polyunsaturated fatty acids; PUFA) compared to rats fed diets containing
beef tallow (high in saturated fatty acids; SFA) (Takeuchi et al., 1995a; Matsuo et al., 1995).
Additionally, Takeuchi et al. (1995b) found that serum T3 levels were higher in rats fed a
safflower-oil diet compared to those fed a beef tallow-fat diet. Lammoglia et al. (1999) demon-
strated that prenatal supplementation of cracked safflower seeds to pregnant cows affected cold
tolerance of newborn calves (fig. 9). Calves were exposed to cold ambient temperatures starting
at 4 h of age. Rectal temperatures in the cold were significantly higher in calves born to cows sup-
plemented with safflower seeds than calves born to cows fed the control diet containing no added
fat. Taken together, these observations suggest that maternal supplementation of a bypass source
of PUFA to pregnant ewes may have the potential to enhance fetal BAT development.
Additionally, studies using fish oils have indicated that the n-3 PUFA eicosapentaenoic
acid (20:5n-3; EPA) and docosahexaenoic acid (22:6n-3; DHA) also are effective in stimu-
lating BAT thermogenesis (Sadurskis et al., 1995; Oudart et al., 1997; Saha et al., 1998;
Kawada et al., 1998). Polyunsaturated fatty acids, and especially n-3 PUFA, stimulate non-
shivering thermogenesis in rodents, and they may do so by increasing NE turnover rate in
BAT. Exposure of BAT to NE reduces the extent of apoptosis in response to warm exposure,
which in turn should cause elevated thermogenesis relative to animals fed beef tallow or no
added n-3 PUFA. Therefore, n-3 PUFA may increase thermogenesis by depressing apoptosis,
or even stimulating brown adipocyte differentiation. This is in contrast to results with splenic
lymphocytes (Avula et al., 1999) and tumor cells (Das, 1999), in which n-3 PUFA promoted
apoptosis, and indicates a tissue-specific effect of n-3 PUFA on BAT.
We recently fed pregnant ewes 2%, 4%, or 8% rumen-protected fat. The fat sources were
high in either saturated/monounsaturated fatty acids or n-3 PUFA (formaldehyde-protected
soy/linseed lipid). The PUFA-fed ewes had higher plasma concentrations of 18:2, 18:3n-3,
and EPA, and lower concentrations of 16:0, 16:1, and 18:1, than ewes fed the saturated/
monounsaturated fatty acid diet. The BAT of lambs born to PUFA-fed ewes had higher

Fig. 9. Effects of feeding supplemental fat (safflower seeds) during late gestation on cold tolerance of
newborn calves. Adapted from Figure 1 of Lammoglia et al. (1999).
318 S. B. Smith and G. E. Carstens

concentrations of 18:2, EPA, and DHA than lambs born to ewes fed the saturated/monoun-
saturated fatty acid diet. However, BAT mass, cytochrome c oxidase activity, and GDP binding
were not affected by level or source of dietary fat. Cold-induced rectal temperature responses
of lambs were not affected by source of prenatal fat. Therefore, unlike results with calves
(Lammoglia et al., 1999), prenatal PUFA supplementation did not affect BAT thermogenic
activity or cold tolerance of newborn lambs. The effects of DHA or EPA on BAT thermogen-
esis in lambs cannot be tested until a rumen-bypass source of these fatty acids is developed.

7.3. Copper supplementation

Although few studies have investigated the impact of prenatal dietary copper on thermo-
metabolism, clinical evidence of a link between copper deficiency and cold intolerance exists
in lambs. Copper plays an essential role in several copper-dependent enzyme systems that
regulate thermometabolism, including cytochrome c oxidase and dopamine-β-hydroxylase.
Cytochrome c oxidase is the terminal enzyme in the electron transport system linking
substrate oxidation to oxidative phosphorylation (ATP synthesis) in mitochondria. Another
copper-dependent enzyme, dopamine-β-hydroxylase, regulates the synthesis of NE from
dopamine in the sympathetic nervous system.
We examined the effects of prenatal dietary copper level on thermometabolism in lambs
(Carstens et al., 1999). Twin-bearing ewes were assigned to low- or high-copper treatments
during the last trimester of gestation. Even though liver copper concentrations in newborn lambs
were reduced 57% by the low-copper treatment (132 vs. 306 ppm copper DM), these
lambs would not be classified as being copper-deficient. Despite the fact that the low-copper
lambs were not copper-deficient, their rectal temperatures at 2 h of age were 3.3°F lower
than lambs born to high-copper ewes (fig. 10). We subsequently found that NE turnover
rates in BAT of lambs at 12 h of age were decreased by the low-copper treatment (0.16 vs
0.3 ng NE/mg BAT/h). This suggests that the low-copper treatment decreased dopamine-
β-hydroxylase enzyme activity which impaired the thermogenic function of BAT. Additional
evidence to support this idea is the finding that low-copper lambs also had lower plasma T3
levels compared to high-copper lambs (fig. 10), even though plasma T4 levels were unaffected
by prenatal copper treatment. An important regulatory aspect of BAT thermogenesis is the
activation of 5′-deiodinase by NE release from the sympathetic nervous system in response
to cold stress. Locally synthesized T3 is a potent regulator of uncoupling protein gene expres-
sion in BAT. Because more than 60% of circulating T3 levels in newborn lambs are derived
from the conversion of T4 to T3 in peripheral tissues by 5′-deiodinase enzyme (Klein et al.,
1980), the fact that plasma T3 levels were depressed in low-copper lambs suggests that
5′-deiodinase activity may have been impaired by the low-copper treatment indirectly through
a reduction in NE stimulation.

8. FUTURE PERSPECTIVES
Research continues to describe the development and metabolism of brown adipose tissue
in lambs and calves. This is important in respect to both the welfare of the newborn animals
and the economic impact of neonatal mortality to the livestock industry. Ideally, production
strategies such as supplemental n-3 PUFA or copper would improve BAT functionality and
thereby increase newborn lamb or calf survival. Conversely, strategies to increase BAT mass
in the neonate may lead to increased adiposity in the mature animal if brown adipocytes
dedifferentiate into white adipocytes during development.
Fig. 10. Effect of prenatal copper treatment on rectal temperature (left panel) and plasma triiodothyronine concentrations (right panel) in newborn lambs. From Figure 1 of
Carstens et al. (1999).
320 S. B. Smith and G. E. Carstens

REFERENCES
Alexander, G., 1978. Quantitative development of adipose tissue in foetal sheep. Aust. J. Biol. Sci. 31,
489–503.
Alexander, G., Bennett, J.W., Gemmell, R.T., 1975. Brown adipose tissue in the new-born calf
(Bos taurus). J. Physiol. 244, 223–234.
Avula, C.P., Zaman, A.K., Lawrence, R., Fernandes, G., 1999. Induction of apoptosis and apoptotic
mediators in Balb/C splenic lymphocytes by dietary n-3 and n-6 fatty acids. Lipids 34, 921–927.
Azzam, S.M., Kinder, J.E., Nielsen, M.K, 1993. Environmental effects on neonatal mortality of beef
calves. J. Anim. Sci. 71, 282–290.
Ballard, F.J., Filsell, O.H., Jarrett, I.G., 1972. Effects of carbohydrate availability on lipogenesis in sheep.
Biochem. J. 226, 193–200.
Bianco, A.C., Silva, J.E., 1988. Cold exposure rapidly induces virtual saturation of brown adipose tissue
nuclear T3 receptors. Amer. J. Physiol. 255, E496–E503.
Brander, F., Keith, J.S., Trayhurn, P., 1993. A 27-mer oligonucleotide probe for the detection and meas-
urement of the mRNA for uncoupling protein in brown adipose tissue of different species. Comp.
Biochem. Physiol. 104B, 125–131.
Carstens, G.E., 1994. Cold thermoregulation in newborn calves. Vet. Clin. N. Amer.-Food Anim. Pr. 10,
69–106.
Carstens, G.E., Eckert, J.C., Greene, L.W., Smith, S.B., 1999. Effects of prenatal dietary copper on
endocrine control of brown fat thermogenesis in newborn lambs. In: Proceedings of IX Symposium
Ruminant Physiology. S. Afr. J. Anim. Sci. 29, 429–430.
Carstens, G.E., Johnson, D.E., Holland, M.D., Odde, K.G., 1987. Effects of prepartum protein nutrition
and birth weight on basal metabolism in bovine neonates. J. Anim. Sci. 65, 745–751.
Carstens, G.E., Mostyn, P.C., Chapman, S.A., Randel, R.D., 1995. Prenatal and postnatal changes in
brown adipose tissue thermogenesis in the neonatal calf. In: Energy Metabolism of Farm Animals.
EAAP Publ. No. 76. Pudoc, Wageningen, The Netherlands, pp. 101–104.
Casteilla, L., Champigny, O., Bouillaud, F., Robelin, J., Ricquier, D., 1989. Sequential changes in the
expression of mitochondrial protein mRNA during the development of brown adipose tissue in bovine
and ovine species: sudden occurrence of uncoupling protein mRNA during embryogenesis and its dis-
appearance after birth. Biochem. J. 257, 665–671.
Casteilla, L., Muzzin, P., Revelli, J.P., Ricquier, D., Giacobino, J.P., 1994. Expression of β1- and
β3-adrenergic-receptor messages and adenylate cyclase β-adrenergic response in bovine perirenal
adipose tissue during its transformation from brown to white fat. Biochem. J. 297, 93–97.
Champigny, O., Ricquier, D., Blondel, O., Mayers, R.M., Briscoe, M.G., Holloway, B.R., 1991. β3 adren-
ergic receptor stimulation restores message and expression of brown-bat mitochondrial uncoupling in
adult dogs. Proc. Natl. Acad. Sci. USA 88, 1074–1077.
Corah, L.R., Dunn, T.G., Kaltenbach, C.C., 1975. Influence of prepartum nutrition on the reproductive
performance of beef females and the performance of their progeny. J. Anim. Sci. 41, 819−824.
Das, U.N., 1999. Essential fatty acids, lipid peroxidation and apoptosis. Prostagland. Leuk. Essent. Fatty
Acids 61, 157–163.
Desautels, M., Heal, S., 1999. Differentiation-dependent inhibition of proteolysis by norepinephrine in
brown adipocytes. Amer. J. Physiol. 277, E215–Ε222.
Géloën, A., Collet, A.J., Guay, G., Bukowiecke, L.J., 1988. β-adrenergic stimulation of brown adipocyte
proliferation. Am. J. Physiol. 254, C175–C182.
Giralt, M., Casteilla, L., Viñas, O., Mampel, T., Iglesias, R., Robelin, J., Villarroya, R., 1989. Iodothyronine
5′-deiodinase activity as an early event of prenatal brown-fat differentiation in bovine development.
Biochem. J. 259, 555–559.
Hight, G.K., 1966. The effects of undernutrition in late pregnancy on beef cattle production. NZ J. Agr. Res.
9, 479–490.
Himms-Hagen, J., 1986. Brown adipose tissue and cold-acclimation. In: Trayhurn, T., Nicholls, D.G. (Eds.),
Brown Adipose Tissue. Edward Arnold, Baltimore, MD, pp. 214–268.
Ingle, D.L., Bauman, D.E., Garrigus, U.S., 1972. Lipogenesis in the ruminant: in vivo site of fatty acid
synthesis in sheep. J. Nutr. 102, 617–624.
Kawada, T., Kayahashi, S., Hida, Y., Koga, K., Nadachi, Y., Fushiki, T., 1998. Fish (Bonito) oil supple-
mentation enhances the expression of uncoupling protein in brown adipose tissue. J. Agr. Food Chem.
46, 1225–1227.
Ontogeny and metabolism of brown adipose tissue 321

Klein, A.H., Oddie, T.H., Fisher, D.A., 1980. Iodothyronine kinetic studies in the newborn lamb. J. Dev.
Physiol. 2, 29–36.
Lammoglia, M.A., Bellows, R.A., Grings, E.E., Bergman, J.W., 1999. Effects of prepartum supplemen-
tary fat and muscle hypertrophy genotype on cold tolerance in newborn calves. J. Anim. Sci. 77,
2227–2233.
Landis, M.D., Carstens, G.E., McPhail, E.G., Randel, R.D., Green, K.K., Slay, L., Smith, S.B., 2002.
Ontogenic development of brown adipose tissue in Angus and Brahman fetal calves. J. Anim. Sci. 80,
591–601.
Lindquist, J.M., Rehnmark, S., 1998. Ambient temperature regulation of apoptosis in brown adipose
tissue. J. Biol. Chem. 273, 30147–30156.
Martin, G.S., Carstens, G.E., Taylor, T.L., Eli, A.G., Tarrant, M., Britain, K., Smith, S.B., 1999.
Metabolism and morphology of brown adipose tissue from Brahman and Angus newborn calves.
J. Anim. Sci. 77, 388–399.
Martin, G.S., Carstens, G.E., Taylor, T.L., Sweatt, C.R., Eli, A.G., Lunt, D.K., Smith, S.B., 1997.
Prepartum protein restriction does not alter norepinephrine-induced thermogenesis or brown adipose
tissue function in newborn calves. J. Nutr. 127, 1929–1937.
Matsuo, T., Shimomura, Y., Saitoh, S., Tokuyama, K., Takeuchi, H., Suzuki, M., 1995. Sympathetic
activity is lower in rats fed a beef tallow diet than in rats fed a safflower diet. Metabolism 44,
934–939.
Mersmann, H.J., 1996. Evidence of classic beta β3-adrenergic receptors in porcine adipocytes. J. Anim.
Sci. 74, 984–992.
Milner, R.E., Wilson, S., Arch, J.R.S., Trayhurn, P., 1988. Acute effects of a β-adrenoceptor agonist (BRL
26830A) on rat brown adipose tissue mitochondria: increased GDP binding and GDP-sensitive
proton conductance without changes in the concentration of uncoupling protein. Biochem. J. 249,
759–763.
Moulin, K., Arnaud, E., Nibbelink, M., Viguerie-Bascands, N., Pénicaud, L., Casteilla, L., 2001a.
Cloning of BUG demonstrates the existence of a brown adipocyte distinct from a white one. Int.
J. Obes. 25, 1413–1441.
Moulin, K., Truel, N., André, M., Arnaud, E., Nibbelink, M., Cousin, B., Dani, C., Pénicaud, L, Casteilla, L.,
2001b. Emergence during development of the white-adipocyte cell phenotype is independent of the
brown-adipocyte phenotype. Biochem. J. 356, 659–664.
Napolitano, L., 1963. The differentiation of white adipose cells: an electron microscopy study. J. Cell
Biol. 18, 663–679.
Nedergaard, J., Connolly, E., Cannon, B., 1986. Brown adipose tissue in the mammalian neonate.
In: Trayhurn, P., Nicholls, D.G. (Eds.), Brown Adipose Tissue. Edward Arnold, London, pp. 152–213.
Nedergaard, J., Herron, D., Jacobsson, A., Rehnmark, S., Cannon, B., 1995. Norepinephrine as a morphogen?
Its unique interaction with brown adipose tissue. Int. J. Dev. Biol. 39, 827–837.
Nougues, J., Reyne, Y., Champigny, O., Holloway, B., Casteilla, L., Ricquier, D., 1993. The β3-adrenoceptor
agonist ICI-D7114 is not as efficient on reinduction of uncoupling protein mRNA in sheep as it is
dogs and smaller species. J. Anim. Sci. 71, 2388–2394.
Okamoto, M., Robinson, J.B., Christopherson, R.J., Young, B.A., 1986. Summit metabolism of newborn
calves with and without colostrum feeding. Can. J. Anim. Sci. 66, 937–944.
Oudart, H., Groscolas, R., Calgari, C., Nibbelink, M., Leray, C., Le Mayo, Y., Malan, A., 1997. Brown
fat thermogenesis in rats fed high-fat diets enriched with n-3 polyunsaturated fatty acids. Int. J. Obes.
21, 955–962.
Puig-Domingo, M., Guerrero, J.M., Vaughan, M.K., Little, J.C., Reiter, R.J., 1989. Activation of cere-
brocortical type II 5′-deiodinase activity in Syrian hamsters kept under short photoperiod and reduced
ambient temperature. Brain Res. Bull. 22, 975–979.
Ridder, T.A., Young, J.W., Anderson, K.A., Lodman, D.W., Odde, K.G., Johnson, D.E., 1991. Effects of
prepartum energy nutrition and body condition on birthweight and basal metabolism in bovine
neonates. J. Anim. Sci. 69, Suppl. 1, 450.
Robinson, J.B., Young, B.A., 1988a. Metabolic heat production of neonatal calves during hypothermia
and recovery. J. Anim. Sci. 66, 2538–2544.
Robinson, J.B., Young, B.A., 1988b. Recovery of neonatal lambs from hypothermia with thermal assis-
tance. Can. J. Anim. Sci. 68, 183–190.
Sadurskis, I., Dicker, A., Cannon, B., Nedergaard, J., 1995. Polyunsaturated fatty acids recruit brown
adipose tissue: increased UCP content and NST capacity. Amer. J. Physiol. 269, E351–E360.
322 S. B. Smith and G. E. Carstens

Saha, S.K., Ohinata, H., Ohno, T., Kuroshima, A., 1998. Thermogenesis and fatty acid composition of
brown adipose tissue in rats rendered hyperthyroid and hypothyroid – with special reference to
docosahexanoic acid. Jpn. J. Physiol. 48, 355–364.
Smith, S.B., Carstens, G.E., Randel, R.D., Mersmann, H.J., Lunt, D.K., 2004. Brown adipose tissue
development and metabolism in ruminants. J. Anim. Sci. 82, 942–954.
Smith, S.B., Prior, R.L., 1986. Comparison of lipogenesis and glucose metabolism between ovine and
bovine adipose tissues. J. Nutr. 116, 1279–1286.
Smith, S.B., Sanders, J.O., Lunt, D.K., 1992. Evaluation of birth and weaning characteristics of halfblood
and three-quarter blood Wagyu-Angus calves. McGregor Field Day Report, Texas Agric. Exp. Station
Tech. Rep. 92-1, pp. 60–64.
Smith, S.B., Zembayashi, M., Lunt, D.K., Sanders, J.O., Gilbert, C.D., 2001. Carcass traits and
microsatellite distributions in offspring of sires from three geographical regions of Japan. J. Anim.
Sci. 79, 3041–3051.
Stott, A.W., Slee, J., 1985. The effect of environmental temperature during pregnancy on thermoregula-
tion in the newborn lamb. Anim. Prod. 41, 341–347.
Symonds, M.E., Bryant, M.J., Clarke, L., Darby, C.J., Lomax, M.A., 1992. Effect of maternal cold expo-
sure on brown adipose tissue and thermogenesis in the neonatal lamb. J. Physiol. 455, 487–502.
Takeuchi, J., Matsuo, T., Tokuyama, K., Shimomura, Y., Suzuki, M., 1995a. Diet-induced thermogenesis
is lower in rats fed a lard diet than in those fed a high oleic sunflower oil diet, a safflower oil diet or
a linseed oil diet. J. Nutr. 125, 920–925.
Takeuchi, J., Matsuo, T., Tokuyama, K., Suzuki, M., 1995b. Serum triiodothyronine concentration and
Na+,K+-ATPase activity in liver and skeletal muscle are influenced by dietary fat type in rats. J. Nutr.
125, 2364–2369.
Trayhurn, P., Thomas, M.E.A., Keith, J.S., 1993. Postnatal development of uncoupling protein, uncou-
pling protein mRNA and GLUT4 in adipose tissues of goats. Amer. J. Physiol. 265, R676–R682.
Tyzbir, R.S., 1984. Altered brown adipose tissue mitochondrial function in neonates born to rats overfed
foods of various protein contents. J. Nutr. 114, 234–237.
Vernon, R.G., Robertson, J.P., Clegg, R.A., Flint, D.J., 1981. Aspects of adipose-tissue metabolism in
foetal lambs. Biochem. J. 196, 819–824.
13 Interorgan lipid and fatty acid metabolism
in growing ruminants

J. K. Drackley

Department of Animal Sciences, University of Illinois,


Urbana, IL 61801, USA

Lipid metabolism is a dynamic and critically important function in growing ruminants.


Preruminants consume a high-fat diet and deposit much of the dietary lipid in adipose tissue.
During postweaning growth, dietary fatty acid intake is low and long-chain fatty acid syn-
thesis increases in adipose tissue depots as the animal approaches physiological maturity.
Deposition of long-chain fatty acids synthesized within adipose or taken up from blood
lipoproteins is extensive in near-mature fattening ruminants. Lipoprotein lipase is a key regu-
latory enzyme in determining interorgan disposition of long-chain fatty acids from circulating
chylomicrons and very low-density lipoproteins in growing ruminants. Considerable research
in recent years has characterized the regulation of fatty acid esterification in adipose tissue.
The role of molecules such as leptin and tumor necrosis factor α, which are synthesized and
secreted by adipose tissue, is only beginning to be elucidated in growing ruminants. The type
of dietary fat affects lipid metabolism in liver of preruminants, but surprisingly little is known
about the developmental changes in hepatic metabolism of fatty acids in growing ruminants.
Although much recent progress has been made in understanding regulation of interorgan lipid
metabolism in growing ruminants, many fundamental questions remain.

1. INTRODUCTION
Lipid metabolism plays a dynamic role during growth in ruminant animals. Lambs and calves
are born with minimal body lipid, accrete body lipid rapidly during suckling of fat-rich milk,
undergo minimal fat deposition during the growth phase after weaning, and then again change
to fat deposition as the animal approaches physiological maturity. During the suckling or
milk-feeding period, the young ruminant (preruminant) functions as a monogastric animal
and largely deposits the long-chain fatty acids (LCFA) from milk fat into adipose tissue for
storage. As skeletal and muscle growth near completion, adipose storage of triacylglycerol (TG)
increases from a combination of both de novo lipogenesis (primarily from ruminally derived
acetate) and deposition of absorbed LCFA. Use of LCFA for fuel in well-fed ruminants is

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
323 © 2005 Elsevier Limited. All rights reserved.
324 J. K. Drackley

relatively minimal, except in heart tissue and in skeletal muscle during exercise. Interorgan
transport of LCFA is accomplished both by circulation of free or nonesterifed fatty acids
(NEFA) and by the various classes of plasma lipoproteins.
Metabolism of LCFA and other lipids assumes obvious importance to the growing ruminant
as a source of membrane components, signaling molecules, and a reserve of readily available
energy. Lipid metabolism occupies a central position in the determination of energetic effi-
ciency of growth and as a result has major impact on the profitability of meat animal
production. Moreover, content and composition of carcass lipid has become an increasingly
important consideration for consumers of ruminant animal products. Consequently, ruminant
lipid metabolism remains of major importance.
This chapter highlights some aspects where recent research has improved our understand-
ing of lipid metabolism in growing ruminants. The major focus is on the interorgan
relationships of lipid metabolism during growth, and is not intended to be a comprehensive
and exhaustive review of the literature. A number of excellent comprehensive and authorita-
tive reviews on various aspects of lipid metabolism in ruminants are available (e.g. Noble,
1978; Bell, 1980; Vernon, 1980; Noble and Shand, 1982; Bauchart, 1993; Chilliard, 1993;
Jenkins, 1993) and key reviews are cited where applicable. For a general discussion of lipid
metabolism in domestic animals, see Drackley (2000).

2. DIGESTION AND ABSORPTION OF DIETARY LIPIDS


2.1. Preruminants

Preruminants fed milk or milk replacer consume a relatively high-fat diet. Bovine milk con-
tains 29–31% fat on a dry solids basis; milk of goats and sheep may contain about 34% and
40% of the solids as fat, respectively (Jenness, 1985). Most commercial milk replacers con-
tain between 12% and 20% fat on a dry solids basis (Davis and Drackley, 1998). Dietary fat
in milk or milk replacers consists primarily of TG, which are digested in the small intestine
and packaged into chylomicrons for distribution throughout the body. The LCFA of dietary
origin are delivered to tissues for oxidative use (primarily heart and skeletal muscle) or for
deposition in adipose tissue. The remaining components of the chylomicron particle (choles-
terol, cholesterol esters, phospholipids, and apoproteins) participate in additional cycles of
lipoprotein metabolism to distribute cholesterol and essential fatty acids (EFA) throughout
the body for cell membrane and steroid hormone biosynthesis.
Preruminants digest the lipids in milk with high efficiency (typically greater than 97%;
Toullec and Mathieu, 1969). A number of other fat sources in milk replacers can be well
digested (>90%) by preruminants if properly emulsified, including tallow, lard, and coconut
oil (Toullec and Mathieu, 1969; Davis and Drackley, 1998). The milk fat of ruminants con-
tains a relatively large proportion of short- and medium-chain fatty acids, probably as a strategy
to maintain fluidity of the fat in the face of the mostly saturated LCFA that reach the maternal
duodenum as a result of ruminal biohydrogenation of polyunsaturated LCFA consumed from
the herbivorous diet.
Milk fat is entrapped in the casein coagulum in the abomasum, and is released as the casein
clot undergoes initial digestion. Within the abomasum, fat digestion is initiated by action of
an acid lipase secreted into the saliva by glands in the glossoepiglottic or pharyngeal area,
which is highly homologous to the gastric lipases of many other species (Gargouri et al.,
1989). About one-third of fatty acids in milk TG are hydrolyzed from the glycerol backbone
in the abomasum (Edwards-Webb and Thompson, 1978). Activity of the acid lipase has been
Interorgan lipid and fatty acid metabolism 325

thought to be greater toward the short- and medium-chain fatty acids (Edwards-Webb and
Thompson, 1978), although it has been shown recently that the acid lipase has selectivity both
for short-chain fatty acids and for fatty acids on the sn-3 position of milk TG (Villenueve et al.,
1996). Furthermore, specificity of the enzyme in these regards appears to vary inversely with
the same characteristics in mothers’ milk; i.e. enzyme specificity is highest in goat kids, in
which the content of short-chain fatty acids is lower than in cows’ milk, and enzyme speci-
ficity is lower in the calf where cows’ milk is rich in short-chain fatty acids (Jenness, 1985).
The metabolic significance of this preduodenal digestion is not certain but several possibil-
ities exist. Short-chain fatty acids are absorbed into the portal vein and are extensively cleared
by the liver. In this regard they are considered “obligate fuels” for the liver (see Chapter 9 by
Odle et al., this volume). In preruminants consuming mothers’ milk, a firm coagulum is formed
in the abomasum, which slows the flow of casein and long-chain fatty acids from the abomasum.
However, lactose and the whey proteins are expelled from the coagulum, and are available for
intestinal digestion much more quickly after a meal. Perhaps the initial digestion of milk fat,
particularly the hydrolysis of short-chain fatty acids, provides a readily available fuel for the
liver to spare amino acids derived from digestion of whey proteins for protein synthesis, and to
provide the ATP necessary to drive that protein synthesis. Moreover, the initial hydrolysis of
short-chain fatty acids from milk fat may increase emulsification of fat droplets in the intestine
(Armand et al., 1994) and may increase subsequent hydrolysis by pancreatic lipases in the small
intestine (Borel et al., 1994). While the role of gastric or preduodenal lipases in total fat digestion
has been shown to be larger than previously considered (Gooden, 1973), this may be especially
true in the young preruminant in which pancreatic lipase activity is immature. The whey protein
β -lactoglobulin may bind and remove fatty acids from the acid lipase to prevent end-product
inhibition (Perez et al., 1992), although the extent to which this occurs is questionable given
that β -lactoglobulin is not retained in the abomasal coagulum.
While intestinal hydrolysis of the partially hydrolyzed dietary lipids traditionally has been
considered to occur by action of the colipase-dependent pancreatic lipase, recent evidence has
demonstrated that the pancreas of most mammalian species also secretes a bile salt-activated
lipase (Wang and Hartsuck, 1993). This enzyme recently has been purified, characterized, and
sequenced from bovine pancreas (Tanaka et al., 1999). The enzyme exerts considerable cat-
alytic activity toward TG in the presence of bile salts, and has basal activity toward shorter-chain
TG even in the absence of bile salts. The enzyme also is active in catalysis of phospholipids;
indeed, the enzyme activity was first identified in the bovine as a lysophosholipase (van den
Bosch et al., 1993). In addition, the enzyme also hydrolyzes cholesterol esters, phosphatidyl-
choline, and fatty acid esters of vitamins A and E (Wang and Hartsuck, 1993). The latter roles
may be the more physiologically relevant in preruminants; alternatively, the apparent redun-
dancy of having two distinct pancreatic enzymes active toward TG may be yet another
example of ensuring that almost all of the dietary TG presented to the young animal is
hydrolyzed for absorption.
A number of fat sources are used in milk replacers (milk substitutes) fed to young
preruminants. Commonly used sources include tallow, lard, and coconut oil (Davis and
Drackley, 1998). Vegetable oils such as corn oil, soybean oil, and sunflower oil are not widely
used because of early research demonstrating that calves grew poorly and developed diarrhea
(scouring) when fed milk replacers containing vegetable oils (Toullec and Mathieu, 1969;
Jenkins et al., 1985, 1986). Canola oil did not seem to cause scouring (Jenkins et al., 1986).
Subsequent research by the same group suggested that the scouring in earlier studies may
have been a result of improper emulsification so that lipid particle size was too large (Jenkins,
1988). Consequently, the degree to which animal-derived fats could be replaced with more
326 J. K. Drackley

highly unsaturated vegetable sources, if properly emulsified, without compromising calf health
or performance would appear to be an unresolved issue.
Free fatty acids also are less acceptable as lipid sources for young preruminants than TG
of the same fatty acid composition (Jenkins et al., 1985). Although emulsification and droplet
size may have been concerns in that study, free fatty acids clearly inhibit feed intake in young
calves (Spanski et al., 1997) as well as in adult cows (Drackley et al., 1992; Christensen
et al., 1994; Bremmer et al., 1998). This effect is specific for unsaturated rather than saturated
fatty acids (Drackley et al., 1992; Christensen et al., 1994; Bremmer et al., 1998), and is dose-
dependent (Overton et al., 1998; Drackley et al., 2000). Presence of the unsaturated free fatty
acids in the duodenum interacts with receptors that in turn signal through glucagon-like
peptide-1 and perhaps cholecystokinin (Drackley et al., 2000; Benson and Reynolds, 2001)
to decrease feed intake. Because dietary TG are not hydrolyzed until more distal regions of
the jejunum, past the site of greatest density of these neuroendocrine cells, unsaturated TG do
not inhibit feed intake as potently as do the same LCFA provided as free acids (Bremmer
et al., 1998). This explains why decreases of dry matter intake have been more pronounced
with postruminal administration of much smaller quantities of free fatty acids than when
much larger quantities of unsaturated TG were administered postruminally (e.g. Gagliostro
and Chilliard, 1991; Drackley et al., 2000). Digestibility of the free fatty acids is high
(Bremmer et al., 1998) and similar to that of TG in young calves, at least when some TG is
present to furnish 2-monoglycerides in the intestine (Spanski et al., 1997).

2.2. Ruminants

Digestion and absorption of LCFA in ruminants have been addressed in a number of reviews
(Noble, 1978; Moore and Christie, 1984; Bauchart, 1993; Jenkins, 1993; Doreau and Chilliard,
1997) and will be discussed only briefly here. Lipid digestion in ruminants begins in the rumen
(see Jenkins, 1993, for review). A number of ruminal microorganisms are actively lipolytic,
resulting in extensive hydrolysis of most dietary complex lipids. The resulting free fatty acids
undergo varying degrees of biohydrogenation, with production of the saturated stearic acid,
as well as smaller quantities of trans-unsaturated monoenes and dienes that leave the rumen
for absorption. The microbial population also synthesizes a variety of fatty acids, principally
odd-chain and branched-chain acids of 15 to 17 carbons, by elongating shorter-chain fatty
acids (<14 carbons) from the diet. The microbes also synthesize phospholipids.
Lipids in digesta reaching the postruminal tract consist mainly of mostly saturated free
fatty acids adsorbed to the surface of feed particles and bacteria, with the remainder being
predominantly phospholipids and sterol esters as components of microbial cells. Because of
the low pH in the abomasum and duodenum of ruminants (pH 2.0–2.5), the free fatty acids
exist in the protonated state, which facilitates their adsorption to the surface of feed particles.
The strong detergent properties of bile salts secreted in the upper duodenum serve to desorb
LCFA from particulate matter, with formation of a liquid crystalline phase. Subsequent for-
mation of lysophosphatidylcholine (lysolecithin) from phosphatidylcholine (lecithin, from
bile or from acid-mediated disruption of rumen microbial cells) by pancreatic phospholipase A2
promotes formation of micelles. Stable micelles formed from LCFA, lysolecithin, and bile
salts function to move lipid across the unstirred water layer of the small intestinal epithelium,
where absorption of free LCFA and lysolecithin can occur by diffusion.
Stearic acid is the predominant lipid to reach the small intestine. Ruminants are able to
absorb saturated LCFA such as palmitic acid and stearic acid with substantially greater effi-
ciency than nonruminants (Moore and Christie, 1984). One reason for this adaptation is the
Interorgan lipid and fatty acid metabolism 327

reliance on lysolecithin as the major micelle stabilizer. Swelling amphiphiles are substances that
can expand the volume of bile salt micelles and their hydrophobic interior in the aqueous envi-
ronment of the small intestinal lumen (Small, 1968). Of all the naturally occurring swelling
amphiphiles (monoglycerides, medium-chain fatty acids, long-chain unsaturated fatty acids,
phospholipids, and lysophospholipids), lysolecithin is the most efficient at increasing the solu-
bility of stearic acid. For example, lysolecithin increases the partitioning of stearic acid into the
micelle by 115%, compared with only 36% for 1-monolein (Freeman, 1984). A second factor
contributing to greater absorption of long-chain saturated fatty acids in ruminants is the lower
pH in the upper small intestine, ranging from about 3.0 in the duodenum to about 6.0 in the
mid-jejunum. The relatively low pH helps minimize formation of calcium soaps of palmitate
and stearate, which has long been known to decrease absorption of these fatty acids in non-
ruminants (Cheng et al., 1949). In ruminants, the major bile salt is taurocholate rather than
glycocholate; the lower pKa of taurocholate (2.0) is likely an advantage since it is less likely
to become insoluble in the acid conditions of the ruminant small intestine than is glycocholate
(pKa 4.7; Harrison and Leat, 1975).
True digestibility of LCFA generally is quite high in ruminants (Moore and Christie, 1984).
Based on results of Palmquist (1991), true digestibility of LCFA may decline with increasing
LCFA intake. As discussed by Bauchart (1993), this decrease suggests that pancreatic phos-
pholipase activity and bile lipids (phospholipids and bile salts) may become limiting for
absorption of large dietary loads of LCFA.

2.3. Changes during the weaning transition from preruminant to ruminant

During the weaning transition, the young animal changes to a relatively low-fat diet, in which
dietary lipids may constitute only 2–6% of dry matter. These lipids typically consist of galacto-
lipids and phospholipids from forages and TG from cereals and oilseeds (Noble, 1978). During
the weaning transition, the dietary supply of glucose largely ceases as rumen microbial fermen-
tation of dietary carbohydrates is established. The resultant short-chain or volatile fatty acids
(VFA), particularly acetate, become the major oxidative fuel for most tissues of the ruminant.
Acetate also becomes the major precursor for lipogenesis. As the young ruminant begins to con-
sume solid food, the ruminal tissue function, capacity, and microbial activity increases (see
Davis and Drackley, 1998), leading eventually to the pattern discussed for ruminants.
During the transition period when the young animal is receiving both a liquid diet and dry
feed, lipids in the liquid (milk) diet continue to reach the abomasum and small intestine via clo-
sure of the esophageal groove. Few data are available that quantify LCFA absorption during the
transition from preruminant to ruminant. Spanski et al. (1997) found that apparent total tract
digestibility of LCFA by calves fed a control (starter) diet averaged about 82% and did not differ
among measurements at 6, 8, and 10 weeks of age. Digestibility of LCFA when a liquid sup-
plement was provided that contained either lard TG or a mixture of lard TG and free fatty acids
from lard was not appreciably different from LCFA digestibility of the control diet.
An enigma exists for transitioning ruminants, as well as for pigs, in that digestibility and
utilization of LCFA is markedly less after weaning to dry feed than for LCFA use from the
liquid diet before weaning. Indeed, these changes seem to occur as abruptly as the weaning
process itself. Before weaning preruminants consume a diet in which a major portion of
dietary lipid is provided in milk or milk replacer, which may contain from 15% to 30% fat on
a dry solids basis. After weaning, dry starter diets may contain only 3% to 6% dietary fat, yet
digestibility and use are much lower even with the lower fat content. For example, Fallon
et al. (1986) showed that average daily gain (ADG) of body weight decreased as a fat
328 J. K. Drackley

supplement (calcium soaps of palm oil fatty acids) was increased from 0 to 20% of the starter
formulation. What factors suddenly become limiting for use of fat?
A portion of the decreased performance might be attributed to interference by fat with
ruminal fermentation; however, differences persist even when fats that should be relatively
inert in the young rumen are fed (Fallon et al., 1986). Increasing fat content of milk replacer
or starter decreases development of starter intake (Kuehn et al., 1994). In contrast, increasing
fat in a liquid diet (as in veal production) increases ADG, whereas increasing fat in starter
decreases ADG (Doppenberg and Palmquist, 1991). Because of ruminal hydrolysis of dietary
lipids, substantial quantities of free fatty acids reach the upper small intestine when young
ruminants consume starter feeds; the presence of free fatty acids in the duodenum may exert
negative feedback on appetite as described earlier.
Likely of major importance is the nature of the fat in the feed, i.e. a highly emulsified milk
fat or fat in milk replacer vs. fat incorporated into a solid feed matrix. A limiting factor may
be that the compounds needed to emulsify lipid and form micelles in the small intestinal
lumen may be secreted in insufficient amounts in the young ruminant. For example, as dis-
cussed earlier, when TG are no longer being hydrolyzed to 2-monoglycerides in the intestine,
lysophosphatidylcholine becomes the predominant stabilizing compound of mixed micelles
(Freeman, 1984). Whether biliary secretion of phospholipids to provide substrate for
phospholipase-mediated production of lysophospholipids is limiting, or whether phospholi-
pase activity itself might be limiting during this transition, has not been investigated. Gooden
(1973) determined that conversion of lecithin to lysolecithin occurred much more rapidly in
intestinal contents of ruminating calves compared with 1- to 2-week old calves, although
much of the lower rate of conversion in milk-fed calves may have been from the presence of
milk TG. The possibility that hepatic synthesis and secretion of bile salts might be inadequate
to disperse lipids for micelle formation also does not appear to have been investigated in
young ruminants.

3. INTESTINAL METABOLISM AND TRANSPORT


OF ABSORBED DIETARY LIPIDS
General aspects of intestinal absorption of LCFA, re-synthesis of TG, and synthesis of
lipoproteins have been reviewed recently by Phan and Tso (2001). Function of these
processes in preruminants and ruminants also has been reviewed (Noble and Shand, 1982;
Bauchart, 1993; Doreau and Chilliard, 1997). Absorption of LCFA into intestinal epithelial
cells generally has been assumed to occur by simple diffusion into and across the lipid-bilayer
membrane, down the concentration gradient maintained by intracellular binding and metabolism.
Recently, several putative transporter proteins for LCFA have been identified in other species
(Chen et al., 2001). These proteins may facilitate uptake across cell membranes, especially at
low extracellular concentrations, and may be subject to metabolic regulation depending on
the physiological state of the animal. To date, however, no data are available on the presence
and role of such proteins in ruminant tissues.
Cytosolic fatty acid binding proteins (FABP) have been identified in most tissues of
rodents and other species that have been examined (see Glatz and van der Vusse, 1996, for
review). Intracellular FABP activity has been identified in the small intestine of the prerumi-
nant calf (Jenkins, 1986). Binding activity in calf intestinal tissue was associated with a
protein fraction with molecular weight of approximately 12 kD, which is in the range reported
for FABP of other species. Functions of intestinal FABP are not entirely clear, but it is intu-
itive that FABP binds LCFA as they desorb from the plasma membrane, to keep intracellular
Interorgan lipid and fatty acid metabolism 329

concentrations of potentially toxic free fatty acids low. In pigs, FABP is induced in the small
intestine by colostrum feeding after birth (Reinhart et al., 1992). Evidence has been presented
in other tissues and species for possible roles of FABP in directing LCFA to appropriate intra-
cellular sites of metabolism, such as mitochondria or endoplasmic reticulum (see Glatz and
van der Vusse, 1996) or nucleus (Wolfrum et al., 2001).
Absorbed LCFA are activated by acyl-CoA synthetase located on the outer leaflet of the
endoplasmic reticulum (Moore and Christie, 1984). After activation of LCFA to their acyl-CoA
esters, they undergo re-synthesis to TG by one of two pathways (fig. 1). In the preruminant that
absorbs both free LCFA and 2-monoglycerides, fatty acid esterification proceeds by the
monoglyceride pathway, in which the remaining two positions on the glyceride molecule are
esterified with acyl-CoA to form TG. In functioning ruminants, no 2-monoglyceride is
absorbed; consequently, TG synthesis in the intestine proceeds by the phosphatidate
(Kennedy) pathway from glycerol-3-phosphate. While few data are available for ruminants,
these pathways are not believed to be major sites for metabolic regulation. Rather, the path-
ways function to repackage the amount of LCFA presented to the intestine into TG for
distribution throughout the body. Teleologically, one can argue that it is prudent for the animal
to efficiently take up all LCFA available from the diet at any time, given the possibility that
dietary energy sources might not be available at some point in the future. Thus, regulation
occurs at disposition and storage sites within the body rather than at the site of assimilation.
Triacylglycerol synthesis occurs in concert with synthesis of the TG-rich lipoproteins, chylo-
microns and very low-density lipoproteins (VLDL). The general scheme for synthesis and
secretion of these lipoproteins has been well described (Tso and Balint, 1986; Davidson and
Shelness, 2000; Phan and Tso, 2001), although specific details in ruminants are lacking. The
primary structural apolipoprotein (apoprotein) of intestinal TG-rich lipoproteins is apoprotein-
(apo)-B48, whereas the liver synthesizes primarily apo-B100. Ruminants synthesize apo-B

Fig. 1. Major pathways of esterification of fatty acids to glycerolipids in growing ruminants. Key enzymes
involved: (1) glycerophosphate acyltransferase (GPAT), which may be mitochondrial or microsomal;
(2) lysophosphatidate acyltransferase; (3) phosphatidate phosphohydrolase; (4) diacylglycerol acyltransferase;
and (5) monoacylglycerol acyltransferase. Reactions 1 to 4 constitute the phosphatidate (Kennedy) pathway
and are found in adipose tissue, liver, and muscle. Activity of reaction 5 (monoacylglycerol acyltransferase)
is largely confined to intestinal enterocytes during digestion of milk triglycerides in preruminants. Pi, inorganic
phosphate. Adapted from Drackley (2000) based on Rule (1995).
330 J. K. Drackley

proteins similar to those of other species. Chylomicrons and VLDL found in intestinal lymph
of ruminants contain only the protein analogous to apo-B48, whereas the VLDL found in
plasma contain both apo-B100- and apo-B48-like proteins (Laplaud et al., 1990, 1991). Both
forms of the protein are translated from the mRNA produced by a single gene. The shorter
form (apo-B48) is produced as a result of a unique post-transcriptional mRNA editing mech-
anism in intestinal cells (Davidson and Shelness, 2000). The apo-B mRNA of bovine intestine
is almost completely edited to the shorter form (95%), whereas less editing (40%) occurs in
sheep intestine (Greeve et al., 1993). Apo-B synthesized and secreted by ruminant liver in the
form of VLDL contains only the apo-B100-like protein (Laplaud et al., 1991).
Whether ruminants synthesize and secrete chylomicrons or just VLDL has been the sub-
ject of considerable debate, but clear evidence for the existence of chylomicrons is available
in preruminants (Laplaud et al., 1990; Bennis et al., 1992). In functioning ruminants fed typ-
ical diets, even those supplemented with fats, it appears that the intestinal cells mainly secrete
VLDL (Bauchart, 1993). However, if polyunsaturated LCFA are infused postruminally in
large amounts, more of the absorbed LCFA will be transported as TG in chylomicrons rather
than VLDL (Harrison et al., 1974). Of interest also is the observation that ruminants, espe-
cially calves fed high-fat milk diets, appear to secrete considerable amounts of VLDL into the
portal vein rather than into the lymphatics (Bauchart et al., 1989; Laplaud et al., 1990).
The ontogeny of ruminant lipoproteins has been characterized by several researchers. No
VLDL were detected in plasma of fetal calves near term (Forte et al., 1981). Both low-density
lipoproteins (LDL) and high-density lipoproteins (HDL) were present in plasma of fetal
calves near term, with LDL being the more abundant lipoprotein class (Forte et al., 1981).
Marcos et al. (1991) found that plasma TG decreased steadily from about day 115 to day 265
of gestation in cattle; TG concentrations near term were very low, while apo-B concentrations
had decreased less, suggesting that VLDL decreased from mid- to late gestation but that LDL
remained unchanged or decreased less. Declining concentrations of LDL were also reported
during gestation in fetal lambs (Turley et al., 1996) so that HDL became the major lipopro-
tein fraction at term (Noble and Shand, 1983). After birth, concentrations of VLDL increase
with consumption of colostrum and milk, while LDL concentrations decrease in calves
(Jenkins et al., 1988; Marcos et al., 1991). During the suckling period, concentrations of LDL
appear to increase in sheep (Turley et al., 1996) and goats (Bennis et al., 1992). With the onset
of suckling, HDL increases rapidly and become the predominant lipoprotein class in calves,
lambs, and kids (Forte et al., 1981; Noble and Shand, 1983; Jenkins et al., 1988; Marcos et al.,
1991; Bennis et al., 1992).

4. TISSUE UTILIZATION OF CIRCULATING TRIGLYCERIDES


Triglycerides in circulating chylomicrons and VLDL are hydrolyzed by the enzyme lipopro-
tein lipase (LPL) found in peripheral tissues. Products of the LPL reaction are free fatty acids
(i.e. NEFA) and monoglycerides, which are hydrolyzed by nonspecific lipases associated
with peripheral tissues. Activity of LPL creates a locally increased NEFA concentration that
increases the likelihood for NEFA uptake by cells, although not all NEFA are taken up by the
tissue and NEFA concentration may increase in venous blood draining the tissue bed. The
unusual LPL enzyme is a glycosylated protein produced by tissue parenchymal cells (e.g.
adipocytes), which is secreted from those cells and becomes active when anchored via
heparin sulfate proteoglycans on the inner surface of capillary endothelial cells (see
Olivecrona and Olivecrona, 1999, for review). In growing ruminants LPL is found predomi-
nantly in adipose tissue, skeletal muscle, and heart. Very low activities of LPL were found in
Interorgan lipid and fatty acid metabolism 331

small intestine, kidney, spleen, adrenals, ovary, and lung in bovines; the LPL mRNA also was
barely detectable in these tissues and was undetectable in brain (Hocquette et al., 1998a).
Only trace amounts of LPL activity were found in bovine liver, and the absence of detectable
mRNA by Northern analysis indicates that this probably represented LPL from blood retained
in the tissue (Hocquette et al., 1998a). Activity and mRNA abundance were higher in heart
and masseter muscle than in other oxidative and glycolytic muscles of calves, but LPL activ-
ity was lower for bovine muscles in general than for rat muscles in parallel experiments
(Hocquette et al., 1998a). Internal adipose depots (perirenal and omental) had greater LPL
activity and mRNA abundance than subcutaneous adipose tissue from fattened calves (Hocquette
et al., 1998a). Activity of LPL in perirenal adipose tissue was similar in cows and ewes, and LPL
is transcribed from similar 3.4 and 3.8 kb mRNA in both species (Bonnet et al., 1998).
Extensive research has been conducted on LPL in laboratory animals, farm animals, and
humans over the last several decades. Despite this intense scrutiny, details of the regulation
of LPL still are not resolved (Olivecrona and Olivecrona, 1999). Because LPL is synthesized
by parenchymal cells of the tissue and then transported across the capillary endothelial cells,
where the enzyme is anchored on the capillary luminal surface to be active, its synthesis and
regulation are understandably complex. Dogma developed on the basis of studies in rats
assumes that LPL in adipose tissue is upregulated during positive energy balance, decreases
during fasting, and increases on refeeding after a fast, whereas in heart and skeletal muscle
the enzyme activity changes less but in opposite direction to that in adipose, i.e. increasing
during fasting and decreasing with refeeding (e.g. Sugden et al., 1993; Cortright et al., 1997).
In this way, heart and skeletal muscles would receive metabolic priority for use of TG circu-
lating in VLDL derived from liver repackaging of NEFA mobilized from adipose tissue.
In both ewes and cows, underfeeding (20% of maintenance requirements) results in decreased
LPL activity and mRNA abundance in perirenal adipose tissue. Contrary to the dogma from lab-
oratory animals, however, LPL activity in heart and skeletal muscle also decreases during
underfeeding, and increases with refeeding (Bonnet et al., 2000; Faulconnier et al., 2001). Similar
changes occur in pigs (Enser, 1973). Bonnet et al. (2000) proposed a linear relationship between
the change in LPL activity in cardiac muscle and the ability of liver slices to secrete TG among
six species (pig, sheep, guinea pig, rabbit, rat, and chicken), which makes teleological sense in
that circulating TG concentrations are maintained during fasting or feed restriction in species that
actively secrete VLDL from liver (e.g. chicken, rat, rabbit) but fall in species that do not (e.g. pig,
sheep). In ruminants, therefore, during feed restriction the energy needs of heart and skeletal
muscle are met more via NEFA and ketone bodies and less from TG.
In contrast to the results for ruminants subjected to underfeeding, altering the plane of nutri-
tion above maintenance does not produce the same changes in LPL among tissues. Andersen
et al. (1996) fed groups of ewe lambs a diet in amounts to support daily gains of 0.15 or 0.25 kg.
Adipose tissue LPL activity was greatest for lambs fed at the highest plane of nutrition,
whereas skeletal muscle LPL was highest for lambs grown at the slower rate. For cardiac
muscle LPL, rates were greatest for lambs grown at the slowest rate and for a group of lambs
fed for compensatory gain (0.33 kg/d) after a period of growth at rates equal to the slowest rate
of gain. These data suggest that the concept of reciprocal regulation between adipose and
muscle LPL is valid when comparing differences in energy balance above maintenance, but
not in more extreme situations where animals are in negative energy balance.
The molecular basis for the differences in response of LPL among tissues within species,
and across species, is not known. The cDNA for bovine LPL has been cloned (Senda et al.,
1987) and the mRNA has been characterized in both sheep and cattle (Hocquette et al., 1998a;
Bonnet et al., 1998). In sheep, the 3.4 kb mRNA was predominantly expressed in adipose
332 J. K. Drackley

tissue, whereas the 3.8 kb mRNA was predominant in cardiac muscle, with no differential
regulation according to nutritional state (Bonnet et al., 2000). Large differences in LPL
expression among different muscles (e.g. heart and masseter vs. rectus abdominis) were par-
alleled by changes in mRNA abundance, suggesting that transcriptional regulation is a major
factor in differences among tissues.
Activity of LPL and the presence of its mRNA were detected in heart, masseter muscle, and
perirenal adipose tissue of bovine fetuses (230–260 days in gestation), but activities and
mRNA were lower than in growing calves (Hocquette et al., 1998a). Changes in LPL activity
and mRNA in adipose tissues and muscles over the transition from preruminant to ruminant
were quantified by Hocquette et al. (2001). Activity of LPL was 2-fold lower in adipose tissue
from weaned calves than from milk-fed calves; mRNA abundance was correspondingly lower
for weaned calves as well. In contrast, neither LPL activity nor mRNA abundance were dif-
ferent between weaned and milk-fed calves in any of seven skeletal muscles studied, with the
notable exception of the masseter muscle. Both LPL activity and mRNA abundance were more
than doubled in masseter of weaned calves. Contraction of the masseter, located in the cheek,
would increase greatly in weaned calves with the increased frequency and workload of chew-
ing; consequently, LCFA use as oxidative fuel might be expected. Activity, but not mRNA
abundance, of LPL tended to decrease in heart from weaned calves.
Together, the available data implicate a role for LPL in controlling use of TG-FA by grow-
ing ruminants. Indeed, for well-fed growing ruminants, Pethick and Dunshea (1993)
calculated that most of the NEFA flux was derived from hydrolysis of lipoprotein TG by LPL.
The potential role of LPL in determining fat deposition, especially for intramuscular fat
(marbling) in fattening ruminants, is still an unanswered question.

5. LIPOPROTEIN METABOLISM
Lipoprotein metabolism in ruminants differs significantly from the more-studied laboratory
species such as rats, mice, and guinea pigs. An excellent comprehensive review (Bauchart,
1993) is available to which the interested reader is referred and in which can be found refer-
ences to other, older reviews. A schematic is presented in fig. 2 to describe the general
patterns of lipoprotein metabolism in ruminants.
The TG-rich lipoproteins (chylomicrons and VLDL) function to deliver LCFA absorbed
from the intestine to peripheral tissues. As discussed earlier (see section 3), intestinally syn-
thesized VLDL predominate in ruminants, whereas chylomicrons would be more important
in the milk-fed preruminant. Following secretion into the lymphatics and entry into the
venous blood system, chylomicrons and intestinal VLDL acquire additional surface-coat
compounds, including phospholipids and apo-C proteins, from circulating HDL. Apo-CI and
apo-CII are important for lipid binding and activation of LPL activity, respectively.
Chylomicrons and VLDL thus activated by apo-CI interact with LPL in the capillaries of
peripheral tissue to catalyze “unloading” of TG in those tissues. Following LPL action, chy-
lomicrons and VLDL are converted to chylomicron remnants and intermediate-density
lipoproteins (IDL), respectively. Chylomicron remnants and IDL containing apo-B48, i.e.
those of intestinal origin, are found in extremely low concentrations in ruminants, suggesting
that they may be more rapidly cleared by the liver. As hydrolysis of TG occurs, the particle
size decreases and thus surface components (phospholipid, apo-A and apo-C) become exces-
sive and are transferred back to HDL.
Most LDL are formed in circulation from IDL in the ruminant. The LDL make up a
small proportion of total lipoproteins (about 10%) in adult ruminants (Bauchart, 1993).
Interorgan lipid and fatty acid metabolism 333

Fig. 2. Schematic diagram showing major aspects of lipoprotein metabolism in growing ruminants.
Chylomicrons (mainly in preruminants) or very low-density lipoproteins (VLDL) secreted from the intestine
or liver acquire apoproteins- (apo)-CI and apo-CII from circulating high-density lipoproteins (HDL).
Triacylglycerols in chylomicrons or VLDL are hydrolyzed by lipoprotein lipase (LPL) in peripheral
tissues, which is activated by apo-CII and allows fatty acid uptake by tissues. Excess surface components
(phospholipids, PL; apoproteins C and A; free cholesterol, chol) that arise as chylomicrons and VLDL
decrease in size during LPL catalysis of triacylglycerols are transferred to HDL. The chylomicron
remnants are cleared by the liver. Remnants of VLDL are called intermediate density lipoproteins (IDL) and
are either cleared by the liver or undergo further triacylglycerol hydrolysis to produce low-density lipoproteins
(LDL). LDL are degraded in liver or after receptor-mediated uptake in peripheral tissues. HDL take up excess
cholesterol (chol) from peripheral tissues and convert it to cholesterol esters by action of lecithin cholesterol
acyltransferase (LCAT); lysolecithin is released into plasma and cholesterol esters enter the core of the HDL.
Action of LCAT to produce cholesterol esters, and the uptake of excess surface components from chylomi-
crons and VLDL, results in increasing size and decreasing buoyant density of lipid-poor heavy HDL (HDLH)
to lipid-rich light HDL (HDLL). HDL can deliver cholesterol and essential fatty acids to tissues or return
cholesterol to liver for conversion to bile salts. Many tissues also possess an HDL receptor that results in
clearance of HDL particles. Adapted from Drackley (2000).

Low concentrations have been attributed to the low activities of cholesterol ester transfer
protein (Ha and Barter, 1982) and hepatic lipase in ruminants (Cordle et al., 1986), which are
involved in the production of LDL in other species. Plasma concentrations of LDL in rumi-
nants are also controlled by expression and activity of tissue LDL receptors. The major sites
for receptor-mediated removal of LDL are bone, intestine, and liver (Rudling and Peterson,
1985). High concentrations of LDL receptors are also found in bovine adrenals and corpora
lutea (Rudling and Peterson, 1985), which indicates that LDL may be important sources of
cholesterol for steroid hormone biosynthesis in these tissues. Because LDL are rich in cho-
lesterol esters and phospholipids containing linoleic acid, tissue uptake of LDL also served to
distribute this EFA for membrane formation.
Bovine HDL are the main lipoprotein class in ruminants, constituting over 80% of total plasma
lipoproteins in both preruminants and ruminants (Bauchart, 1993). The synthesis of HDL
remains an enigma even in more well-studied species (Fielding and Fielding, 2001), but precur-
sor particles (apo-AI) are synthesized by liver and small intestine (Fielding and Fielding, 1995).
334 J. K. Drackley

Genesis of pre-HDL particles appears to occur in circulation, either from chylomicrons and
VLDL or within the interstitial spaces (Fielding and Fielding, 1995). Initial complexes of
apo-AI and phospholipids (pre-β-HDL) acquire cholesterol from tissue membranes by as yet
undefined mechanisms that may involve some combination of simple diffusion, facilitated dif-
fusion, active transport, or other mechanisms (see Fielding and Fielding, 2001, for review). The
lecithin-cholesterol acyltransferase (LCAT) reaction then functions to catalyze transfer of an
unsaturated fatty acid (primarily linoleic) from the 2-position of phosphatidylcholine (lecithin)
to cholesterol, both polar surface-coat lipids of HDL. The more nonpolar cholesterol ester
migrates to the core of the HDL, while the lysolecithin is shed to serum albumin. Continued
transfer of cholesterol esters to the core of the HDL causes their expansion in size, forming
“light” HDL. The lack of cholesterol ester transfer protein, which would transfer cholesterol
esters to LDL or VLDL, in ruminants and the high turnover of surface-coat lipids from VLDL
result in the large size of HDL (Bauchart, 1993). The presence of the very large and less dense
(i.e. high lipid content) HDL in ruminants results in overlapping density ranges for LDL and
HDL, thus making separation of LDL and HDL by density gradient ultracentrifugation proce-
dures impossible in ruminants. Alternative approaches have been developed that rely on
combinations of affinity chromatography and centrifugal separations (see Bauchart, 1993 for
review).
The HDL of ruminants possesses apo-AIV, which is secreted by intestinal cells in TG-rich
lipoproteins and then transferred to HDL in lymph and plasma (Bauchart et al., 1989).
Originally described as an activator of peripheral LPL, apo-AIV has been implicated in recent
years as a regulator of food intake in humans and laboratory animals (see Tso et al., 2001, for
review). An intestinal fat load stimulates synthesis and secretion of apo-AIV, which in turn
acts to suppress additional feed intake. The potential roles of apo-AIV in regulation of feed
intake in ruminants, or its response to supplemental fat, have not been investigated.
Because of their predominance in ruminants, HDL function as the main distribution vehi-
cle for cholesterol and EFA. Specific receptors for HDL have been identified and
characterized (Graham and Oram, 1987). This receptor recognizes apo-AI but not LDL or
apo-E; the apo-E protein is not expressed to any extent in ruminants (Bauchart, 1993). Uptake
of HDL by bovine liver, and probably other tissues, is regulated by the density of the HDL
receptors, which are almost always occupied with HDL, rather than by concentration of HDL
in plasma (Mendel et al., 1986). Cholesterol uptake by tissues is used for membrane synthe-
sis, steroidogenesis, or, in the liver, bile salt synthesis. Uptake of cholesterol from peripheral
tissues by HDL followed by clearance of HDL by the liver constitutes a cycle that has been
referred to as “reverse cholesterol transport” in other species and is one component of the
regulation of cholesterol homeostasis in the body (Fielding and Fielding, 1995).
Relatively little is known about cholesterol synthesis and homeostasis in ruminants. The
major site of cholesterol synthesis appears to be the small intestine (Nestel et al., 1978), which
is in keeping with the low rate of production of VLDL by the ruminant liver (Kleppe et al.,
1988; Graulet et al., 1998; Gruffat-Mouty et al., 1999), although the ruminant liver does syn-
thesize some cholesterol. Transport of cholesterol from intestine and liver to other tissues by
LDL and HDL furnishes the needs of those tissues for membrane structure and steroid
hormone biosynthesis.
Although beyond the scope of this chapter, metabolism of the EFA, i.e. those LCFA pro-
duced from elongation and desaturation of linoleic acid (18:2n-6) and linolenic acid
(18 : 3n-3), is a particularly fascinating topic in ruminants (Noble, 1984). Little transfer of
LCFA occurs across the ruminant placenta, so the fetuses and young are born with what
would be considered severely deficient status for EFA in other mammalian species (see Noble
Interorgan lipid and fatty acid metabolism 335

and Shand, 1982, for review). Linoleic and linolenic acids in plasma phospholipids and cho-
lesterol esters are extremely low at birth, but then increase substantially by 3 weeks with little
additional change after weaning (Jenkins et al., 1988). After birth, the young receive some
EFA from colostrum and milk, but again intakes are insufficient relative to other species
because of the low EFA content in ruminant milk. Because functional deficiencies of EFA do
not seem to occur in preruminants or ruminants, the ruminant animal must have developed
extremely efficient mechanisms for capturing and maintaining EFA within the body.
Mechanisms for the conservation and concentration of EFA include lower rates of oxida-
tion compared with more abundant saturated LCFA (Lindsay and Leat, 1977), which may be
mediated via the low affinity of mitochondrial dehydrogenases for EFA (Reid and Husbands,
1985). Other mechanisms include the high affinity of phospholipid esterification enzymes for
incorporation of EFA (Lindsay and Leat, 1977), the high affinity of lecithin cholesterol acyl-
transferase for linoleic acid (Noble et al., 1972), and the slow turnover of the phospholipid
and cholesterol ester pools in plasma (Palmquist, 1976).
Interorgan transport and metabolic conversions of the lipoproteins serve as a vehicle for dis-
tribution of EFA to cells throughout the body for incorporation into cell membranes. Other
methods by which tissues acquire EFA include uptake of EFA as free fatty acids, uptake of the
EFA-acyl group from lysophosphatidylcholine, and local desaturation and elongation of free
linoleic acid (see review by Zhou and Nilsson, 2001). The extent to which peripheral tissues
of ruminants acquire EFA by uptake or transfer of phospholipids from lipoproteins to cell
membranes vs. receptor-mediated uptake of LDL or HDL is not well characterized. Zhou et al.
(2002) demonstrated that plasma free arachidonic acid was the major source of arachidonic
acid taken up by extrahepatic tissues in rats. Regardless, increasing dietary intake of fats and
oils rich in n-6 or n-3 LCFA results in corresponding increases in contents of these LCFA in
phosphatidylcholine and cholesterol esters in plasma of preruminants, and corresponding
increases in the n-6 and n-3 LCFA in phophatidylcholine and phosphatidylethanolamine in
liver and muscle (Jenkins and Kramer, 1990). Similar changes occur in functioning ruminants
if dietary sources of EFA are protected from ruminal biohydrogenation (Ashes et al., 1995). It
appears that phosphatidylethanolamine in muscle, primarily found in the inner leaflet of
plasma membranes, may be particularly important for concentration of n-3 LCFA (Jenkins and
Kramer, 1990; Ashes et al., 1995).
While a considerable body of research has appeared in the last decade describing inter-
organ lipoprotein metabolism in growing ruminants, these studies mostly were conducted on
calves that were maintained in the preruminant state on milk diets long after what would be
conventional in North American production systems. Research on lipoprotein metabolism in
growing ruminants is lacking. The significance and role in growth, therefore, is still largely
as speculative as it was over twenty years ago (Kris-Etherton and Etherton, 1982).

6. TISSUE FATTY ACID METABOLISM IN GROWING RUMINANTS


Excellent comprehensive reviews of lipid metabolism in adipose tissue (Vernon, 1980; Chilliard,
1993) and liver, muscle, and other tissues (Bell, 1980) are available. Discussion here will mainly
consider recent research on lipid metabolism in key organs relative to growth of ruminants.

6.1. Skeletal muscle and heart

During growth of ruminants, muscle protein accretion is afforded a higher metabolic priority
than is lipid deposition. For example, Smith et al. (1992) fed groups of ovariectomized
336 J. K. Drackley

Angus–Hereford heifers different amounts of the same diet to supply 0.76, 1.43, 1.74, and
2.05 times the estimated requirements for metabolizable energy for a 128-day feeding period.
Retention of dietary N increased linearly with increasing intake, whereas measurements of
adipose lipid synthesis increased only for groups fed 1.74 and 2.05 times maintenance. Rates of
protein deposition in muscle (and whole carcass) decrease as the animal approaches maturity
(Campbell, 1988).
Skeletal muscles contain differing proportions of fibers that are primarily oxidative (red
muscle), primarily glycolytic (white muscle), or a mixture of both. Metabolic properties may
vary even within the same muscle (Brandstetter et al., 1997). So-called red muscle is character-
ized by a higher density of mitochondria than is found in white muscle. Furthermore, at least
two distinct subpopulations of mitochondria exist in ruminant muscle (Piot et al., 2000). The
subsarcolemmal mitochondria are located just beneath the muscle membrane or sarcolemma,
whereas the intermyofibrillar mitochondria are found inserted into the myofibrillar matrix.
Piot et al. (2000) found that enzymes of oxidative metabolism and LCFA oxidation were pres-
ent in greater specific activity in intermyofibrillar mitochondria than in subsarcolemmal
mitochondria. Capacity of heart tissue for LCFA oxidation was much greater than that of
skeletal muscles (rectus abdominis and longissimus thoracis) from preruminant calves, due
to higher density of both intermyofibrillar and subsarcolemmal mitochondria and to greater
specific activities of oxidative enzymes within the mitochondria from heart.
In fed, growing ruminants, glucose, acetate, and β-hydroxybutyrate (BHBA) are the major
fuels for heart and skeletal muscle, potentially accounting for 31–57%, 15–29%, and 18%,
respectively, of fuel for muscle oxidation at rest (Hocquette et al., 1998b; Hocquette and
Bauchart, 1999). In contrast, oxidation of NEFA at rest accounts for only about 5% of oxida-
tive needs. Uptake of NEFA is linearly related to arterial concentration up to about 1 mM
(Bell and Thompson, 1979). Muscle uptake of NEFA likely is facilitated by muscle-type
FABP (Moore et al., 1993). Uptake exceeds the amount oxidized immediately for energy,
with the excess stored in an intracellular TG pool (Bell and Thompson, 1979). Although not
investigated in ruminants, hormone-sensitive lipase is present in muscle from other species
and responds to catecholamine stimulation to initiate intracellular hydrolysis of stored TG for
oxidation within the muscle (Langfort et al., 1998).
Rate-limiting control of NEFA oxidation lies at the enzyme carnitine palmitoyltransferase-1
(CPT-1), which governs entrance of fatty acyl-CoA into mitochondria. Malonyl-CoA is pro-
duced by the muscle form of acetyl-CoA carboxylase and inhibits muscle CPT-1 (Winder,
2001). While not yet characterized in ruminants, in laboratory rodents and humans malonyl-
CoA is produced by the muscle form of acetyl-CoA carboxylase (Chien et al., 2000) and
degraded by malonyl-CoA decarboxylase (Young et al., 2001). Regulation of malonyl-CoA
concentration in muscle thereby represents an elegant control system that coordinates utiliza-
tion of glucose and LCFA depending on substrate availability and muscle energy needs (see
Winder, 2001, for review).
Oxidative capacity per unit of tissue is lower in muscle from cattle than in muscle from rats
(Ottemann-Abbamonte et al., 1998; Piot et al., 1998). Oxidation of palmitate by isolated
strips of muscle in vitro appeared to be saturated at a palmitate concentration of 0.5 mM for
muscle from adult cows, but was not yet saturated at 2.0 mM for rat muscle (Ottemann-
Abbamonte et al., 1998). Piot et al. (1998) attributed the greater oxidative capacity of rat
muscle to a greater mitochondrial density; peroxisomal oxidative capacity was not different
between rats and preruminant calves but was lower for 15-month-old growing bulls. In that
same study, total oxidation capacity of muscle homogenates was about 1.7-fold greater for
preruminant calves than for growing bulls. Piot et al. (1998) proposed that the difference
Interorgan lipid and fatty acid metabolism 337

between preruminant and older ruminant muscle was related both to decreased relative quan-
tity of mitochondria as well as to changes in the oxidative properties of muscle mitochondria.
Bartelds et al. (1998) studied the in vivo flux of energy supplying nutrients to the heart in
fetal, newborn (1 to 4 days), and 7-week-old lambs. In fetal lambs, NEFA contributed no energy
to the heart. In newborn lambs, the supply of NEFA to the heart increased 10-fold, but heart still
did not oxidize NEFA for energy. Glucose was the major energy source in both fetal and new-
born lambs, accounting for 89% and 69% of oxygen consumption, respectively. By 7 weeks, the
flux of NEFA through heart was increased 3-fold, and the supply and use of ketone bodies like-
wise was increased. Similar developmental data for skeletal muscle are lacking.

6.2. Adipose tissue

Adipose tissue represents a complex assortment of anatomical depots, including internal


(e.g. perirenal, omental), intermuscular, subcutaneous, and intramuscular. In growing rumi-
nants, adipose tissue functions to accrete excess energy in the form of TG (see Chapter 10 by
Mersmann and Smith, this volume). The LCFA that are incorporated into adipose tissue TG
during growth and fattening can derive from de novo lipogenesis within the adipose tissue or
via uptake from blood. From a biological perspective, adipose tissue functions as a reserve of
energy for situations of dietary shortage. Recent research has demonstrated that adipose is not
an inert storage vessel, however, but also communicates with other organs and systems of the
body by synthesizing and secreting a variety of mediators such as leptin, tumor-necrosis
factor α (TNFα), and adipsin (see Vernon et al., 1999; Chilliard et al., 2000). Given its impor-
tance in determining eating quality of ruminant meat, and in the production economics of
ruminant agriculture, it is not surprising that lipid metabolism in adipose tissue has been
widely studied.
Lipogenesis is low during the milk-feeding period (Pothoven et al., 1975), then increases
during growth as rates of lean tissue growth decline (Pothoven and Beitz, 1973; Smith et al.,
1984, 1987). Internal adipose tissues have greater rates of de novo lipogenesis in younger ani-
mals, whereas in more mature animals subcutaneous depots have the greater activity (Ingle
et al., 1972; Pothoven and Beitz, 1973; Hansen et al., 1995). Increasing feed intake increases
lipogenesis in subcutaneous adipose (Mills et al., 1989; Smith et al., 1992); conversely, feed
restriction during growth decreases lipogenic rates (Pothoven et al., 1975). The intermediary
metabolism of lipogenesis in adipose tissue appears to differ between sheep and cattle in sev-
eral ways (Smith and Prior, 1986); regulation in sheep and cattle adipose tissue is discussed
in detail by Smith (1995).
Differential activity of lipogenesis among adipose depots and with stage of growth may
relate to adipocyte size. Hood and Allen (1978) found that lipogenic enzyme activities were
positively correlated with the cell volume of adipocytes across several anatomical sites in
sheep. More recently, Barber et al. (2000) showed that expression of mRNA for LPL and
acetyl-CoA carboxylase-α (ACCα), the rate-limiting and regulated step in LCFA synthesis,
per 106 cells was highly correlated with the size of adipocytes isolated from seven subcuta-
neous and internal depots from wethers at slaughter. These data agree with other observations
that lipogenesis is lower in adipose tissue with smaller adipose cell diameters, such as intra-
muscular adipose or in adipose tissue from young animals (Smith, 1995). However, factors
other than adipocyte size also must impact metabolic activity among different adipose tissue
depots. For example, in fattened sheep lipogenesis from acetate was about 10-fold lower in
perirenal adipose than in subcutaneous adipose tissue, but mean cell size and numbers of cells
per gram of adipose tissue differed by only 10% and 24%, respectively (Hansen et al., 1995).
338 J. K. Drackley

Classic experiments by Hanson and Ballard (1967) demonstrated that acetate and not glu-
cose was the major carbon source for LCFA synthesis in adipose tissues of ruminants. A
possible exception is intramuscular adipose depots in which glucose may account for 50–75%
of LCFA carbon (Smith and Crouse, 1984). However, overall rates of de novo LCFA synthe-
sis are low in intramuscular adipose, with rates of incorporation of acetate into LCFA only
10–50% of those in subcutaneous depots (Smith and Crouse, 1984; Miller et al., 1991).
Although rates of esterification of LCFA to glycerol-3-phosphate are lower in intramuscular
adipose than in subcutaneous, differences in LCFA synthesis are much more pronounced
(Smith et al., 1998). This suggests that uptake of preformed LCFA from blood TG by action
of LPL may be more important for lipid deposition in intramuscular adipocytes.
Uptake of preformed LCFA into adipose tissue can be quite substantial in preruminants
consuming high-fat milk-based diets. For example, it can be calculated from data of Tikofsy
et al. (2001), in which preruminant calves were isocalorically fed milk replacers containing
14.8%, 21.6%, or 30.6% fat, that about 45% of the additional fat intake was deposited in the
body. The amount of LCFA originating from the diet in functioning ruminants is much more
limited but not insignificant. For example, typical forage- and grain-based diets for growing
ruminants contain only 3–4% LCFA, which would supply about 6–10% of the digestible
energy intake. Even with near-maximal incorporation of supplemental fat into diets for rumi-
nants, LCFA will usually supply less than 20% of the energy. However, the amount of LCFA
supplied may still contribute a substantial portion of fat deposited in adipose tissues.
An example of the estimated contribution of dietary LCFA supply to adipose fat deposition
is shown in table 1, based on data from Zinn (1992). In this example, cattle fed basal diets
based on corn or wheat had digestible LCFA intakes of about 144 g/d, whereas cattle fed
corn- or wheat-based diets supplemented with 6% yellow grease would have digestible LCFA

Table 1
Estimation of contribution of dietary fat to adipose lipid deposition and suppression of
de novo fatty acid synthesis in fattening beef cattle a

Diet

Basal Basal plus


Variable (no supplemental fat) 6% yellow grease

Dry matter (DM) intake, kg/d 7.82 7.42


Dietary fatty acids (FA)b, % of DM 2.45 7.45
FA intake, g/d 192 553
Estimated digestible FA intakec, g/d 144 415
Estimated FA deposited in adiposed, g/d 118 340
Measured fat gain, g/d 480 550
Estimated FA gaine, g/d 432 495
Dietary FA as percentage of FA gain 27 69
Estimated de novo FA synthesis f, g/d 314 155
% Suppression of de novo synthesis – 51
a Data for intake, diet composition, and fat gain are means from Zinn (1992).
b Reported dietary ether extract concentration multiplied by assumed 85% fatty acid content.
c Assumes dietary fatty acids are 75% digestible (Zinn, 1989).
d Digestible energy of fat equals metabolizable energy; assumes that efficiency of use of metabolizable energy for

fat deposition (i.e. NEg) is 82% (Zinn, 1989).


e Assumes adipose tissue triacylglycerol is 90% fatty acids.
f
Estimated fatty acid gain minus dietary fatty acids deposited.
Interorgan lipid and fatty acid metabolism 339

intakes of about 415 g/d. Assuming that absorbed LCFA are used in fattening cattle with an
efficiency of 82% (Zinn, 1989), cattle fed basal and fat-supplemented diets would deposit 118
and 340 g/d of dietary LCFA. Whole-body LCFA gains averaged 432 and 495 g/d, respec-
tively, which indicates that 27% and 69% of the total LCFA deposited originated from dietary
LCFA for control and fat-supplemented diets, respectively. By difference, whole-body
de novo synthesis of LCFA would need to be decreased by about 51% when the fat-supplemented
diet was fed. Page et al. (1997) found that de novo synthesis of LCFA from acetate was decreased
by 29% in subcutaneous adipose tissue from steers fed a diet containing 30% whole cottonseed
(which supplied an additional 6% lipid to the diet) compared with adipose tissue from control
steers, suggesting that decreased lipogenesis of the magnitude estimated here (table 1) is not
biologically unrealistic.
Other research suggests that dietary lipid supplements also may decrease lipogenesis in
sheep. Hood et al. (1980) found that rumen-protected safflower oil, but not rumen-protected
tallow or palm oil, decreased lipogenesis from acetate by 43% in perirenal adipose tissue. In
subcutaneous adipose tissue, protected safflower oil decreased in vitro lipogenesis by 75%;
differences due to protected tallow (−42%) and protected palm oil (−32%) were not statisti-
cally significant (Hood et al., 1980). Lipogenesis, measured both in vitro and in vivo, was
decreased in sheep fed calcium salts of palm oil (Moibi et al., 2000a). Decreased lipogenesis
by dietary lipid was accompanied by a 30% decrease in fatty acid synthase activity in sub-
cutaneous, mesenteric, and perirenal adipose tissues; in contrast, activity of acetyl-CoA
carboxylase actually was increased by supplemental fat (Moibi et al., 2000b). On the other
hand, lipogenesis in isolated adipocytes was not different between control sheep and those fed
a diet containing 5% soybean oil (Jenkins et al., 1994). Although the data are not conclusive,
together these observations suggest that increased dietary LCFA, even if not protected from
ruminal biohydrogenation, may suppress de novo lipogenesis from acetate in ruminant adi-
pose tissue. This differs from the situation in rodents, in which only unsaturated LCFA have
been shown to be effective at suppressing lipogenesis (see Clarke, 2001, for review).
Regardless of the source of LCFA available to adipose tissue, the ultimate pathway for
accretion of lipid stores is the series of enzymatic steps involved with attachment of LCFA to
glycerol-3-phosphate (fig. 1). Considerable progress has been made in the last decade in char-
acterizing the esterification pathway (see review by Rule, 1995, and references therein).
Esterification activity is greater in bovine subcutaneous adipose tissue than in liver (Wilson
et al., 1992). Evidence indicates that phosphatidate phosphohydrolase may be the rate-regulating
step in adipose but that glycerol-3-phosphate acyltransferase may be more likely to control
esterification in liver (Wilson et al., 1992; Smith et al., 1998). Changes in palmitate esterifi-
cation and activity of glycerol-3-phosphate acyltransferase parallel changes in fat deposition
and carcass fat thickness with different rates of gain and degrees of maturity in both bovine
and ovine adipose tissue (Bouyekhf et al., 1992; West et al., 1994; Andersen et al., 1996).
Palmitate esterification in vitro was decreased by >48% by the catecholamines clenbuterol,
norepinephrine, and isoproterenol as well as by cAMP in subcutaneous adipose tissue from ewes
fed at maintenance; however, for tissue from ewes fed on a high-energy diet for 6 weeks, only
isoproterenol and cAMP inhibited esterification but to a lesser degree than for maintenance-fed
lambs (Bouyekhf et al., 1993). A period of 72 h of starvation decreased LCFA esterification
and phosphatidate phosphohydrolase activity by about 50% in bovine subcutaneous adipose
tissue, but not in intramuscular adipose tissue (Smith et al., 1998).
Changes in adipose esterification rates in parallel with positive energy balance and during
fattening suggest a major control by insulin. However, insulin has only modest effects
on esterification activity in vitro (Jacobi and Miner, 2002), suggesting that removal of
340 J. K. Drackley

antagonistic effects during positive energy balance may be the more dominant regulatory
influence. Other factors may be involved in stimulating esterification. For example, acylation-
stimulating protein is an 8.9 kD protein generated within adipose tissue from components of
the alternative complement pathway (see Cianflone et al., 1999 for review). This protein acts
in an autocrine fashion to promote uptake and esterification of LCFA by adipocytes from
humans and laboratory rodents as well as cell lines. Recently, Jacobi and Miner (2002)
demonstrated that human acylation-stimulating protein increased acetate incorporation into
LCFA by 15–30% and oleate incorporation into TG by 10–25% in bovine adipose tissue, and
that this stimulation was not affected by feed restriction. It will be of interest to isolate the
bovine counterpart of acylation-stimulating protein and to determine its role in growth under
different dietary regimes. For example, in other species, a factor carried by chylomicrons after
a meal is important in activating the synthesis of acylation-stimulating protein (Cianflone
et al., 1999); does increased synthesis of acylation-stimulating protein during fat feeding in
ruminants contribute to the general increase in body fat deposition under those conditions?
In growing ruminants consuming adequate feed, the primary state is lipid accretion so that
lipid mobilization is a less important process. This contrasts with lactating ruminants, which
regularly undergo periods of intense lipid mobilization after parturition to support lactation
(McNamara, 1994). Situations do occur with growing ruminants that demand lipid mobiliza-
tion from adipose tissues to meet energy needs, for example, animals grazing poor-quality
pasture (due to environmental conditions or poor management). Agents that increase intra-
cellular cAMP concentrations, such as epinephrine and norepinephrine, are the primary stimuli
for increased lipolysis in sheep and cattle (Etherton et al., 1977) and exert lipolytic activity via
binding to β-adrenergic receptors (Houseknecht et al., 1996). The adrenergic receptors, through
protein kinase A, lead to phosphorylation and activation of hormone-sensitive lipase (HSL).
In laboratory animals, proteins called perilipins also are phosphorylated by β-adrenergic acti-
vation (Clifford et al., 2000). When phosphorylated, perilipins appear to increase access of
activated HSL to the lipid droplet, and thereby promote lipolysis.
Stimulatory effects of catecholamines on lipolysis in ruminants are enhanced by fasting
(DiMarco et al., 1991). Glucagon does not stimulate lipolysis in adipose tissue from either
sheep or cattle (Etherton et al., 1977). Somatotropin increases responsiveness of adipose
tissue to catecholamine-stimulated lipolysis (for review see Etherton and Bauman, 1998), but
somatotropin, insulin-like growth factors I and II, prolactin, and placental lactogen are with-
out effect on lipolysis in ruminants (Houseknecht et al., 1996). Prostaglandins (specifically
PGE2) also are not involved with regulation of lipolysis in bovine adipose tissue, as demon-
strated by the lack of effect of both exogenous PGE2 and indomethacin, which blocks
prostaglandin synthetase, on basal and epinephrine-stimulated lipolysis (DiMarco et al.,
1991). Although insulin is known to suppress lipolysis, direct effects of insulin to inhibit basal
or stimulated lipolysis have been difficult to demonstrate in vitro (DiMarco et al., 1991).
Basal and stimulated lipolytic rates of adipose tissues increase with growth and fattening
(Smith et al., 1984; Rule et al., 1992) but are not affected by diet (high forage vs. high grain) at
equal energy intake (Smith et al., 1984). Lipolytic rates generally are higher in subcutaneous
adipose than in internal depots (Etherton et al., 1977; Rule et al., 1992), although differences
among depots seem to diminish with increased age or body size (Rule et al., 1992). As animals
grow and fatten, lipolysis becomes less complete, resulting in greater release of NEFA than of
glycerol (Smith et al., 1984); this may indicate greater re-esterification activity within
adipocytes that contributes to increased fat deposition. Hansen et al. (1995) found that glycerol
release was greater from subcutaneous adipose than from perirenal adipose from sheep
at market weight, but that fatty acid release and tissue fatty acid pool size were greater in
Interorgan lipid and fatty acid metabolism 341

perirenal than in subcutaneous adipose. Consequently, lipolysis was less complete in the inter-
nal depot than in the subcutaneous depot.
An exciting and expanding field concerns the impact of various factors produced by
adipocytes that may act in autocrine, paracrine, or endocrine fashion to alter metabolism
(Vernon et al., 1999; Chilliard et al., 2000). One of these, leptin, has received tremendous
attention in both the biomedical arena (for review see Reidy and Weber, 2000) and in animal
agriculture (for reviews see Hossner, 1998; Houseknecht et al., 1998; Chilliard et al., 2001;
Ingvartsen and Boisclair, 2001). As adipocytes fill with lipid, leptin synthesis and secretion
increases. Leptin acts centrally to decrease feed intake and increase thermogenesis, and acts
directly to increase lipolysis in adipose tissue and NEFA oxidation in peripheral tissues.
Another autocrine or paracrine factor produced by adipocytes is TNFα, which acts to increase
lipolysis and suppress LPL (Gasic et al., 1999). Expression of TNFα mRNA appears to change
in bovine subcutaneous adipose tissue with degree of fatness (Drackley et al., unpublished
observations) as in other species. The role of these adipose-derived regulatory factors in
counteracting or decreasing the efficiency of fattening in ruminants will be an area of consider-
able interest in the future.

6.3. Liver

Because the liver of ruminant animals is not a major site of de novo synthesis of fatty acids
(Hanson and Ballard, 1967), its role lies in processing and redistributing exogenous and
endogenous LCFA taken up from the diet and from plasma lipoproteins, as well as in metab-
olism of NEFA mobilized from adipose tissue. In growing ruminants, the latter usually is a
minor process. Glycolytic and lipogenic enzymes are present in the milk-fed calf, but specific
activities decrease sharply after weaning (Pearce and Unsworth, 1980). The rate of VLDL
synthesis and secretion in liver of ruminants is lower than in rats (Kleppe et al., 1988; Graulet
et al., 1998), which is associated with the low rate of hepatic de novo fatty acid synthesis as
observed in other species (Pullen et al., 1990).
In fasted preruminant calves, considerable uptake of VLDL by the liver has been noted
(Bauchart et al., 1989). Furthermore, heavy, lipid-poor HDL were also taken up by the liver
of fasting calves, with concomitant production and release of lighter, lipid-enriched (prima-
rily with cholesterol esters) HDL particles (Bauchart et al., 1989); no appreciable flux of LDL
was noted. Uptake of VLDL by the liver changed to secretion as dietary tallow intake
increased in calves (Auboiron et al., 1995). Addition of supplemental methionine, which
might be postulated to be limiting for hepatic synthesis of apo-B, slightly increased secretion
of VLDL (Auboiron et al., 1995). Secretion of VLDL by calf liver slices occurred at 6- to 18-fold
lower rates than in rat liver slices (Graulet et al., 1998), despite similar rates of synthesis
of apoB in liver slices (Gruffat-Mouty et al., 1999). This lower rate of VLDL synthesis or
secretion may be attributable to a low rate of synthesis of TG in the microsomal compartment
responsible for VLDL assembly, with no limitation in the ability to synthesize and store
cytosolic TG (Graulet et al., 1998).
Hepatic lipid metabolism in preruminants may be sensitive to the LCFA profile of the diet,
as replacement of tallow with hydrogenated coconut oil (rich in medium-chain saturated
fatty acids and deficient in unsaturated LCFA) or soybean oil (high in polyunsaturated
LCFA) caused accumulation of TG in the liver of calves (Jenkins and Kramer, 1986;
Leplaix-Charlat et al., 1996; Piot et al., 1999). This suggests that changes in the profile of
dietary LCFA that are markedly different from that of bovine milk fat cause alterations in
hepatic lipid metabolism.
342 J. K. Drackley

Accumulation of TG in liver of preruminant calves fed milk replacer containing soybean


oil appeared to result from secretion of VLDL that were enriched with cholesterol esters
rather than TG, which consequently decreased hepatic secretion of TG (Leplaix-Charlat et al.,
1996). Because of the low oxidation of linoleic acid (Lindsay and Leat, 1977; Reid and
Husbands, 1985), high dietary linoleate supply could have decreased hepatic LCFA oxidation
and so contributed to TG accumulation. Liver uptake of very large and light HDL enriched
with cholesterol esters was observed, and appeared to be increased by including cholesterol
in the diet with soybean oil. This corresponded with simultaneous secretion of heavy HDL.
Liver slices from calves fed coconut oil-enriched milk replacers had lower rates of LCFA
oxidation and greater rates of LCFA esterification than did liver slices from calves fed an
equal amount of tallow (Graulet et al., 2000). Liver from coconut oil-fed calves also had
decreased rates of apo-B synthesis (Gruffat-Mouty et al., 2001); consequently, greater rates
of TG synthesis coupled with lower rates of VLDL export likely explain the accumulation of
TG in liver of calves fed coconut oil. Since milk fat contains more short- and medium-chain
fatty acids, which are extensively metabolized by the liver, than does tallow, and less PUFA
than does soybean oil, it would be of interest to compare the effects caused by coconut oil and
soybean oil with milk-fat controls rather than tallow controls to discern more about the native
metabolism in liver of preruminant calves.
Regulation of hepatic LCFA metabolism in ruminants has been reviewed extensively (Bell,
1980; Bauchart, 1993; Grummer, 1993; Bauchart et al., 1996; Hocquette and Bauchart, 1999;
Drackley et al., 2001). Liver removal of NEFA increases as arterial concentration of NEFA
increases due to changing intake in growing beef steers (Lapierre et al., 2000). Increased
uptake of NEFA mobilized from adipose tissue may lead to increases in TG formation and
ketone body synthesis, although these changes are much smaller in magnitude than those
observed during the negative energy balance of early lactation in dairy cows. For example, a
9-day starvation period in growing steers only increased liver TG concentration from 0.47% to
1.38% of wet weight, whereas plasma NEFA concentration increased to 1.03 mM (Lyle et al.,
1984). Ketone body concentrations in that study were only modestly elevated (BHBA =
6.9 mg/dl). Consequently, consideration of hepatic TG accumulation or excessive ketone
body production is not of major concern in growing ruminants.
Oxidation of LCFA in liver of preruminants and ruminants occurs in mitochondria and
peroxisomes. Peroxisomes are subcellular organelles that possess a pathway for β-oxidation
of fatty acids in which the first oxidation step is not coupled to ATP production but which
results in the dissipation of heat (see Reddy and Hashimoto, 2001, for review). Peroxisomal
β-oxidation is believed to function to oxidize fatty acids that are poor substrates for mito-
chondrial oxidative enzymes and to help process fatty acids when cellular uptake greatly
exceeds the cells’ need for energy. Peroxisomal β-oxidation in liver of ruminants was first
studied by Grum et al. (1994), who found that hepatic tissue capacity for peroxisomal
β-oxidation constituted about 50% of the total β-oxidation capacity in liver from mature dairy
cows, which was much greater than capacity measured in retired female breeder rats (26%).
Subsequent studies showed that peroxisomal β-oxidation constituted from 44% to 48% of
total β-oxidation in liver from sheep at market weight (Hansen et al., 1995). Absolute rates of
peroxisomal β-oxidation and the contribution to total β-oxidation capacity in isolated liver
tissue were lower in preruminant calves than in 15-month-old bulls (Piot et al., 1998). The
potential role and importance of hepatic peroxisomal β-oxidation during growth is uncertain.
Surprisingly little is known about the development of hepatic tissue capacity for metabo-
lism of LCFA during the transition from preruminant through weaning to functioning
ruminant. Ontogenic studies of key regulatory enzymes and pathways for fatty acid oxidation
Interorgan lipid and fatty acid metabolism 343

and esterification, as well as for VLDL assembly and secretion, would seem to be of interest
and importance to a full understanding of the metabolism of ruminant growth.

7. FUTURE PERSPECTIVES
Regulation of interorgan lipid metabolism during growth will continue to be an important
topic for biological scientists studying ruminant animals. Although much has been learned
about metabolism of LCFA and other lipids during the last decade, there are numerous areas
that lack fundamental understanding. Notable among these is the inability to alter rates of
development and filling of intramuscular adipocytes, which is critically important to eating
qualities of meat. An understanding of genetic and physiological mechanisms that control
adipose development among different depots would be of enormous benefit in development
of strategies to minimize lipid accumulation in internal and subcutaneous depots while max-
imizing marbling. The role of exogenous (dietary) LCFA, delivered by intestinally derived
chylomicrons and VLDL, as substrates for TG synthesis in intramuscular adipose also remains
of interest given the apparently low rate of de novo lipogenesis in this tissue. Another area
where information is unexpectedly incomplete is the developmental changes in lipid meta-
bolism in key organs, especially the liver, as ruminants make the transition from preruminant
through weaning to functioning ruminant.
Additional progress will certainly be made in the near future as techniques of molecular biol-
ogy are increasingly applied to questions concerning ruminant growth. The increasing ease of
measuring specific mRNA concentrations, as well as transcription rates, will likely lead to a
more complete understanding of how lipid metabolism is regulated through transcriptional and
post-transcriptional means. Furthermore, high-throughput techniques such as microarray analy-
sis of global changes in gene expression can be applied strategically, both to test hypotheses and
to search for novel genes that are regulated differently between physiological states, genetic
backgrounds, disease states, or applications of nutritional or pharmacological treatments.
Finally, impacts of the external animal environment on lipid metabolism during growth is an
area that should see additional exploration. Increasing knowledge of the multiway communi-
cation among the neuroendocrine system, the central nervous system, the immune system, and
intermediary metabolism of numerous organs (see e.g. Johnson, 1997; Kelley, 2001; and
Chapter 4 by Johnson and Escobar, this volume) offers exciting possibilities to improve both
efficiency of ruminant animal production and animal well-being. Such efforts are critical to the
long-term sustainability of ruminant animal production.

REFERENCES
Andersen, M.K., Bailey, J.W., Wilken, C., Rule, D.C., 1996. Lipoprotein lipase and glycerophosphate
acyltransferase in ovine tissues are influenced by growth and energy intake regimen. J. Nutr. Biochem.
7, 610–616.
Armand, M., Borel, P., Dubois, C., Senft, M., Peyrot, J., Salducci, J., Lafont, H., Lairon, D., 1994.
Characterization of emulsions and lipolysis of dietary lipids in the human stomach. Amer. J. Physiol.
266, G372–G381.
Ashes, J.R., Fleck, E., Scott, T.W., 1995. Dietary manipulation of membrane lipids and its implications
for their role in the production of second messengers. In: Von Englehardt, W., Leonhard-Marek, S.,
Breves, G., Giesecke, D. (Eds.), Ruminant Physiology: Digestion, Metabolism, Growth and
Reproduction. Ferdinand Enke Verlag, Stuttgart, Germany, pp. 373–386.
Auboiron, S., Durand, D., Robert, J.C., Chapman, M.J., Bauchart, D., 1995. Effects of dietary fat and
L-methionine on the hepatic metabolism of very low density lipoproteins in the preruminant calf, Bos
spp. Reprod. Nutr. Dev. 35, 167–178.
344 J. K. Drackley

Barber, M.C., Ward, R.J., Richards, S.E., Salter, A.M., Buttery, P.J., Vernon, R.G., Travers, M.T., 2000.
Ovine adipose tissue monounsaturated fat content is correlated to depot-specific expression of the
stearoyl-CoA desaturase gene. J. Anim. Sci. 78, 62–68.
Bartelds, B., Gratama, J.W., Knoester, H., Takens, J., Smid, G.B., Aarnoudse, J.G., Heymans, H.S.,
Kuipers, J.R., 1998. Perinatal changes in myocardial supply and flux of fatty acids, carbohydrates,
and ketone bodies in lambs. Amer. J. Physiol. 274, H1962–H1969.
Bauchart, D., 1993. Lipid absorption and transport in ruminants. J. Dairy Sci. 76, 3864–3881.
Bauchart, D., Durand, D., Laplaud, P.M., Forgez, P., Goulinet, S., Chapman, M.J., 1989. Plasma lipopro-
teins and apolipoproteins in the preruminant calf, Bos spp.: density distribution, physicochemical
properties, and the in vivo evaluation of the contribution of the liver to lipoprotein homeostasis.
J. Lipid Res. 30, 1499–1514.
Bauchart, D., Gruffat, D., Durand, D., 1996. Lipid absorption and hepatic metabolism in ruminants. Proc.
Nutr. Soc. 55, 39–47.
Bell, A.W., 1980. Lipid metabolism in liver and selected tissues and in the whole body of ruminant
animals. Prog. Lipid Res. 18, 117–164.
Bell, A.W., Thompson, G.E., 1979. Free fatty acid oxidation in bovine muscle in vivo: effects of cold
exposure and feeding. Amer. J. Physiol. 237, E309–E315.
Bennis, A., de La Farge, F., Bézille, P., Valdiguié, P., Rico, A.G., Braun, J.P., 1992. Effects of age of
newborn and delivery by female goats on plasma lipids and lipoproteins. Small Ruminant Res. 9,
243–253.
Benson, J.A., Reynolds, C.K., 2001. Effects of abomasal infusion of long-chain fatty acids on splanch-
nic metabolism of pancreatic and gut hormones in lactating dairy cows. J. Dairy Sci. 84, 1488–1500.
Bonnet, M., Faulconnier, Y., Fléchet, J., Hocquette, J.F., Leroux, C., Langin, D., Martin, P., Chilliard, Y.,
1998. Messenger RNAs encoding lipoprotein lipase, fatty acid synthase and hormone-sensitive lipase
in the adipose tissue of underfed-refed ewes and cows. Reprod. Nutr. Dev. 38, 297–307.
Bonnet, M., Leroux, C., Faulconnier, Y., Hocquette, J.F., Bocquier, F., Martin, P., Chilliard, Y., 2000.
Lipoprotein lipase activity and mRNA are upregulated by refeeding in adipose tissue and cardiac
muscle of sheep. J. Nutr. 130, 749–756.
Borel, P., Armand, M., Ythier, P., Dutot, G., Melin, C., Senft, M., Lafont, H., Lairon, D., 1994. Hydrolysis
of emulsions with different triglycerides and droplet sizes by gastric lipases in vitro: effect on
pancreatic lipase activity. J. Nutr. Biochem. 5, 124–133.
Bouyekhf, M., Rule, D.C., Hu, C.Y., 1992. Glycerolipid biosynthesis in adipose tissue of the bovine
during growth. Comp. Biochem. Physiol. B 103, 101–104.
Bouyekhf, M., Rule, D.C., Hu, C.Y., 1993. Effect of catecholamines on lipolysis and esterification in
vitro in adipose tissue of sheep fed low and high energy diets. J. Nutr. Biochem. 4, 80–85.
Brandstetter, A., Picard, B., Geay, Y., 1997. Regional variations of muscle fiber characteristics in
m. semitendinosus of growing cattle. J. Muscle Res. Cell Motil. 18, 57–62.
Bremmer, D.R., Ruppert, L.D., Clark, J.H., Drackley, J.K., 1998. Effects of chain length and unsatura-
tion of fatty acid mixtures infused into the abomasum of lactating dairy cows. J. Dairy Sci. 81,
176–188.
Campbell, R.G., 1988. Nutritional constraints to lean tissue accretion in farm animals. Nutr. Res. Rev.
1, 233–253.
Chen, M., Yang, Y., Braunstein, E., Georgeson, K.E., Harmon, C.M., 2001. Gut expression and regula-
tion of FAT/CD36: possible role in fatty acid transport in rat enterocytes. Amer. J. Physiol. 281,
E916–E923.
Cheng, A.L.S., Morehouse, M.G., Duell, J.J., 1949. The effect of the level of dietary calcium and mag-
nesium on the digestibility of fatty acids, simple triglycerides and some natural and hydrogenated fats.
J. Nutr. 37, 237–250.
Chien, D., Dean, D., Saha, A.K., Flatt, J.P., Ruderman, N.B., 2000. Malonyl-CoA content and fatty acid
oxidation in rat muscle and liver in vivo. Amer. J. Physiol. 279, E259–E265.
Chilliard, Y., 1993. Dietary fat and adipose tissue metabolism in ruminants, pigs, and rodents: a review.
J. Dairy Sci. 76, 3897–3931.
Chilliard, Y., Bonnet, M., Delavaud, C., Faulconnier, Y., Leroux, C., Djiane, J., Bocquier, F., 2001. Leptin
in ruminants: gene expression in adipose tissue and mammary gland, and regulation of plasm
concentration. Domest. Anim. Endocrinol. 21, 271–295.
Chilliard, Y., Ferlay, A., Faulconnier, Y., Bonnet, M., Rouel, J., Bocquier, F., 2000. Adipose tissue metab-
olism and its role in adaptations to undernutrition in ruminants. Proc. Nutr. Soc. 59, 127–134.
Interorgan lipid and fatty acid metabolism 345

Christensen, R.A., Drackley, J.K., LaCount, D.W., Clark, J.H., 1994. Infusion of four long-chain fatty
acid mixtures into the abomasum of lactating dairy cows. J. Dairy Sci. 77, 1052–1069.
Cianflone, K., Maslowska, M., Sniderman, A.D., 1999. Acylation stimulating protein (ASP), an
adipocyte autocrine: new directions. Sem. Cell Dev. Biol. 10, 31−41.
Clarke, S.D., 2001. Polyunsaturated fatty acid regulation of gene transcription: a molecular mechanism
to improve the metabolic syndrome. J. Nutr. 131, 1129–1132.
Clifford, G.M., Londos, C., Kraemer, F.B., Vernon, R.G., Yeaman, S.J., 2000. Translocation of hormone-
sensitive lipase and perilipin upon lipolytic stimulation of rat adipocytes. J. Biol. Chem. 275,
5011–5015.
Cordle, S.R., Kroper, W.J., Yeaman, S.J., 1986. Isolation and characterization of bovine lipoproteins:
incubation with hepatic lipase and isolated adrenal zona fasiculata cells. Biochem. Soc. Trans. 13,
879–882.
Cortright, R.N., Muoio, D.M., Dohm, G.L., 1997. Skeletal muscle lipid metabolism: a frontier for new
insights into fuel homeostasis. J. Nutr. Biochem. 8, 228–245.
Davidson, N.O., Shelness, G.S., 2000. Apolipoprotein B: mRNA editing, lipoprotein assembly, and
presecretory degradation. Annu. Rev. Nutr. 20, 169–193.
Davis, C.L., Drackley, J.K., 1998. The Development, Nutrition, and Management of the Young Calf. Iowa
State University Press, Ames, IA.
DiMarco, N.M., Rule, D.C., Whitehurst, G.B., Beitz, D.C., 1991. Effect of indomethacin, epinephrine,
prostaglandin E2 and insulin on lipolysis in bovine adipose tissue in vitro. Int. J. Biochem. 23, 1231−1235.
Doppenberg, J., Palmquist, D.L., 1991. Effect of dietary fat level on feed intake, growth, plasma metabo-
lites and hormones of calves fed dry or liquid diets. Livest. Prod. Sci. 29, 151–166.
Doreau, M., Chilliard, Y., 1997. Digestion and metabolism of dietary fat in farm animals. Brit. J. Nutr.
78, Suppl. 1, S15–S35.
Drackley, J.K., 2000. Lipid metabolism. In: D’Mello, J.P.F. (Ed.), Farm Animal Metabolism and
Nutrition: Critical Reviews. CAB International, Wallingford, UK, pp. 97–119.
Drackley, J.K., Klusmeyer, T.H., Trusk, A.M., Clark, J.H., 1992. Infusion of long-chain fatty acids varying
in saturation and chain length into the abomasum of lactating dairy cows. J. Dairy Sci. 75, 1517–1526.
Drackley, J.K., Overton, T.R., Douglas, G.N., 2001. Adaptations of glucose and long-chain fatty acid metab-
olism in liver of dairy cows during the periparturient period. J. Dairy Sci. 84, Suppl. E, E100–E112.
Drackley, J.K., Thire, S., Litherland, N.B., Beaulieu, A.D., 2000. Dry matter intake is decreased more by
abomasal infusion of unsaturated free fatty acids than by unsaturated triglycerides. J. Dairy Sci. 83,
Suppl. 1, 286 (Abstr.).
Edwards-Webb, J.D., Thompson, S.Y., 1978. Studies on lipid digestion in the preruminant calf. 3. The
action of salivary lipase on milk fat in the abomasum. Brit. J. Nutr. 40, 125–131.
Enser, M., 1973. Clearing-factor lipase in muscle and adipose tissue of pigs. Biochem. J. 136, 381–385.
Etherton, T.D., Bauman, D.E., 1998. Biology of somatotropin in growth and lactation of domestic animals.
Phyiol. Rev. 78, 745–761.
Etherton, T.D., Bauman, D.E., Romans, J.R., 1977. Lipolysis in subcutaneous and perirenal adipose
tissue from sheep and dairy steers. J. Anim. Sci. 44, 1100–1106.
Fallon, R.J., Williams, P.E.V., Innes, G.M., 1986. The effects on feed intake, growth and digestibility of
nutrients of including calcium soaps of fat in diets for young calves. Anim. Feed Sci. Tech. 14,
103–115.
Faulconnier, Y., Bonnet, M., Bocquier, F., Leroux, C., Chilliard, Y., 2001. Effects of photoperiod and
feeding level on adipose tissue and muscle lipoprotein lipase activity and mRNA level in dry non-
pregnant sheep. Brit. J. Nutr. 85, 299–306.
Fielding, C.J., Fielding, P.E., 1995. Molecular physiology of reverse cholesterol transport. J. Lipid Res.
36, 211–228.
Fielding, C.J., Fielding, P.E., 2001. Cellular cholesterol efflux. Biochim. Biophys. Acta 1533, 175–189.
Forte, T.M., Bell-Quint, J.J., Cheng, F., 1981. Lipoproteins of fetal and newborn calves and adult steers:
a study of developmental changes. Lipids 16, 240–245.
Freeman, C.P., 1984. The digestion, absorption and transport of fats–non-ruminants. In: Wiseman, J. (Ed.),
Fats in Animal Nutrition. Butterworths, London, pp. 105–122.
Gagliostro, G., Chilliard, Y., 1991. Duodenal rapeseed oil infusion in early and midlactation cows. 2.
Voluntary intake, milk production, and composition. J. Dairy Sci. 74, 499–509.
Gargouri, Y., Moreau, H., Verger, R., 1989. Gastric lipases: biochemical and physiological studies.
Biochim. Biophys. Acta 1006, 255–271.
346 J. K. Drackley

Gasic, S., Tian, B., Green, A., 1999. Tumor necrosis factor α stimulates lipolysis in adipocytes by decreas-
ing Gi protein concentrations. J. Biol. Chem. 274, 6770–6775.
Glatz, J.F., van der Vusse, G.J., 1996. Cellular fatty acid-binding proteins: their function and physiological
significance. Prog. Lipid Res. 35, 243–282.
Gooden, J.M., 1973. The importance of lipolytic enzymes in milk-fed and ruminating calves. Austr. J. Biol.
Sci. 26, 1189–1199.
Graham, D.L., Oram, J.F., 1987. Identification and characterization of a high density lipoprotein-binding
protein in cell membranes by ligand blotting. J. Biol. Chem. 262, 7439–7442.
Graulet, B., Gruffat, D., Durand, D., Bauchart, D., 1998. Fatty acid metabolism and very low density
lipoprotein secretion in liver slices from rats and preruminant calves. J. Biochem. 124, 1212–1219.
Graulet, B., Gruffat-Mouty, D., Durand, D., Bauchart, D., 2000. Effects of milk diets containing beef
tallow or coconut oil on the fatty acid metabolism of liver slices from preruminant calves. Brit. J.
Nutr. 84, 309–318.
Greeve, J., Altkemper, I., Dieterich, J.H., Greten, H., Windler, E., 1993. Apolipoprotein mRNA editing
in 12 different mammalian species: hepatic expression is reflected in low concentrations of apo B-
containing plasma lipoproteins. J. Lipid Res. 34, 1367–1383.
Gruffat-Mouty, D., Graulet, B., Durand, D., Samson-Bouma, M.E., Bauchart, D., 1999. Apolipoprotein
B production and very low density lipoprotein secretion by calf liver slices. J. Biochem. 126,
188–193.
Gruffat-Mouty, D., Graulet, B., Durand, D., Samson-Bouma, M.E., Bauchart, D., 2001. Effects of dietary
coconut oil on apolipoprotein synthesis and VLDL secretion by calf liver slices. Brit. J. Nutr. 86,
13–19.
Grum, D.E., Hansen, L.R., Drackley, J.K., 1994. Peroxisomal β-oxidation of fatty acids in bovine and rat
liver. Comp. Biochem. Physiol. B 109, 281–292.
Grummer, R.R., 1993. Etiology of lipid-related metabolic disorders in periparturient dairy cows. J. Dairy
Sci. 76, 3882–3896.
Ha, Y.C., Barter, P.J., 1982. Differences in plasma cholesteryl ester transfer activity in sixteen vertebrate
species. Comp. Biochem. Physiol. B 71, 265–269.
Hansen, L.R., Drackley, J.K., Berger, L.L., Grum, D.E., Cremin, J.D. Jr., Lin, X., Odle, J., 1995. Prenatal
androgenization of lambs: II. Metabolism in adipose tissue and liver. J. Anim. Sci. 73, 1701–1712.
Hanson, R.W., Ballard, F.J., 1967. The relative significance of acetate and glucose as precursors for lipid
synthesis in liver and adipose tissue from ruminants. Biochem. J. 105, 529–536.
Harrison, F.A., Leat, W.M.F., 1975. Digestion and absorption of lipids in non-ruminant and ruminant
animals: a comparison. Proc. Nutr. Soc. 34, 203–210.
Harrison, F.A., Leat, W.M.F., Foster, A., 1974. Absorption of maize oil infused into the duodenum of the
sheep. Proc. Nutr. Soc. 33, 103 (Abstr.).
Hocquette, J.F., Bauchart, D., 1999. Intestinal absorption, blood transport and hepatic and muscle metab-
olism of fatty acids in preruminant and ruminant animals. Reprod. Nutr. Dev. 39, 27–48.
Hocquette, J.F., Graulet, B., Olivecrona, T., 1998a. Lipoprotein lipase activity and mRNA levels in
bovine tissues. Comp. Biochem. Physiol. B121, 201–212.
Hocquette, J.F., Graulet, B., Vermorel, M., Bauchart, D., 2001. Weaning affects lipoprotein lipase
activity and gene expression in adipose tissues and in masseter but not in other muscles of the calf.
Brit. J. Nutr. 86, 433–441.
Hocquette, J.F., Ortigues-Marty, I., Pethick, D., Herpin, P., Fernandez, X., 1998b. Nutritional and hor-
monal regulation of energy metabolism in skeletal muscles of meat-producing animals. Livest. Prod.
Sci. 56, 115–143.
Hood, R.L., Allen, C.E., 1978. Lipogenesis in isolated intramuscular adipose tissue from four bovine
muscles. J. Anim. Sci. 46, 1626–1633.
Hood, R.L., Cook, L.J., Mills, S.C., Scott, T.W., 1980. Effect of feeding protected lipids on fatty acid
synthesis in ovine tissues. Lipids 15, 644–650.
Hossner, K.L., 1998. Cellular, molecular and physiological aspects of leptin: potential application in
animal production. Can. J. Anim. Sci. 78, 463–472.
Houseknecht, K.L., Baile, C.A., Matteri, R.L., Spurlock, M.E., 1998. The biology of leptin: a review.
J. Anim. Sci. 76, 1405–1420.
Houseknecht, K.L., Bauman, D.E., Vernon, R.G., Byatt, J.C., Collier, R.J., 1996. Insulin-like growth
factors-I and -II, somatotropin, prolactin, and placental lactogen are not acute effectors of lipolysis
in ruminants. Domest. Anim. Endocrinol. 13, 239–249.
Interorgan lipid and fatty acid metabolism 347

Ingle, D.L., Bauman, D.E., Garrigus, U.S., 1972. Lipogenesis in the ruminant: in vitro study of tissue
sites, carbon source and reducing equivalent generation for fatty acid synthesis. J. Nutr. 102, 609–616.
Ingvartsen, K.L., Boisclair, Y.R., 2001. Leptin and the regulation of food intake, energy homeostasis and
immunity with special focus on periparturient ruminants. Domest. Anim. Endocrinol. 21, 215–250.
Jacobi, S.K., Miner, J.L., 2002. Human acylation-stimulating protein and lipid biosynthesis in bovine
adipose tissue explants. J. Anim. Sci. 80, 751–756.
Jenkins, K.J., 1986. Fatty acid-binding protein in liver and small intestine of the preruminant calf.
J. Dairy Sci. 69, 155–159.
Jenkins, K.J., 1988. Factors affecting poor performance and scours in preruminant calves fed corn oil.
J. Dairy Sci. 71, 3013–3020.
Jenkins, K.J., Kramer, J.K.G., 1986. Influence of low linoleic and linolenic acids in milk replacer on calf
performance and lipids in blood plasma, heart, and liver. J. Dairy Sci. 69, 1374–1386.
Jenkins, K.J., Kramer, J.K.G., 1990. Effects of dietary corn oil and fish oil concentrate on lipid compo-
sition of calf tissues. J. Dairy Sci. 73, 2940−2951.
Jenkins, K.J., Griffith, G., Kramer, J.K.G., 1988. Plasma lipoproteins in neonatal, preruminant, and
weaned calf. J. Dairy Sci. 71, 3003–3012.
Jenkins, K.J., Kramer, J.K.G., Emmons, D.B., 1986. Effect of lipids in milk replacers on calf perform-
ance and lipids in blood plasma, liver, and perirenal fat. J. Dairy Sci. 69, 447–459.
Jenkins, K.J., Kramer, J.K.G., Sauer, F.D., Emmons, D.B., 1985. Influence of triglycerides and free fatty
acids in milk replacers on calf performance, blood plasma, and adipose lipids. J. Dairy Sci. 68,
669–680.
Jenkins, T.C., 1993. Lipid metabolism in the rumen. J. Dairy Sci. 76, 3851–3863.
Jenkins, T.C., Thies, E.J., Fotouhi, N., 1994. Dietary soybean oil changes lipolytic rate and composition
of fatty acids in plasma membranes of ovine adipocytes. J. Nutr. 124, 566–570.
Jenness, R., 1985. Biochemical and nutritional aspects of milk and colostrum. In: Larson, B.L. (Ed.),
Lactation. Iowa State University Press, Ames, IA, pp. 164–197.
Johnson, R.W., 1997. Inhibition of growth by pro-inflammatory cytokines: an integrated view. J. Anim.
Sci. 75, 1244–1255.
Kelley, K.W., 2001. It’s time for psychoneuroimmunology. Brain Behav. Immun. 15, 1–6.
Kleppe, B.B., Aiello, R.J., Grummer, R.R., Armentano, L.E., 1988. Trigyceride accumulation and very
low density lipoprotein secretion by rat and goat hepatocytes in vitro. J. Dairy Sci. 71, 1813–1822.
Kris-Etherton, P.M., Etherton, T.D., 1982. The role of lipoproteins in lipid metabolism of meat animals.
J. Anim. Sci. 55, 804–817.
Kuehn, C.S., Otterby, D.E., Linn, J.G., Olson, W.G., Chester-Jones, H., Marx, G.D., Barmore, J.A., 1994.
The effect of dietary energy concentration on calf performance. J. Dairy Sci. 77, 2621–2629.
Langfort, J., Ploug, T., Ihlemann, J., Enevoldsen, L.H., Stallknecht, B., Saldo, M., Kjaer, M., Hom, C.,
Galbo, H., 1998. Hormone-sensitive lipase (HSL) expression and regulation in skeletal muscle. Adv.
Exp. Med. Biol. 441, 219–228.
Lapierre, H., Bernier, J.F., Dubreuil, P., Reynolds, C.K., Farmer, C., Ouellet, D.R., Lobley, G.E., 2000.
The effect of feed intake level on splanchnic metabolism in growing beef steers. J. Anim. Sci. 78,
1084–1099.
Laplaud, P.M., Bauchart, D., Durand, D., Beaubatie, L., Chapman, M.J., 1991. Intestinal lymph and plasma
lipoproteins in the preruminant calf: partial resolution of particle heterogeneity in the 1.040−1.090 g/mL
interval. J. Lipid Res. 32, 1429–1439.
Laplaud, P.M., Bauchart, D., Durand, D., Chapman, M.J., 1990. Lipoproteins and apolipoproteins in
intestinal lymph of the preruminant calf, Bos spp., at peak lipid absorption. J. Lipid Res. 31,
1781–1792.
Leplaix-Charlat, L., Durand, D., Bauchart, D., 1996. Effects of diets containing tallow and soybean oil
with and without cholesterol on hepatic metabolism of lipids and lipoproteins in the preruminant calf.
J. Dairy Sci. 79, 1826–1835.
Lindsay, D.B., Leat, W.M.F., 1977. Oxidation and metabolism of linoleic acid in fed and fasted sheep.
J. Agr. Sci. 89, 215–221.
Lyle, R.R., DeBoer, G., Harrison, R.O., Young, J.W., 1984. Plasma and liver metabolites and glucose
kinetics as affected by prolonged ketonemia-glucosuria and fasting in steers. J. Dairy Sci. 67,
2274–2282.
Marcos, E., Mazur, A., Cardot, P., Coxam, V., Rayssiguier, Y., 1991. Developmental changes in plasma
apolipoproteins B and A-I in fetal bovines. Biol. Neonate 59, 22–29.
348 J. K. Drackley

McNamara, J.P., 1994. Lipid metabolism in adipose tissue during lactation: a model of a metabolic con-
trol system. J. Nutr. 124, 1383S–1391S.
Mendel, C.M., Kunitake, S.T., Kane, J.P., 1986. Discrimination between subclasses of human high-
density lipoproteins by the HDL binding sites of bovine liver. Biochim. Biophys. Acta 875, 59–68.
Miller, M.F., Cross, H.R., Lunt, D.K., Smith, S.B., 1991. Lipogenesis in acute and 48-hour cultures of
bovine intramuscular and subcutaneous adipose tissue explants. J. Anim. Sci. 69, 162–170.
Mills, S.E., Lemenager, R.P., Horstman, L.A., 1989. Adipose tissue lipogenesis in growing steers adapted
to different levels of feed intake. J. Anim. Sci. 67, 3011–3017.
Moibi, J.A., Christopherson, R.J., Okine, E.K., 2000a. In vivo and in vitro lipogenesis and aspects
of metabolism in ovines: effect of environmental temperature and dietary lipid supplementation. Can.
J. Anim. Sci. 80, 59–67.
Moibi, J.A., Christopherson, R.J., Okine, E.K., 2000b. Effect of environmental temperature and dietary
lipid supplement on activity and protein abundance of acetyl-CoA carboxylase and fatty acid synthase
in skeletal muscle and adipose tissues of sheep. Can. J. Anim. Sci. 80, 69–77.
Moore, J.H., Christie, W.W., 1984. Digestion, absorption and transport of fats in ruminant animals.
In: Wiseman, J. (Ed.), Fats in Animal Nutrition. Butterworths, London, pp. 123–149.
Moore, K.K., Cameron, P.J., Ekeren, P.A., Smith, S.B., 1993. Fatty acid binding protein in bovine longissimus
dorsi muscle. Comp. Biochem. Physiol. B 104, 259–266.
Nestel, P.J., Poyser, A., Hood, R.L., Mills, S.C., Willis, M.R., Cook, L.J., Scott, T.W., 1978. The effect of
dietary fat supplements on cholesterol metabolism in ruminants. J. Lipid Res. 19, 899–909.
Noble, R.C., 1978. Digestion, absorption and transport of lipids in ruminant animals. Prog. Lipid Res.
17, 55–91.
Noble, R.C., 1984. Essential fatty acids in the ruminant. In: Wiseman, J. (Ed.), Fats in Animal Nutrition.
Butterworths, London, pp. 185–200.
Noble, R.C., Shand, J.H., 1982. Fatty acid metabolism in the neonatal ruminant. In: Draper, H.H. (Ed.),
Advances in Nutritional Research, Vol. 4. Plenum Press, New York, pp. 287–337.
Noble, R.C., O’Kelly, J.C., Moore, J.H., 1972. Observations on the lecithin:cholesterol acyl transferase
system in bovine plasma. Biochim. Biophys. Acta 270, 519–528.
Noble, R.C., Shand, J.H., 1983. A comparative study of the distribution and fatty acid composition of the
lipoproteins in the fetal and maternal plasma of sheep. Biol. Neonate 44, 10–20.
Olivecrona, T., Olivecrona, G., 1999. Lipoprotein and hepatic lipases in lipoprotein metabolism.
In: Betteridge, D.J., Illingworth, D.R., Shepherd, J. (Eds.), Lipoproteins in Health and Disease.
Edward Arnold, London, pp. 223–246.
Ottemann-Abbamonte, C.J., Drackley, J.K., Beaulieu, A.D., Overton, T.R., Douglas, G.N., Palmquist, D.L.,
1998. Effect of acetate on palmitate oxidation in bovine and rat skeletal muscle tissue. FASEB J.
12, A835 (Abstr.).
Overton, T.R., Beaulieu, A.D., Zhu, J., Ortiz-Gonzalez, G., Drackley, J.K., Barbano, D.M., 1998. Dose
response to increasing amounts of high-oleic sunflower fatty acids infused into the abomasum of
lactating dairy cows. J. Dairy Sci. 81, Suppl. 1, 352 (Abstr.).
Page, A.M., Sturdivant, C.A., Lunt, D.K., Smith, S.B., 1997. Dietary whole cottonseed depresses lipoge-
nesis but has no effect on stearoyl-coenzyme desaturase activity in bovine subcutaneous adipose
tissue. Comp. Biochem. Physiol. B 118, 79–84.
Palmquist, D.L., 1976. A kinetic concept of lipid transport in ruminants. J. Dairy Sci. 59, 355–363.
Palmquist, D.L., 1991. Influence of source and amount of dietary fat on digestibility in lactating cows.
J. Dairy Sci. 74, 1354–1360.
Pearce, J., Unsworth, E.F., 1980. The effects of diet on some hepatic enzyme activities in the pre-ruminant
and ruminating calf. J. Nutr. 110, 255–261.
Perez, M.D., Sanchez, L., Aranda, P., Ena, J.M., Oria, R., Calvo, M., 1992. Effect of β-lactoglobulin on
the activity of pregastric lipase: a possible role for this protein in ruminant milk. Biochim. Biophys.
Acta 1123, 151–155.
Pethick, D.W., Dunshea, F.R., 1993. Fat metabolism and turnover. In: Forbes, J.M., France, J. (Eds.),
Quantitative Aspects of Ruminant Digestion and Metabolism. CAB International, Wallingford, UK,
pp. 291−311.
Phan, C.T., Tso, P., 2001. Intestinal lipid absorption and transport. Frontiers Biosci. 6, 299–319.
Piot, C., Hocquette, J.F., Veerkamp, J.H., Durand, D., Bauchart, D., 1999. Effects of dietary coconut oil
on fatty acid oxidative capacity of the liver, the heart and skeletal muscles in the preruminant calf.
Brit. J. Nutr. 82, 299–308.
Interorgan lipid and fatty acid metabolism 349

Piot, C., Hocquette, J.F., Herpin, P., Veerkamp, J.H., Bauchart, D., 2000. Dietary coconut oil affects more
lipoprotein lipase activity than the mitochondria oxidative capacities in muscles of preruminant
calves. J. Nutr. Biochem. 11, 231–238.
Piot, C., Veerkamp, J.H., Bauchart, D., Hocquette, J.F., 1998. Contribution of mitochondria and peroxi-
somes to palmitate oxidation in rat and bovine tissues. Comp. Biochem. Phyiol. B 121, 185–194.
Pothoven, M.A., Beitz, D.C., 1973. Effect of adipose tissue site, animal weight, and long-term fasting on
lipogenesis in the bovine. J. Nutr. 103, 468–475.
Pothoven, M.A., Beitz, D.C., Thornton, J.H., 1975. Lipogenesis and lipolysis in adipose tissue of ad libitum
and restricted-fed beef cattle during growth. J. Anim. Sci. 40, 957–962.
Pullen, D.L., Liesman, J.S., Emery, R.S., 1990. A species comparison of liver slice synthesis and secretion
of triacylglycerol from non-esterified fatty acids in media. J. Anim. Sci. 68, 1395–1399.
Reddy, J.K., Hashimoto, T., 2001. Peroxisomal β-oxidation and peroxisome proliferator-activated recep-
tor: an adaptive metabolic system. Annu. Rev. Nutr. 21, 193–230.
Reid, J.C.W., Husbands, D.R., 1985. Oxidative metabolism of long-chain fatty acids in mitochondria
from sheep and rat liver: evidence that sheep conserve linoleate by limiting its oxidation. Biochem.
J. 225, 233–237.
Reidy, S.P., Weber, J.M., 2000. Leptin: an essential regulator of lipid metabolism. Comp. Biochem.
Phyisiol. A 125, 285–297.
Reinhart, G.A., Simmen, F.A., Mahan, D.C., White, M.E., Roehrig, K.L., 1992. Intestinal development and
fatty acid binding protein activity of newborn pigs fed colostrum or milk. Biol. Neonate 62, 155–163.
Rudling, M.J., Peterson, C.O., 1985. LDL receptors in bovine tissues assayed as the heparin-sensitive
binding or 125I-labeled LDL in homogenates: relation between liver LDL receptors and serum
cholesterol in the fetus and post-term. Biochim. Biophys. Acta 836, 96–104.
Rule, D.C., 1995. Adipose tissue glycerolipid biosynthesis. In: Smith, S.B., Smith, D.R. (Eds.), The
Biology of Fat in Meat Animals. American Society of Animal Science, Champaign, IL, pp. 129–143.
Rule, D.C., Thornton, J.H., McGilliard, A.D., Beitz, D.C., 1992. Effect of adipose tissue site, animal size,
and fasting on lipolysis in bovine adipose tissue in vitro. Int. J. Biochem. 24, 789–793.
Senda, M., Oka, K., Brown, W.V., Qasba, P.K., Furuichi, Y., 1987. Molecular cloning and sequence of a
cDNA coding for bovine lipoprotein lipase. Proc. Natl. Acad. Sci. USA 84, 4369–4373.
Small, D.M., 1968. A classification of biological lipids based upon their interaction in aqueous systems.
J. Amer. Oil Chem. Soc. 45, 108–119.
Smith, S.B., 1995. Substrate utilization in ruminant adipose tissues. In: Smith, S.B., Smith, D.R. (Eds.),
The Biology of Fat in Meat Animals. American Society of Animal Science, Champaign, IL,
pp. 166–188.
Smith, S.B., Crouse, J.D., 1984. Relative contributions of acetate, lactate and glucose to lipogenesis in
bovine intramuscular and subcutaneous adipose tissue. J. Nutr. 114, 792–800.
Smith, S.B., Prior, R.L., 1986. Comparisons of lipogenesis and glucose metabolism between ovine and
bovine adipose tissues. J. Nutr. 116, 1279–1286.
Smith, S.B., Jenkins, T., Prior, R.L., 1987. Carcass composition and adipose tissue metabolism in grow-
ing sheep. J. Anim. Sci. 65, 1525–1530.
Smith, S.B., Lin, K.C., Wilson, J.J., Lunt, D.K., Cross, H.R., 1998. Starvation depresses acylglycerol
biosynthesis in bovine subcutaneous but not intramuscular adipose tissue homogenates. Comp.
Biochem. Physiol. B 120, 165–174.
Smith, S.B., Prior, R.L., Ferrell, C.L., Mersmann, H.J., 1984. Interrelationships among diet, age, fat
deposition and lipid metabolism in growing steers. J. Nutr. 114, 153–162.
Smith, S.B., Prior, R.L., Koong, L.J., Mersmann, H.J., 1992. Nitrogen and lipid metabolism in heifers fed
at increasing levels of intake. J. Anim. Sci. 70, 152–160.
Spanski, N.A., Drackley, J.K., Davis, C.L., Jaster, E.H., 1997. Utilization of supplemental triglycerides
or free fatty acids by calves from 4 to 10 weeks of age. J. Dairy Sci. 80, 573–585.
Sugden, M.C., Holness, M.J., Howard, R.M., 1993. Changes in lipoprotein lipase activities in adipose
tissue, heart and skeletal muscle during continuous or interrupted feeding. Biochem. J. 292, 113–119.
Tanaka, H., Mierau, I., Ito, F., 1999. Purification and characterization of bovine pancreatic bile salt-
activated lipase. J. Biochem. 125, 883–890.
Tikofsky, J.N., Van Amburgh, M.E., Ross, D.A., 2001. Effect of varying carbohydrate and fat content of
milk replacer on body composition of Holstein bull calves. J. Anim. Sci. 79, 2260–2267.
Toullec, R., Mathieu, C.M., 1969. Digestive utilization of fats and their principal fatty acids by calves:
influence on body composition. Ann. Biol. Anim. Biochem. Biophys. 9, 139–160 (in French).
350 J. K. Drackley

Tso, P., Balint, J.A., 1986. Formation and transport of chylomicrons by enterocytes to the lymphatics.
Amer. J. Phyiol. 250, G715–G726.
Tso, P., Liu, M., Kalogeris, T.J., Thomson, A.B.R., 2001. The role of apolipoprotein A-IV in the regulation
of food intake. Annu. Rev. Nutr. 21, 231–254.
Turley, S.D., Burns, D.K., Rosenfeld, C.R., Dietschy, J.M., 1996. Brain does not utilize low density
lipoprotein-cholesterol during fetal and neonatal development in the sheep. J. Lipid Res. 37,
1953–1961.
van den Bosch H., Aarsman, A.J., de Jong, J.G., van Deenem, L.L., 1993. Studies on lysophospholipases. I.
Purification and some properties of a lysophospholipase from beef pancreas. Biochim. Biophys. Acta
296, 94–104.
Vernon, R.G., 1980. Lipid metabolism in the adipose tissue of ruminant animals. Prog. Lipid Res. 19,
23–106.
Vernon, R.G., Barber, M.C., Travers, M.T., 1999. Present and future studies on lipogenesis in animals and
human subjects. Proc. Nutr. Soc. 58, 541–549.
Villeneuve, P., Pina, M., Graille, J., 1996. Determination of pregastric lipase specificity in young rumi-
nants. Chem. Phys. Lipids 83, 161–168.
Wang, C.S., Hartsuck, J.A., 1993. Bile salt-activated lipase: a multiple function lipolytic enzyme.
Biochim. Biophys. Acta 1166, 1–19.
West, T.R., Riley, M.L., Rule, D.C., 1994. Palmitate esterification and glycerophosphate acyltransferase
activity in adipose tissue of growing lambs. J. Anim. Sci. 72, 81–86.
Wilson, J.J., Young, C.R., Smith, S.B., 1992. Triacylglycerol biosynthesis in bovine liver and subcuta-
neous adipose tissue. Comp. Biochem. Physiol. B 103, 511–516.
Winder, W.W., 2001. Energy-sensing and signaling by AMP-activated protein kinase in skeletal muscle.
J. Appl. Physiol. 91, 1017–1028.
Wolfrum, C., Borrmann, C.M., Borchers, T., Spener, F., 2001. Fatty acids and hypolipidemic drugs regu-
late peroxisome proliferator-activated receptors alpha- and gamma-mediated gene expression via liver
fatty acid binding protein: a signaling path to the nucleus. Proc. Natl. Acad. Sci. USA 98, 2323–2328.
Young, M.E., Goodwin, G.W., Ying, J., Guthrie, P., Wilson, C.R., Laws, F.A., Taegtmeyer, H., 2001.
Regulation of cardiac and skeletal muscle malonyl-CoA decarboxylase by fatty acids. Amer. J. Physiol.
280, E471–E479.
Zhou, L., Nilsson, A., 2001. Sources of eicosanoid precursor fatty acid pools in tissues. J. Lipid Res. 42,
1521–1542.
Zhou, L., Vessby, B., Nilsson, A., 2002. Quantitative role of plasma free fatty acids in the supply of
arachidonic acid to extrahepatic tissues in rats. J. Nutr. 132, 2626–2631.
Zinn, R.A., 1989. Influence of level and source of dietary fat on its comparative feeding value in finish-
ing diets for feedlot steers: metabolism. J. Anim. Sci. 67, 1038–1049.
Zinn, R.A., 1992. Comparative feeding value of supplemental fat in steam-flaked corn- and steam-flaked
wheat-based finishing diets for feedlot steers. J. Anim. Sci. 70, 2959–2969.
PART IV
Carbohydrate and
energy metabolism
14 Environmental and hormonal regulation
of energy metabolism in early
development of the pig

P. Herpin, I. Louveau, M. Damon and J. Le Dividich

INRA, Joint Research Unit for Calf and Pig Production,


35590 Saint-Gilles, France

This chapter is concerned with the development and the environmental regulation of the energy
metabolism in the neonatal pig, a species devoid of brown adipose tissue. The newborn pig is
poorly insulated and maintenance of its homeothermic balance is essentially related to its ability
to produce heat. The skeletal muscle is the major contributor to the energy metabolism, with shiv-
ering being the main mechanism of cold-induced thermogenesis. The piglet does not express
UCP1. UCP2 and UCP3 are both expressed in skeletal muscle, but there is no evidence that they
play a role in the cold-induced thermogenesis. Major factors involved in the postnatal improve-
ment in the cold-induced thermogenesis include changes in muscle structure and cardiovascular,
biochemical and hormonal adjustments. Muscle maturation is suggested by the marked post-
natal increase in myofibrillar mass and the transitory expression of α-cardiac heavy chain myosin
and triad proliferation. Moreover, muscle mitochondria are functional at birth, and both the
increase in mitochondrial mass and their ultrastructural change account for the increased oxida-
tive potential of the muscle. Cardiovascular adjustments include the redistribution of the cardiac
output towards the skeletal muscle at the expense of the skin, liver and intestine. The effects of the
shift in the energy source from carbohydrate to fat at birth on the development of metabolic path-
ways including gluconeogenesis, lipogenesis and fatty acid oxidation are also examined.
Emphasis is given to the key role of CPT-1 in the regulation of fatty acid oxidation. Finally, the
actions of thyroid, HPA and, to a lesser extent, somatotropic axes on the regulation of the energy
metabolism are considered. It is concluded that the newborn pig is immature at birth in several
respects. As a whole, factors involved in the improvement of its postnatal thermogenic capacity
are all suggestive of the enhancement of this maturity within the first postnatal days.

1. INTRODUCTION
The neonatal period is attended by important modifications in several physiological functions
associated with dramatic changes in energy metabolism and nutrition. Before birth, the maternal
Biology of Metabolism in Growing Animals
D.G. Burrin and H. Mersmann (Eds.)
353 © 2005 Elsevier Limited. All rights reserved.
354 P. Herpin et al.

organism provides an appropriate environment so that the foetus does not have to actively reg-
ulate body temperature. At birth the newborn pig experiences a sudden and dramatic cold stress
since the gap between the dam’s uterus and the ambient temperature (Ta) may be as high as
10–12°C. The foetus receives a continuous supply of substrates for growth and metabolism
among which glucose is the main energy substrate (Père et al., 1995). At birth, the piglet is
abruptly switched to intermittent feeding of high-fat, low-carbohydrate colostrum and milk.
Because the ability to alter metabolic rate is a prerequisite for survival, successful adaptation
to this critical period implies that the piglet is able to activate its thermoregulatory mechanisms
and to provide energy to its heat-producing tissues. This, in turn, requires profound changes in
major metabolic pathways including oxidative, lipogenic and gluconeogenic pathways associ-
ated with changes in nutrition and hormonal status.
During the past years, several reviews have been devoted to development of the metabolic
responses to cold (Mount, 1968; Curtis, 1974) and the metabolic patterns in the neonatal pig
(Mersmann, 1974). Data presented in this chapter rely mostly on recent results which provide
new insights on the adaptation of the energy, fatty acids (oxidation, storage) and glucose
metabolism during the neonatal period and their environmental and hormonal control. The
actions of the thyroid, hypothalamic–pituitary–adrenal (HPA) and, to a lesser extent, soma-
totropic axis will be developed even though other hormones like insulin play an important
role in the control of metabolism.

2. OVERVIEW OF THE ENERGY METABOLISM


IN THE NEONATAL PIG
2.1. General aspects

Maintenance of homeothermia results from a dynamic balance between heat loss and heat
production. The newborn pig is poorly insulated, being virtually hairless and devoid of sub-
cutaneous fat, and although cold exposure induces some vasoconstriction, it does not reduce
the cardiac output to the skin (Lossec et al., 1999). This poor ability of the piglet to conserve
heat is reflected by the fact that each 1°C coldness is associated with a 2 kJ/h/kg BW increase
in heat production, which is 2.6-fold higher than at the time of weaning (Le Dividich et al.,
1998). In fact, the weight-specific requirement for energy of the neonatal pig is maximal at
birth. This is due to the high rate of heat production associated with thermoregulation, phys-
ical activity related to the establishment of the nursing order, and a high potential for growth.
At birth, the typical ambient temperature (Ta) of 24–20°C is 10–12°C below the lower criti-
cal temperature of the piglet and is close to the temperature of 18°C at which the metabolic
rate is maximal (Berthon et al., 1993). This period of cold stress is commonly associated with
a temporary fall in rectal temperature, the extent of which and the time taken for recovery
being dependent on body weight and Ta. In practice, this period of hypothermia may be a
cause of mortality by weakening the piglet and predisposing it to crushing by the sow and
starvation that is illustrated by the usual high level of mortality in the first 48 h after birth.
However, cold resistance improves markedly in the early postnatal period in fed piglets but
not in those consuming little or no colostrum (Le Dividich et al., 1991a; Herpin et al., 1994).
Because thermal insulation does not change substantially during the first postnatal days
(Berthon, 1994), maintenance of the homeothermic balance is largely dependent on the ability
of the newborn to produce heat. The newborn pig responds quickly and vigorously to cold stress
as exemplified by the 30% increase in metabolic rate at 18°C compared with 31°C within the
first 20 min after birth (Noblet and Le Dividich, 1981), with the difference increasing to 100%
Energy metabolism in early development of the pig 355

Table 1
Minimal (mmR) and maximal (MmR) metabolic rate, lower critical
temperature (LCT) and ambient temperature at which maximal
metabolic rate (TMmR) is reached in pigs aged from 2–48 h

Age, h

2 24 48

mmR, kJ/h/kg BW 12.9 16.9 20.2


MmR, kJ/h/kg BW 36.6 43.3 >46.8
LCT, °C 34.2 33.1 30.2
TMmR, °C 17.8 12.8 <10
Adapted from Berthon et al. (1993, 1994).

in the first 90 min (Herpin et al., 1994). Also (table 1), maximal and minimal metabolic rate
are increased by 56% and 28% respectively, during the first 48 h of life (Berthon, 1994).
Together, these indicate that mechanisms responsible for heat production are active soon after
birth. However, maintenance of a high metabolic rate during cold stress is closely dependent
on both the availability of energy substrates and on the ability of the piglets to utilize these as
an energy source. Once the piglet displays an efficient thermoregulatory behaviour and once
a regular milk intake is established, cold stress is usually of minor importance during the
remaining suckling period. The energy metabolism is then regulated essentially by the
amount of milk intake (Marion and Le Dividich, 1999).

2.2. Evidence for the major importance of muscle in energy metabolism

To cope with the neonatal cold challenge, skeletal muscle plays a central role in neonatal
energy metabolism (fig. 1). Indeed, the calculated contribution of skeletal muscle to total

Fig. 1. Total body O2 consumption and contribution of muscle to total body O2 consumption in relation to
age and ambient temperature (TN, thermoneutral; C, cold). Adapted from Lossec et al. (1998b).
356 P. Herpin et al.

oxygen consumption averages 34–40% at thermal neutrality and 50–64% in the cold. The
skeletal muscle is the major contributor to regulatory thermogenesis, accounting for 97% of
the cold-induced increase in heat production at 5 days of life (Lossec et al., 1998b). In addi-
tion, this stimulation of muscle energy metabolism is accompanied by a redistribution of
cardiac output towards skeletal muscles at the expense of digestive tract and liver blood flows
(Lossec et al., 1999) which, however, might be detrimental to the digestive processes through
the possible shortening of energy supply in the cold.

3. MECHANISMS OF HEAT PRODUCTION


3.1. Overview of heat production mechanisms

During cold exposure, maintenance of the homeothermic balance is achieved through the
development of specific heat-producing mechanisms. Two main mechanisms are usually
reported: shivering and non-shivering thermogenesis (NST). NST has been defined as “heat-
producing mechanisms due to processes that do not involve muscular contractions, such as
those involved in ion pumping or mitochondrial loose coupling mediated by uncoupling
protein-1 (UCP1) in brown fat”.
Shivering is defined as an involuntary rhythmic contraction of skeletal muscle myofibrils
involving no voluntary movements or external work. Heat production during shivering involves
biochemical mechanisms close to those associated with the contraction of the skeletal muscle
myofibril: heat is produced during both the hydrolysis of ATP and the associated processes of
ATP re-synthesis, and all energy substrates can support shivering. Tremendous amounts of heat
can be produced during shivering but this mechanism is not very efficient because it occurs at
the periphery of the body, and therefore enhances thermolysis, and it impairs physical move-
ments. In addition, heat produced during physical activity and meal consumption can contribute
significantly to the extra thermoregulatory heat produced in the cold.

3.2. Contribution of feeding to extra heat production

A major role of colostrum is the provision of immunoglobulins. However, owing to the low
body energy stores at birth, it is obvious that colostrum is of utmost importance in the provi-
sion of energy in the first day of life. This is evidenced by the fact that in cold conditions both
body temperature and heat production are positively related to the amount of colostrum intake
(Noblet and Le Dividich, 1981). Moreover, ingestion of colostrum results in metabolic heat
production (Gentz et al., 1970). This metabolic response, referred to as postprandial thermo-
genesis, represents the energy cost associated with digestion, absorption and processing of
nutrients. However, its contribution to thermoregulation is marginal, accounting for about
10% of the extra heat produced in the cold (Herpin et al., 1994), probably because of the high
(0.91) efficiency of colostral metabolizable energy (ME) for growth (Le Dividich et al.,
1994). In contrast, the efficiency of milk ME for growth is lower (0.73; Marion and
Le Dividich, 1999), suggesting a possible higher contribution of the thermal effect of milk to
the extra heat produced in the cold.

3.3. Shivering or non-shivering thermogenesis

The nature of heat production mechanisms in cold-exposed newborn mammals has long been
an open question. It has been known for years that the newborn pig shivers vigorously from
Energy metabolism in early development of the pig 357

birth (Mount, 1968) and is at variance with most other newborn mammals in that it contains
very little adipose tissue of any type at birth, and appears to have no brown fat. The unlikely
presence of brown fat and existence of NST are suggested by (i) the absence of calorigenic
response to the injection of noradrenaline during the first week of life (LeBlanc and Mount,
1968; Brück et al., 1969) and (ii) the immunoblotting studies of Trayhurn et al. (1989) showing
that piglets do not express UCP1 in various tissues. However, some doubts subsist, because
small quantities of adipose tissue resembling brown adipose tissue have been detected in
3-month-old pigs (Dauncey et al., 1981). In skeletal muscle, the situation is even more intriguing
because loose-coupled mitochondria have been detected in skeletal muscle of 2-month-old
cold-adapted pigs (Herpin and Barré, 1989). Interestingly, UCP homologues have been recently
identified in numerous mammalian tissues including skeletal muscle (review in Ricquier and
Bouillaud, 2000), which reinforces interest in the molecular mechanisms underlying cellular
thermogenesis.
These proteins (UCP) allow the dissipation of part of the proton electrochemical gradient
generated by the electron transfer chain across the inner mitochondrial membrane and can
thus increase heat production by uncoupling respiration from ATP synthesis (Boss et al.,
1998b). This role is probably significant since the proton leak, in part sustained by UCPs,
contributes up to 50% to skeletal muscle basal respiration rate and nearly 30% to standard
metabolic rate in the rat (Brand et al., 1994). UCP1 is exclusively expressed in brown adipose
tissue and plays a vital role in protection against cold. The UCP homologues, UCP2 and
UCP3, are respectively 59% and 57% identical to UCP1 in their amino acid sequences. UCP2
is widely expressed in human and rat tissues whereas UCP3, which shares 73% identity with
UCP2, is highly expressed in muscle (Fleury et al., 1997). Both proteins are able to uncouple
respiration when they are recombinantly expressed in yeast (Fleury et al., 1997; Gong et al.,
1997) and in myoblasts (Boss et al., 1998a).

3.3.1. Search for uncoupling proteins in piglet skeletal muscle

Despite the absence of UCP1 in pigs, UCP2 and UCP3 are expressed in skeletal muscle
(Damon et al., 2000a). However, in agreement with most recent studies, cold stress has no
effect on UCP3 expression in pig skeletal muscle (Damon, unpublished observation).
Results that do not support a classic thermogenic uncoupling role for UCP2 or UCP3 include
the absence of an abnormality in thermoregulation in UCP3 or UCP2 knockout mice
(Arsenijevic et al., 2000; Gong et al., 2000) and a rise in their skeletal muscle transcript
levels during fasting, food restriction or chronic exercise. Thus, the physiological impor-
tance of UCP2 and UCP3 in regulatory cold-induced thermogenesis is still a matter of
debate in the pig, as in other species, and the presence of UCP2 and UCP3 in pig muscle is
not conclusive of the existence of NST. As far as we know, the most recent investigations
suggest that UCP3 has the potential to act as a molecular determinant in the regulation of
resting metabolic rate by 3,5,3′-triiodothyronine (T3) (De Lange et al., 2001), and to play a
role in the export of fatty acids from mitochondria (Himms-Hagen and Harper, 2001). In
piglet muscle, the up-regulation of UCP3 mRNA by acute T3 treatment is less marked than
in rats (Damon et al., 2000a).

3.3.2. Search for the existence of non-shivering thermogenesis in the piglet

To establish the possible existence of non-shivering thermogenesis, it is necessary to measure


simultaneously the magnitude of shivering and the level of heat production at temperatures
358 P. Herpin et al.

Fig. 2. Changes in heat production (  ) and shivering intensity (integrated EMG,  ) with ambient temperature
at 2 h and 5 days of age. Adapted from Berthon et al. (1994).

ranging from thermoneutrality to cold (Barré et al., 1985). As illustrated in fig. 2, this allows
comparison of the lower critical temperature (LCT) with the temperature threshold for shiv-
ering (STT). The lack of delay between the cold-induced increase in heat production (LCT)
and the onset of shivering (STT), and the strict linearity of the relationship between meta-
bolic rate and shivering intensity within the range of ambient temperature studied (fig. 2),
confirm the absence of NST and the main role of shivering in neonatal thermogenesis
(Berthon et al., 1994). This result is observed at 2 h as well as at 5 days of life and is not
modified after 2 days of cold exposure (Berthon et al., 1994). One interesting point in fig. 2
is the reduction of shivering intensity between 2 h and 5 days of age. In the absence of reg-
ulatory non-shivering thermogenesis, this reduction of shivering intensity with age for a
given metabolic rate suggests that the thermogenic efficiency of shivering, i.e. the heat
power of shivering, increases with postnatal age in the cold. In other words, more heat is
produced per unit increase in the electrical activity of shivering at 5 days than at 2 h of life
(Berthon et al., 1994), which should definitely contribute to the improvement of piglet
thermostability after birth.
In pigs aged 5–6 weeks and exposed to cold for 3 weeks, shivering is progressively
replaced by NST. In the absence of brown fat, skeletal muscles are potential candidates for
such adaptations, especially when one considers that skeletal muscle mass contributes sub-
stantially to total body weight. Two mechanisms that are likely to contribute to heat production
have been identified in skeletal muscles from cold-acclimated weaned piglets. First, it is now
well accepted that cellular O2 consumption and ion pumping are intimately linked (Gregg and
Milligan, 1982), and it has been shown that Na+/K+-ATPase-dependent respiration represents
about 20% of whole-muscle O2 consumption. The activity of this pump is increased by 75%
in the cold and this increase accounts for 70% of the total increase in muscle respiration
(Herpin et al., 1987a; Harrison et al., 1994), which is probably associated with a similar
increase in heat production. Second, heat can also be directly produced at the level of ATP
synthesis. Indeed, a loose coupling between oxidation and phosphorylation is observed in
subsarcolemmal mitochondria from rhomboideus muscle (Herpin et al., 1987b). This muscle
is located between the shoulders, i.e. at the same place that brown fat, with its mitochondrial
uncoupling, occurs in cold-acclimated rats. However, the biological significance and impor-
tance of those mechanisms is difficult to assess and the existence of non-shivering thermogenesis
has not been confirmed in vivo.
Energy metabolism in early development of the pig 359

Fig. 3. Overview of the structures and metabolic pathways involved in muscle contraction and heat production
during shivering.

3.4. Mechanisms involved in the postnatal improvement of shivering efficiency

All the components of skeletal muscle energy metabolism can be potentially involved in the
postnatal increase in the capacity to produce heat during shivering. Indeed, to produce heat,
muscle fibres need to exhibit both an optimal coupling between excitation and contraction, an
optimal power of contraction, an adequate supply of substrates and oxygen, and finally an
efficient ATP synthesis (fig. 3).

3.4.1. Mechanisms and structures involved in muscle contraction

During shivering, repetitive contractions of muscle fibres produce heat by myosin cross-bridge
cycling, Ca2+ cycling and Na+/K+ transport. Obviously, the power of contraction of muscle
fibres is directly related to myofibril mass and, unfortunately, piglet fibres definitely lack
myofibrils at birth (Bradley et al., 1980; Handel and Stickland, 1987; Herpin et al., 2002).
Recent results show that myofibril volume density is quite low in longissimus lumborum
(LL, 32%) and rhomboideus (RH, 40%) muscles at birth, and increases markedly within 5 days
(fig. 4). The high rate of muscle protein synthesis in the newborn is reported to be restricted
entirely to the myofibrillar protein compartment (Fiorotto et al., 2000). This marked postnatal
increase in myofibril mass is probably one of the key events in the development of muscle func-
tion and should contribute to the enhancement of muscle contraction potential during shivering.
With regards to muscle fibre types, endurance and high aerobic capacities are important
features for thermogenesis because of the continuous nature of shivering. Structures enhanc-
ing endurance are mitochondria and their fuel source, lipids, and they are mostly present in
slow-oxidative (type I) and fast-oxidoglycolytic (type IIa) fibres. Moreover, concerning the
economy of contraction, fast muscles consume more energy than slow when working for the
same time against the same load (Crow and Kushmerick, 1982) and fast-cycling muscles have
360 P. Herpin et al.

Fig. 4. (A) Fibre (Vfibre) and (B) myofibril (Vmyofibril) volume density in longissimus lumborum (LL) and
rhomboideus (RH) muscles of newborn (open bars) and 5-day-old piglets exposed to thermoneutral (striped
bars) or cold (solid bars) conditions. Means sharing a common superscript did not differ significantly
(P > 0.05). The effect of muscle type on Vfibre is significant (P < 0.05). Adapted from Herpin et al. (2002).

relatively uneconomical force development owing to their high cross-bridge cycling rate and
short contraction time (Alpert and Mulieri, 1986). As a whole, fast-contracting high-
endurance fibers (type IIa fibres) may produce more heat per unit time than other fibres when
used in shivering. Interestingly, the proportion of type IIa fibres is very high in piglet muscles
at birth (Lefaucheur and Vigneron, 1986). In addition, transitory expression of α-cardiac
myosin heavy chain (MHC), with contractile properties intermediate between type I and
type IIa MHC, has been recently demonstrated in muscle fibres of piglets (Lefaucheur et al.,
1997). Moreover, the expression of this myosin isoform is markedly increased within the first
5 days of life and is stimulated by chronic cold exposure (Lefaucheur et al., 2002). Although
our knowledge of α-cardiac MHC properties is very limited, its marked up-regulation in cold-
exposed shivering piglets points out its potential role in the enhancement of shivering
efficiency after birth.
Also, the excitation–contraction coupling apparatus appears to be immature at birth in pigs,
as already shown in birds (Eppley and Russell, 1995). Triads correspond to the junctional
association of transverse tubules with sarcoplasmic reticulum terminal cisternae in mature
skeletal muscle (fig. 3) and thereby play a crucial role in calcium release in excitation–
contraction coupling (Flucher, 1992). Electron microscopic examination showed that they
proliferate rapidly in LL and RH muscles of cold-exposed piglets after birth (+70% to +90%
within 5 days) (fig. 5), suggesting that the efficiency of excitation–contraction coupling

Fig. 5. Number of triad profiles per unit fibre area (Ntriad) in longissimus lumborum (LL) and rhomboideus
(RH) muscles of newborn (open bars) and five-day old piglets exposed to thermoneutral (striped bars) or cold
(solid bars) conditions. Means sharing a common superscript did not differ significantly (P > 0.05). Adapted
from Herpin et al. (2002).
Energy metabolism in early development of the pig 361

increased in the cold (Herpin et al., 2002). In other words, for a given electrical stimulation
more of the sarcoplasmic reticulum is likely activated during contraction and calcium release
and cycling are higher, thereby promoting an increase in ATP hydrolysis and heat production
through cross-bridge cycling, calcium reuptake and myofibril contraction (Block, 1994).
Therefore, triad proliferation is probably of utmost importance in the adaptive and improved
response of skeletal muscle to sustained shivering during the early neonatal period.

3.4.2. Muscle blood supply

Blood flow through muscles is highly adaptable and represents an important determinant of
oxygen and nutrient supply and consumption in all species (Hoppeler et al., 1981). It should
be sufficient to supply oxygen and nutrient during shivering activity and to support heat pro-
duction mechanisms. When comparing muscle blood flow in 1- and 5-day-old piglets, it
appears clearly that muscle cardiovascular adjustments in the cold are limited in the newborn
pig (table 2).
Measurement of blood flow in a representative muscle compartment, i.e. the hindquarters,
in conscious piglets shows that muscle blood flow increases with age and short-term cold
exposure. However, changes in blood flow in response to a similar cold challenge were 3
times higher in 5-day-old (+65%) than in 1-day-old (+25%) piglets, suggesting that blood
supply to the shivering muscle was considerably improved with age (Lossec et al., 1998b).
Measurement of blood flow in individual muscles using coloured microspheres confirms
these results (Lossec et al., 1999). Interestingly, a preferential redistribution of cardiac output
towards skeletal muscle was only observed at 5 days of life, at the expense of the small intes-
tine, the liver and the skin; this cardiovascular response was more pronounced in the most
oxidative skeletal muscles studied (RH vs. LL). This should favour, and probably potentiate,
the efficiency of shivering. Indeed, a redistribution of cardiac output to the most thermogenic

Table 2
Cardiac output and its fractional distribution Fco to selected tissues or whole organs in 1- and
5-day-old pigs exposed to thermoneutral (TN) or cold (C) environment

1-day-old 5-day-old

TN C TN C

Cardiac output, mL/min/kg 436 535*,a 448 550*


Fco for 10 g tissue (%)
Longissimus thoracis 0.42 0.55* 0.42 0.61*
Rhomboideus 0.47 0.64* 0.49 0.87*
Subcutaneous adipose tissue 0.25 0.22 0.21 0.22
Skin 0.25 0.25 0.35 0.27*
Fco to whole organs, %
Heart 1.87 2.45* 3.24 4.61*
Liver 3.97 3.75 6.44 5.18*
Small intestine 13.6 14.0 13.7 10.8*
Brain 2.72 2.86 3.27 3.12
Adrenals 0.14 0.14 0.13 0.15
Thyroid 0.08 0.09 0.08 0.11*
a given age. *denotes significant effect (P < 0.05) of cold.
a At

Adapted from Lossec et al. (1999).


362 P. Herpin et al.

organs has already been reported in rats (Foster and Frydman, 1979), pigs (Mayfield et al.,
1986) and ducklings (Duchamp and Barré, 1993). Additional investigations using electron
microscopy have shown no increase in skeletal muscle capillary bed with age or cold expo-
sure (Herpin et al., 2002). This suggests that the observed postnatal improvement of
cardiovascular adjustments required for shivering thermogenesis may be accommodated by
existing capillaries. Therefore, present changes are more likely related to the postnatal matu-
ration of the nervous and hormonal regulation of muscle blood flow.

3.4.3. Activity and ultrastructure of muscle mitochondria

The plasticity of mitochondrial density and activity according to age or cold acclimation is
well documented. During shivering, optimal supply of ATP via oxidative phosphorylation is
necessary when glycogen stores are exhausted. In contrast to liver mitochondria (Mersmann
et al., 1972), muscle mitochondria are functional from birth in pigs and are changing prima-
rily quantitatively during the first 5 days of life (Berthon et al., 1996a; Schmidt and Herpin,
1997). The oxidative potential of pig muscle increases gradually after birth, but no consistent
changes in mitochondrial respiration, respiratory control and phosphorus:oxygen ratio are
evidenced during this period and after short-term cold exposure (Schmidt and Herpin, 1997).
However, biochemical characteristics of intermyofibrillar (IMF) and subsarcolemmal (SS)
mitochondria differ from birth. The higher respiration rate and higher respiratory control ratio
shown by IMF compared with those shown by SS mitochondria are principally due to the
higher activities involved in substrate oxidation because there is no difference in the proton
leak between both populations (Lombardi et al., 2000).
Thus, the actual event responsible for the postnatal increase in skeletal muscle oxidative
potential is the enhancement of mitochondrial mass (Schmidt and Herpin, 1997), as already
reported in various tissues and species during the neonatal period (Mersmann et al., 1972;
Eppley and Russell, 1995). Between birth and 5 days of life, mitochondrial mass increased
by 49% in glycolytic LL muscle and by 93% in oxidative RH muscle. In LL muscle this
increase was only supported by the proliferation of IMF mitochondria, whereas both types
of mitochondria (IMF and SS) proliferate in RH muscle. The mechanisms underlying these
changes have been elucidated by electron microscopic examination. Within 5 days (fig. 6),
there is an increase in both the number of mitochondria and the surface of the inner mem-
brane and cristae of each mitochondrion (Herpin et al., 2002). Indeed, the number of
respiratory chain and F1-ATPase units is directly related to this parameter (Hoppeler, 1986).
Interestingly, this postnatal change in the surface of the inner membrane is more marked
in RH than in LL muscle, and is further enhanced when piglets are exposed to cold for
5 days. As a whole, this should contribute to the enhancement of muscle endurance during
contractile activity associated with shivering, and to the postnatal acquisition of muscle
metabolic type.

4. SOURCES OF ENERGY
The requirement for energy in the newborn pig is met by body energy reserves, colostrum and
milk. During the neonatal period the protein accretion is very high, and the potential for pro-
tein deposition is probably beyond that allowed by milk intake (Le Dividich and Sève, 2001).
The rate of amino acid catabolism is very low during this period and is not enhanced in cold
conditions (Herpin et al., 1992). The contribution of energy derived from amino acid catabo-
lism is therefore of marginal importance and will not be discussed.
Energy metabolism in early development of the pig 363

Fig. 6. Stereological parameters of mitochondria in longissimus lumborum (LL) and rhomboideus (RH)
muscles of newborn (open bars) and 5-day-old piglets exposed to thermoneutral (striped bars) or cold (solid
bars) conditions.(A) Number of mitochondria per unit fiber area (Nmitochondria); (B) Mitochondria volume
density (Vmitochondria); (C) Surface of outer mitochondrial membranes per unit tissue volume
(Somm/Vtissue); (D) Surface of inner mitochondrial membrane and cristae per unit mitochondrial volume
(Si+c/Vmitochondria). Means sharing a common superscript did not differ significantly (P > 0.05). The effect
of muscle type on Si+c/Vmitochondria is significant (P < 0.05). Adpated from Herpin et al. (2002).

At birth, body energy stores are present as glycogen and fat (fig. 7). Because FFA are
poorly transferred across the swine placenta (Thulin et al., 1989), the amount of fat reserves
in the newborn pig is low, ranging from 15 to 20 g/kg body weight (BW). Most (45%) of this
stored fat is structural fat and is not available for mobilization. Total body glycogen stores
range from 30–38 g/kg BW with the major part (~90%) being located in muscle. From an

Fig. 7. Available energy stores at birth and cumulative available energy derived from ingested colostrums.
Adapted from Mellor and Cockburn (1986) and Le Dividich et al. (1997).
364 P. Herpin et al.

energy point of view, glycogen represents >90% of the available stored energy. Nevertheless,
available energy derived from body stores is low, accounting for <10% of that of the newborn
infant (Mellor and Cockburn, 1986).
Soon after birth the pig is fed at intervals with colostrum for 24–36 h and thereafter milk.
The first suckling occurs in the 20–30 min following birth and colostrum intake can be very
high immediately after birth since the first three sucklings account for ~25% of the total
colostrum ingested during the first day (Fraser and Rushen, 1992; Le Dividich et al., 1997).
Although highly variable, total consumption of colostrum in the first day of life is in the range
of 310–340 g/kg BW and may be as high as 460 g/kg BW (Le Dividich et al., 1997). Fat
accounts for 35–50% and 55–65% of total energy of colostrum and milk, respectively, and
represents the main source of energy. Colostrum and milk fats are highly digestible (~100%).
They are composed of long-chain fatty acids (>C14), but are devoid of medium-chain fatty
acid (MCFA). Both the fatty acid profile and amount of colostrum and milk fat can be manip-
ulated by the source of dietary fat provided to the dam in late gestation and throughout
lactation (Averette et al., 1999), but there is no evidence for mammary MCFA transfer
(Newcomb et al., 1991). Moreover, lactose is the predominant carbohydrate in colostrum and
milk, but its content is lower in colostrum than in milk (3.1–3.9% vs. 4.8–5.5%).
This abrupt shift in the source of energy substrates from mainly glucose to colostral and
milk fat implies that the piglet is rapidly capable of (i) oxidizing fat to provide energy for heat
production, (ii) depositing fat for thermal insulation and energy through its mobilization at
weaning, and (iii) providing glucose to its glucose-dependent tissues.

5. EFFECTS OF THE SHIFT IN THE ENERGY SOURCE ON LIPID


OXIDATION, LIPOGENESIS AND GLUCONEOGENESIS
5.1. Lipid oxidation

During the first postnatal hours, piglets rely almost entirely on carbohydrate to meet their
thermoregulatory needs (Mount, 1968). However, in usual environmental conditions at birth,
75% of liver glycogen and 41% of muscle glycogen is mobilized by 12 h postpartum (Elliot
and Lodge, 1977). Cold exposure hastens the depletion in both tissues (Herpin et al., 1992),
and increases the rate of glucose turnover (Lossec et al., 1998a) and of peripheral glucose
uptake (Duée et al., 1988; Lossec et al., 1998a). However, colostrum intake increases avail-
ability of lipids. The ensuing increase in plasma NEFA (Le Dividich et al., 1991b) and glycerol
(Bengtsson et al., 1969) secondary to the increased activity of the adipose tissue hormone-
sensitive lipase (Horn et al., 1973; Steffen et al., 1978) is associated with a progressive
decline in respiratory quotient during the first postnatal day in both thermoneutral and cold
environments (fig. 8), providing evidence for an early involvement of lipids as an energy
source (Noblet and Le Dividich, 1981; Berthon et al., 1993). At 48 h of age, fatty acids
account for ~60% of the energy metabolism, increasing to 90% in the 7-day-old pig fed at
maintenance (Marion and Le Dividich, 1999). Biochemical adjustments associated with the
improved ability of the piglet to oxidize fat in the skeletal muscle include:
1. An increase in muscle lipid content within the first postnatal days. For example, muscle
lipid content nearly doubled within 5 days in LL and RH muscle, in agreement with the
increase in the number of lipid droplets per unit tissue area (Herpin et al., 2002). At birth
these lipid droplets are scarce, but at 5 days they are wedged between the myofibrils and
the IMF mitochondria, a position that is ideal for optimizing the provision of energy for
oxidative metabolism and sustained shivering.
Energy metabolism in early development of the pig 365

Fig. 8. Respiratory quotient of the neonatal pig in relation to age and ambient temperature. Adapted from
Berthon et al. (1993) and Schmidt and Herpin (1997).

2. An increase in NEFA uptake by the shivering muscle at 5 days of life whereas this
uptake is negligible in 1-day-old piglets (Lossec et al., 1998b).
3. An increase in mitochondrial (+100%) and peroxisomal (+160%) β-oxidation potential
in LL and RH muscle homogenates between birth and 5 days of life (Herpin, unpublished
observation). Interestingly, a similar increase in mitochondrial and peroxisomal β-oxidation
potential has been previously reported in liver, kidney and heart of young pigs
(Yu et al., 1997).
Deeper investigations into the molecular and biochemical regulation of fatty acid oxidation
in piglet skeletal muscle showed that oleic, linoleic and palmitic acids are readily oxidized
from birth by isolated skeletal muscle mitochondria (Schmidt and Herpin, 1998). MCFA
(C8–C10), which are now being introduced in colostrum and milk substitutes, are readily oxi-
dized by the liver (for review, see Odle, 1997). However, in vivo studies in respiration
chambers (Léon et al., 1998) provide evidence that substitution of MCFA for long-chain fatty
acids in colostrum does not improve the energy status of the newborn, even in cold conditions
(fig. 9). It is suggested that MCFA are poorly oxidized by skeletal muscle as indicated by
in vitro studies using isolated muscle mitochondria (Schmidt and Herpin, 1997). Surprisingly,
the mitochondrial potential is not increased with age, which suggests that the enhancement of
fatty acid oxidation potential with age is mostly supported by the above-mentioned postnatal
proliferation of muscle mitochondria.
Complex changes in the expression and activity of carnitine palmitoyltransferase I (CPT I),
which is the limiting enzyme of fatty acid β-oxidation, have also been observed (Odle et al.,
1995; McGarry and Brown, 1997). In piglets, CPT I activity increases postnatally in SS
muscle mitochondria and is modulated by malonyl-CoA in IMF mitochondria (Schmidt and
Herpin, 1998). Indeed, between birth and 5 days of life, both the sensitivity of CPT I to mal-
onyl-CoA inhibition and the tissue level of malonyl-CoA decreased, which could partly
relieve CPT I inhibition and enhances fatty acid utilization. Further, during cold stress, the
decrease in the tissue levels of malonyl-CoA is even more marked in the most oxidative
muscle. A surprising result is the difference in CPT I sensitivity to malonyl-CoA between
piglets and rats: in piglets, sensitivity to malonyl-CoA is much lower in muscle than in the
366 P. Herpin et al.

Fig. 9. Effect of substitution of medium- (MCFA) for long- (LCFA) chain fatty acids in colostrum on heat
production (a) and respiratory quotient (b) in the newborn pig in relation to environmental temperature. Data
adjusted to a common ME intake of 960 kJ/kg BW per 26 h. Adapted from Léon et al. (1998).

liver (Schmidt and Herpin, 1998; Nicot et al., 2001) whereas the opposite has been obtained
in rats. At the molecular level, two isoforms with tissue-specific expression and sensitivity to
malonyl-CoA inhibition are usually described.
Recently, partial liver (CPT1-L) and muscle (CPT1-M) cDNA sequences have been suc-
cessfully cloned and the co-expression of these two isoforms in skeletal muscle of neonatal
piglets has been demonstrated (Damon et al., 2000b). The expression of CPT1-L in pig skele-
tal muscle could provide a partial answer to the difference of sensitivity to malonyl-CoA
between rat and pig muscles (Schmidt and Herpin, 1998). However, co-expression of both
isoforms should result in an intermediate and not a reverse sensitivity to malonyl-CoA inhi-
bition. Recent data support another seductive hypothesis. In fact, in yeast expressing pig
CPT1-L, kinetic characteristics (Km’s for carnitine and palmitoyl-CoA) were similar to those
of human and rat CPT1-L whereas sensitivity to malonyl-CoA inhibition was found to be
closer to that of rat and human CPT1-M isoforms (Nicot et al., 2001). It then appears that pig
CPT1-L possesses specific biochemical properties despite its high degree of homology with
CPT1-L from other mammals. Thus pig CPT1-M could also behave as CPT1-L of other mam-
mals in terms of malonyl-CoA sensitivity. It has been speculated that CPT1-L was prone to
hormonal regulation whereas CPT1-M was more regulated by nutritional factors (Cook et al.,
2001). Thus co-expression of both isoforms in pig skeletal muscle could allow a fine tuning
of lipid utilization.

5.2. Lipogenesis

The neonatal pig has a remarkable capacity to deposit large amounts of fat soon after birth.
Depending on the colostrum fat content, carcass fat content increases by 25–100% during the
first day of life (Le Dividich et al., 1997). During the suckling phase, fat accretion occurs at
a mean rate of 30–35 g/day, depending mainly on the amount of ingested milk (Marion and
Le Dividich, 1999) and on the milk fat content (Jones et al., 1999). De novo lipogenesis is
marginal, albeit being limited not by enzyme activities or insulin-regulated glucose trans-
porter (GLUT 4), but by the substrate availability (Gerfault et al., 2000). Colostrum and milk
are the major routes for lipid acquisition (Sarkar et al., 1985). Adipose tissue lipoprotein lipase,
Energy metabolism in early development of the pig 367

an enzyme playing a key role in fat storage, contributes to the ability of the suckling pig to
deposit large amounts of fat. Its activity, already high at birth, is increased 3–4-fold in the early
neonatal period (Steffen et al., 1978; Le Dividich et al., 1997). It is suggested that sow’s milk is
designed to promote fat accretion in the young pig. Stored fat is mostly subcutaneous fat that
provides thermal insulation to the young pig and energy through its mobilization during the
period of low feed intake following weaning (Le Dividich and Sève, 2001).

5.3. Gluconeogenesis

The glucose requirement of the neonatal pig is very high, reaching ~15 g/kg BW per day
(Flecknell et al., 1980; Pégorier et al., 1984). It is 50% higher than in lambs and human
infants. Moreover, the glucose requirement is enhanced ~30% in the cold (Duée et al., 1988;
Lossec et al., 1998a). Glucose requirements are met by (i) liver glycogenolysis, since only
liver glycogen is able to release glucose into the blood, (ii) colostrum and milk and (iii) glu-
coneogenesis. Provision of glucose by the first two sources meets 50–60% of the
requirements during the first day of life, which underlines the importance of the gluco-
neogenic pathway in the glucose homeostasis of the piglet.
Gluconeogenesis is the process by which glucose is synthesized from various precursors.
The liver is the main site of gluconeogenesis. The developmental pattern of hepatic gluco-
neogenesis has been the subject of several extensive reviews (Girard, 1986; Girard et al.,
1992) and will not be discussed in detail in this chapter. In brief, key enzymes involved in the
pathway, i.e. pyruvate carboxylase, phosphoenolpyruvate carboxykinase, fructose-1,6-
diphosphatase and glucose-6-phosphatase (G6Pase), have a substantial activity (35-105% of
adult values) at birth. In both fed and unfed pigs, enzyme activity increases markedly during
the first postnatal day. Also, the insulin:glucagon molar ratio decreases after birth in both fed
and unfed piglets, thus providing an appropriate environment for an active gluconeogenesis.
However, the level of plasma NEFA is much higher in the fed pigs. In fact, colostrum ingestion
is essential to sustain a high rate of hepatic gluconeogenesis. Based on the amount of glucose
available from lactose digestion (glucose + galactose), colostrum provides at least 40–45% of
the glucose requirement and ~80–90% on the assumption that all galactose is converted into
glucose by the G6Pase. However, piglets fasted from birth (Goodwin, 1957) or fed a low-fat
colostrum (Herpin et al., 1992) are unable to sustain normal glycaemia. It is suggested that fatty
acid oxidation plays a major role in glucose homeostasis (Girard, 1986) through the supply of
ATP and co-factors (NADH and acetyl-CoA) for catalysing key reactions. This is convincingly
demonstrated by the in vivo and in vitro studies of Pégorier et al. (1985) and Duée et al. (1985).
In contrast, Lepine et al. (1991) failed to find any stimulatory effects of fatty acid oxidation on
the rate of glucose production by isolated hepatocytes from piglets.

6. REGULATION OF ENERGY METABOLISM DURING


EARLY DEVELOPMENT
6.1. The thyroid axis

Thyroid hormones (TH) are known to play a major role in the regulation of metabolic adap-
tations and growth. They exert their effects primarily through interactions with nuclear TH
receptors (TR) which occur as a series of isoforms controlling the transcription of thyroid
hormone-responsive genes (Lazar, 1993; Wrutniak-Cabello et al., 2001). The ontogenic
368 P. Herpin et al.

profile of the thyroid system suggests that TH metabolism is fully developed at birth. Plasma
concentrations of both total and free TH, thyroid gland weights and hepatic 5′-deiodinase
activity all increase during late gestation (Berthon et al., 1993). It is relevant to notice that
receptors are already present in skeletal muscle, but not in the liver, at 80 days of gestation,
suggesting that porcine muscle can potentially respond to TH much earlier in development
than the liver (Duchamp et al., 1994). During the first 6 h after birth, there is a surge in T3,
free T3 and T4 plasma concentrations and, apart from a transient decline at 12 h, TH concen-
trations remain elevated during the first 2 days and then decline slightly over the next 2 weeks
(Slebodzinski et al., 1981; Berthon et al., 1993, 1996b).
The finding that the postnatal surge in plasma TH levels precedes the physiological rise in
heat production in the newborn suggests a close relationship between perinatal thyroid status
and neonatal thermogenic capacity (Berthon et al., 1993). This is exemplified by the findings
that (i) hypothyroidism at birth is associated with depressed thermoregulatory capabilities of
the newborn (Berthon et al., 1993) and (ii) a single injection of T4 induces an increase in
metabolic rate (Slebodzinski, 1979). TH control the oxidative capacities in the newborn
through a short-term regulation of mitochondial respiration (Herpin et al., 1996) and a long-
term regulation of mitochondriogenesis (Mutvei et al., 1989). The thyroid axis is also
regulated by nutrition and TH actions might be complementary to catecholamine actions
during cold-induced thermogenesis. Cold-exposed newborn pigs fed a limited amount of milk
exhibit high catecholamines but low T3 levels whereas the opposite is observed in piglets fed
a high milk intake (Herpin et al., 1995; Berthon et al., 1996b). These adaptations are assumed
to optimize the utilization of either body stores (low intake) or exogenous substrates (high
intake). The marked effects of food intake on the thyroid axis are also observed during the
whole suckling period: a low intake reduces thyroid gland activity, circulating TH concentra-
tions and nuclear TR abundance in muscle (Dauncey, 1990; Morovat and Dauncey, 1995).

6.2. HPA axis

Circulating levels of glucorticoids and catecholamines are very high at the time of birth and
dramatically decrease thereafter (Kaciuba-Uscilko, 1972; Randall, 1983). Cortisol and cate-
cholamines are potent stimulators of catabolism and one can speculate that these high levels
induce mobilization of glycogen stores after birth. However, response of catechalomine to
cold exposure during the first 5 days of life is variable, with both no change (Lossec, 1998)
and a marked increase (Duée et al., 1988; Le Dividich et al., 1991a) being reported.
Moreover, the response to cold is found to be impaired in moderately hypothermic piglets
(Mayfield et al., 1989) or, as mentioned above, to be dependent on the level of milk intake
(Herpin et al., 1995). A lipolytic response is only detected at 2–4 days of age (Curtis and Rogler,
1970; Persson et al., 1971). Neither norepinephrine nor epinephrine administration elicits a
thermogenic response in the neonatal pig (LeBlanc and Mount, 1968; Persson et al., 1971).
Clearly, these observations indicate that the actual role of catecholamines in the neonatal
thermogenesis requires further investigation.

6.3. Somatotropic axis

Even though the regulation of energy metabolism by the somatotropic axis is well docu-
mented in the growing pig (Louveau and Bonneau, 2001), there is no evidence from the
literature that the somatotropic axis contributes to the cold-induced thermogenesis. However,
because of a high potential for protein synthesis and growth of the neonatal pig (Le Dividich
Energy metabolism in early development of the pig 369

and Sève, 2001), one might expect that the energy metabolism is to some extent regulated by
the somatotropic axis. Plasma GH concentrations are very high at birth and decrease sharply
during the next 2–3 days (Scanes et al., 1987; Carroll et al., 1998). Although the significance
of these high levels of plasma GH is not completely understood, GH could contribute to the
maintenance of protein accretion in the newborn pig, even in negative energy balance (Herpin
et al., 1992). Plasma IGF-I concentrations increase significantly during the first 3 weeks after
birth (Lee et al., 1991, 1993; Louveau et al., 1996). After 24 h of feeding, IGFBP profile
changes with the abundance of plasma IGFBP-3 predominating (Lee et al., 1991). Changes
in GH and IGF-I receptor levels are also observed during this period, with GH receptor
increasing over the first 10 days of life in liver and IGF-I receptor decreasing in skeletal
muscle and other tissues (Breier et al., 1989; Lee et al., 1993; Louveau et al., 1996;
Schnoebelen-Combes et al., 1996). These profiles are modulated by thyroid status (Duchamp
et al., 1996). In addition, the somatotropic axis appears to be functional and responsive to GH
administration in neonatal pigs, although the responsiveness is reduced compared to older
pigs (Harrell et al., 1999). The administration of GH at a dose that is commonly used in older
pigs has little or no effect on growth rate or plasma IGF-I or IGFBP-3 (Harrell et al., 1999;
Dunshea et al., 2001). Perhaps the lack of response in the growth rate is not surprising owing
to the already high rate of protein synthesis.
Changes in nutritional status during the neonatal period are associated with several changes
in the GH–IGF-I axis. Both moderate and severe feed restriction (Dauncey et al., 1994;
Louveau and Le Dividich, 2002) in the suckling period decrease plasma IGF-I and IGFBP-3
levels. These data indicate that circulating IGF-I is directly related to energy intake in neonatal
pigs as observed in older animals. Even though the regulation of receptors may represent an
important mechanism of control within the GH–IGF-I axis, the few available data indicate
that the regulation of IGF-I and GH receptors is tissue-specific and dependent on the type of
undernutrition during the suckling period (Louveau and Le Dividich, 2002).

7. CONCLUDING COMMENTS AND FUTURE PERSPECTIVES


This chapter provides new insights on the development of the energy metabolism in a species
devoid of brown fat. Key factors involved in the poor abilities of the newborn pig to withstand
cold stress include mainly the relative immaturity of the newborn pig and the availability of
energy substrates. Improvement of its thermogenic capacities within the first postnatal days
parallels maturation of the skeletal muscle metabolism and function and of the cellular
machinery.
In the future, in the light of improving survival, it should be relevant to select piglets on
physiological traits related to maturity. This is convincingly attested by the first findings
(Leenhouvers, 2001) that selection of pigs with different genetic merit for survival leads to
piglets with a higher maturity at birth. Attempts made to improve the energy available at birth
resulted only in moderate increase in energy stores at birth. However, effects of sow nutrition
during pregnancy on fetal muscle development during the critical stages of fetal development
warrant future investigation. In addition, we suggest that more research should be focused on
factors initiating and controlling quantity and quality of colostrum and milk produced by the
sow. However, during the past decades, selection for lean tissue growth has led to less mature
pigs at birth (Herpin et al., 1993). Selection of sows for higher litter size has resulted in prob-
lems of increased intrapartum deaths, proportion of weak piglets and competition at the udder
(Quiniou et al., 2002). Therefore, our efforts will be in vain if the survival of the piglet
continues to be challenged unwisely by the pig industry.
370 P. Herpin et al.

REFERENCES
Alpert, N.R., Mulieri, L.A., 1986. Functional consequences of altered cardiac myosin isoforms. Med. Sci.
Sport Exerc. 18, 309–313.
Arsenijevic, D., Onuma, H., Pecqueur, C., Raimbault, S., Manning, B.S., Miroux, B., Couplan, E., Alves-
Guerra, M.C., Goubern, M., Surwit, R., Bouillaud, F., Richard, D., Collins, S., Ricquier, D., 2000.
Disruption of the uncoupling protein-2 gene in mice reveals a role in immunity and reactive oxygen
species production. Nat. Genet. 26, 435–439.
Averette, L.A., Odle, J., Monaco, M.H., Donavan, S.M., 1999. Dietary fat during pregnancy and lacta-
tion increases milk fat and insulin-like growth factor 1 concentrations and improves neonatal growth
rates in swine. J. Nutr. 129, 2123–2129.
Barré, H., Geloen, A., Chatonnet, J., Dittmar, A., Rouanet, J.L., 1985. Potentiated muscular thermogen-
esis in cold-acclimated Muscovy duckling. Amer. J. Physiol. 249, R533–R538.
Bengsston, G., Gentz, J., Hallarainen, J., Hellstrom, R., Persson, B., 1969. Plasma levels of FFA, glycerol,
β-hydroybutyrate and blood glucose during the postnatal development of the pig. J. Nutr. 97, 311–315.
Berthon, D., 1994. Thermogenesis in the neonatal pig. Ph.D. Thesis, Université Claude Bernard, Lyon,
122. pp.
Berthon, D., Herpin, P., Duchamp, C., Dauncey, M.J., Le Dividich, J., 1993. Modification of thermogenic
capacity in neonatal pig by changes in thyroid status during late gestation. J. Dev. Physiol. 19, 253–261.
Berthon, D., Herpin, P., Bertin, R., De Marco, F., Le Dividich, J., 1996a. Metabolic changes associated with
sustained 48-hr shivering thermogenesis in the newborn pig. Comp. Biochem. Physiol. 114, 327–335.
Berthon, D., Herpin, P., Le Dividich, J., 1994. Shivering thermogenesis in the neonatal pig. J. Thermal
Biol. 19, 413–418.
Berthon, D., Herpin, P., Le Dividich, J., Dauncey, M.J., 1996b. Interactive effects of thermal environment
and energy intake on thyroid hormone and metabolism in newborn pigs. Biol. Neonate 69, 51–59.
Block, B.A., 1994. Thermogenesis in muscle. Annu. Rev. Physiol. 56, 535–577.
Boss, O., Muzzin, P., Giacobino, J.P., 1998b. The uncoupling proteins, a review. Eur. J. Endocrinol. 139, 1–9.
Boss, O., Samec, S., Kuhne, F., Bijlenga, P., Assimacopoulos-Jeannet, F., Seydoux, J., Giacobino, J.P.,
Muzzin, P., 1998a. Uncoupling protein-3 expression in rodent skeletal muscle is modulated by food
intake but not by changes in environmental temperature. J. Biol. Chem. 273, 5–8.
Bradley, R., Ward, P.S., Bailey, J., 1980. The ultrastructural morphology of the skeletal muscles of normal
pigs and pigs with splayleg from birth to one week of age. J. Comp. Pathol. 90, 433–446.
Brand, M.D., Chien, L.F., Ainscow, E.K., Rolfe, D.F., Porter, R.K., 1994. The causes and functions of
mitochondrial proton leak. Biochim. Biophys. Acta 1187, 132–139.
Breier, B.H., Gluckman, P.D., Blair, H.T., McCutcheon, S.N., 1989. Somatotrophic receptors in hepatic
tissue of the developing pig. J. Endocrinol. 123, 25–31.
Brück, K., Wünnenberg, W., Zeisberger, E., 1969. Comparison of cold adaptive metabolic modifications
in different species with special reference to the miniature pig. Fed. Proc. 28, 1035–1041.
Carroll, J.A., Veum, T.L., Matteri, R.L., 1998. Endocrine responses to weaning and changes in post-
weaning diet in the young pig. Domest. Anim. Endocrinol. 15, 183–194.
Cook, G.A., Edwards, T.L., Jansen, M.S., Bahouth, S.W., Wilcox, H.G., Park, E.A., 2001. Differential
regulation of carnitine palmitoyltransferase-I gene isoforms (CPT-I alpha and CPT-I beta) in the rat
heart. J. Mol. Cell. Cardiol. 33, 317–329.
Crow, M.T., Kushmerick, M.J., 1982. Chemical energy of slow- and fast-twitch muscles of the mouse.
J. Gen. Physiol. 79, 147–166.
Curtis, S.E., 1974. Responses of the piglet to perinatal stressors. J. Anim. Sci. 38, 1031–1035.
Curtis, S., Rogler, J.C., 1970. Thermoregulatory ontogeny in piglets: sympathetic and adipokinetic
response to cold. Amer. J. Physiol. 218, 149–152.
Damon, M., Vincent, A., Lombardi, A., Herpin, P., 2000a. First evidence of uncoupling protein-2
(UCP-2) and -3 (UCP-3) gene expression in piglet skeletal muscle and adipose tissue. Gene 246,
133–141.
Damon, M., Vincent, A., Herpin, P., 2000b. Characterization and expression of skeletal and hepatic car-
nitine palmitoyl transferase I genes in piglet skeletal muscle. Nutr. Clin. Metabol. 14, Suppl. 2, 158S.
Dauncey, M.J., 1990. Thyroid hormones and thermogenesis. Proc. Nutr. Soc. 49, 203–215.
Dauncey, M.J., Burton, K.A., Tivey, D.R., 1994. Nutritional modulation of insulin-like growth factor-I
expression in early postnatal piglets. Pediat. Res. 36, 77–84.
Dauncey, M.J., Wooding, F.B., Ingram, D.L., 1981. Evidence for the presence of brown adipose tissue in
the pig. Res. Vet. Sci. 31, 76–81.
Energy metabolism in early development of the pig 371

De Lange, P., Lanni, A., Beneduce, L., Moreno, M., Lombardi, A., Silvestri, E., Goglia, F., 2001.
Uncoupling protein-3 is a molecular determinant for the regulation of resting metabolic rate by
thyroid hormone. Endocrinology 142, 3414–3420.
Duchamp, C., Barré, H., 1993. Skeletal muscle as the major site of nonshivering thermogenesis in cold-
acclimated ducklings. Amer. J. Physiol. 265, R1076–R1083.
Duchamp, C., Burton, K.A., Herpin, P., Dauncey, M.J., 1994. Perinatal ontogeny of porcine nuclear
thyroid hormone receptors and its modulation by thyroid status. Amer. J. Physiol. 267, E687–E693.
Duchamp, C., Burton, K.A., Herpin, P., Dauncey, M.J., 1996. Perinatal ontogeny of porcine growth
hormone receptor gene expression is modulated by thyroid status. Eur. J. Endocrinol. 134, 524–532.
Duée, P.H., Pégorier, J.P., Le Dividich, J., Girard, J., 1988. Metabolic and hormonal response to acute
cold exposure in newborn pig. J. Dev. Physiol. 10, 371–381.
Duée, P.H., Pégorier, J.P., Péret, J., Girard, J., 1985. Separate effects of fatty acid oxidation and glucagon
on gluconeogenesis in isolated hepatocytes from newborn pigs. Biol. Neonate 47, 77–83.
Dunshea, F.R., Suster, D., Kerton, D.J., Leury, B.J., 2001. Porcine somatotropin treatment of neonatal
pigs alters lifetime fat but not lean accretion. In: Cranwell, P.D. (Ed.), Manipulating Pig Production
VIII. Australasian Pig Science Association, Werribee, Victoria, Australia, p. 42.
Elliot, J.I., Lodge, G.A., 1977. Body composition and glycogen, reserves in the neonatal pig during the
first 96 hours post-partum. Can. J. Anim. Sci. 57, 141–150.
Eppley, Z.A., Russell, B., 1995. Perinatal changes in avian muscle: implications from ultrastructure for
the development of endothermy. J. Morphol. 225, 357–367.
Flecknell, P.A., Wootton, R., John, M., 1980. Total body glucose metabolism in the conscious, unre-
strained piglet and its relation to body and organ weight. Brit. J. Nutr. 44, 193–203.
Fleury, C., Neverova, M., Collins, S., Raimbault, S., Champigny, O., Levi-Meyrueis, C., Bouillaud, F.,
Seldin, M.F., Surwit, R.S., Ricquier, D., Warden, C.H., 1997. Uncoupling protein-2: a novel gene
linked to obesity and hyperinsulinemia. Nat. Genet. 15, 269–272.
Fiorotto, M.L., Davis, T.A., Reeds, P., Burrin, D., 2000. Nonnutritive factors in colostrum enhance
myofibrillar protein synthesis in the newborn pig. Pediat. Res. 48, 511–517.
Flucher, B.E., 1992. Structural analysis of muscle development: transverse tubules, sarcoplasmic reticulum,
and the triad. Dev. Biol. 154, 245–260.
Foster, D.O., Frydman, M.L., 1979. Tissue distribution of cold-induced thermogenesis in conscious
warm- or cold-acclimated rats reevaluated from changes in tissue blood flow: the dominant role of
brown adipose tissue in the replacement of shivering by non-shivering thermogenesis. Can. J. Physiol.
Pharmacol. 57, 257–270.
Fraser, D., Rushen, J., 1992. Colostrum intake by the newborn piglet. Can. J. Anim. Sci. 72, 1–13.
Gentz, J., Bengtsson, G., Hakkarainen, J., Hellström, R., Persson, B., 1970. Factors influencing oxygen
consumption in the newborn pig with special reference to feeding. Biol. Neonate 16, 328–341.
Gerfault, V., Louveau, I., Mourot, J., Le Dividich, J., 2000. Lipogenic enzyme activity in subcutaneous
adipose tissue and skeletal muscle from neonatal pig consuming maternal or formula milk. Reprod.
Nutr. Dev. 40, 103–111.
Girard, J., 1986. Gluconeogenesis in late and early life. Biol. Neonate 59, 257–267.
Girard, J., Férré, P., Pégorier, J.P., Duée, P.H., 1992. Adaptations of glucose and fatty acid metabolism
during perinatal period and suckling-weaning transition. Physiol. Rev. 72, 501–562.
Gong, D.W., He, Y., Karas, M., Reitman, M., 1997. Uncoupling protein-3 is a mediator of thermogenesis
regulated by thyroid hormone, beta3-adrenergic agonists, and leptin. J. Biol. Chem. 272, 24129–24132.
Gong, D.W., Monemdjou, S., Gavrilova, O., Leon, L.R., Marcus-Samuels, B., Chou, C.J., Everett, C.,
Kozak, L.P., Li, C., Deng, C., Harper, M.E., Reitman, M.L., 2000. Lack of obesity and normal
response to fasting and thyroid hormone in mice lacking uncoupling protein-3. J. Biol. Chem. 275,
16251–16257.
Goodwin, R.F.W., 1957. The relationship between the concentration of blood sugar and some vital func-
tions in the newborn pig. J. Physiol. (Lond.) 136, 208–217.
Gregg, V.A., Milligan, L.P., 1982. Role of Na+/K+-ATPase in muscular energy expenditure of warm- and
cold-exposed sheep. Can. J. Anim. Sci. 62, 123–132.
Handel, S.E., Stickland, N.C., 1987. The effect of low birthweight on the ultrastructural development of
two myofiber types in the pig. J. Anat. 150, 129–143.
Harrell, R.J., Thomas, M.J., Boyd, R.D., Czerwinski, S.M., Steele, N.C., Bauman, D.E., 1999. Ontogenic
maturation of the somatotropin/insulin-like growth factor axis. J. Anim. Sci. 77, 2934–2941.
Harrison, A.P., Clausen, T., Duchamp, C., Dauncey, M.J., 1994. Roles of skeletal muscle morphology and
activity in determining Na+–K+-ATPase concentration in young pigs. Amer. J. Physiol. 266, R102–R111.
372 P. Herpin et al.

Herpin, P., Barré, H., 1989. Loose-coupled subsarcolemmal mitochondria from muscle rhomboideus in
cold-acclimated piglets. Comp. Biochem. Physiol. B 92, 59–65.
Herpin, P., McBride, B.W., Bayley, H.S., 1987a. Effects of cold exposure on energy metabolism in the
young pig. Can. J. Physiol. Pharmacol. 65, 236–245.
Herpin, P., Bertin, R., Le Dividich, J., Portet, R., 1987b. Some regulatory aspects of thermogenesis in
cold-exposed piglets. Comp. Biochem. Physiol. 87, 1073–1081.
Herpin, P., Le Dividich, J., van Os, M., 1992. Contribution of colostral fat to thermogenesis and glucose
homeostasis in the newborn pig. J. Dev. Physiol. 17, 133–141.
Herpin, P., Le Dividich, J., Amaral, N., 1993. Effect of selection for lean tissue growth on body compo-
sition and physiological state of pig at birth. J. Anim. Sci. 71, 2645–2653.
Herpin, P., Le Dividich, J., Berthon, D., Hulin, J.C., 1994. Assessment of thermoregulatory and post-
prandial thermogenesis over the first 24 hours after birth in pigs. Exp. Physiol. 79, 1011–1019.
Herpin, P., Berthon, D., Bertin, R., De Marco, F., Dauncey, M.J., Le Dividich, J., 1995. Cold-induced
changes in circulating levels of catecholamines and thyroid hormones are modulated by energy intake
in newborn pigs. Exp. Physiol. 80, 877–880.
Herpin, P., Berthon, D., Duchamp, C., Dauncey, M.J., Le Dividich, J., 1996. Effect of thyroid status in
the perinatal period on oxidative capacities and mitochondrial respiration in porcine liver and skele-
tal muscle. Reprod. Fertil. Dev. 8, 147–155.
Herpin, P., Lossec, G., Schmidt, I., Cohen-Adad, F., Duchamp, C., Lefaucheur, L., Goglia, F., Lanni, A.,
2002. Effect of age and cold exposure on morphofunctional characteristics of skeletal muscle in
neonatal pigs. Pflügers Arch. Eur. J. Physiol. 444, 610–618.
Himms-Hagen, J., Harper, M.-E., 2001. Physiological role of UCP3 may be export of fatty acids from
mitochondria when fatty acid oxidation predominates: an hypothesis. Exp. Biol. Med. 226, 78–84.
Hoppeler, H., 1986. Exercise-induced ultrastructural changes in skeletal muscle. Int. J. Sports Med. 7,
187–204.
Hoppeler, H., Mathieu, O., Krauer, R., Claassen, H., Armstrong, R.B., Weiber, E.R., 1981. Design of the
mammalian respiratory system. VI. Distribution of mitochondria and capillaries in various muscles.
Resp. Physiol. 44, 87–111.
Horn, G.W., Foley, C.W., Seerley, R.W., Munnel, J.F., 1973. Role of adipose tissue lipolysis in postnatal
amelioration of thermogenesis in domestic and wild piglets. J. Anim. Sci. 37, 1356–1361.
Jones, G., Edwards, S.A., Traver, S., Jagger, S., Hoste, S., 1999. Body composition changes in piglets at
weaning to nutritional modification of sow milk composition and effects on post-weaning perform-
ance. Paper presented at the 50th Annual Meeting of the EAAP, Zürich, 5 pp.
Kaciuba-Uscilko, H., 1972. Hormonal regulation of thermogenesis in the new-born pig. Biol. Neonate
21, 245–258.
Lazar, M.A., 1993. Thyroid hormone receptors: multiple forms, multiple possibilities. Endocr. Rev. 14,
184–193.
LeBlanc, J., Mount, L.E., 1968. Effects of noradrenaline and adrenaline on oxygen consumption rate and
arterial blood pressure in the newborn pig. Nature 217, 77–78.
Le Dividich, J., Esnault, Th., Lynch, B., Hoo-Paris, R., Castex, Ch., Peiniau, J., 1991b. Effect of colostral
fat on fat deposition and plasma metabolites in newborn pig. J. Anim. Sci. 69, 2480–2488.
Le Dividich, J., Herpin, P., Paul, E., Strullu, F., 1997. Effect of fat content in colostrum on voluntary
colostrum intake and fat utilization in the newborn pig. J. Anim. Sci. 75, 707–713.
Le Dividich, J., Herpin, P., Rosario-Ludovino, R.M., 1994. Utilization of colostrum energy by the new-
born pig. J. Anim. Sci. 72, 2082–2089.
Le Dividich, J., Mormède, P., Catheline, M., Caritez, J.C., 1991a. Body composition and cold resistance of the
neonatal pig from European (Large White) and Chinese (Meishan) breeds. Biol. Neonate 59, 268–277.
Le Dividich, J., Noblet, J., Herpin, P.H., van Milgen, J., Quiniou, N., 1998. Thermoregulation. In:
Wiseman, J., Varley, M.A., Charlick, J.P. (Eds.), Progress in Pig Science. Nottingham University
Press, Nottingham, UK, pp. 229–263.
Le Dividich, J., Sève, B., 2001. Energy requirement of the young pig. In: Varley, M.A., Wiseman,
J. (Eds.), The Weaner Pig. CAB International, Wallingford, UK, pp. 17–44.
Lee, C.Y., Bazer, F.W., Etherton, T.D., Simmen, F.A., 1991. Ontogeny of insulin-like growth factors
(IGF-I and IGF-II) and IGF-binding proteins in porcine serum during fetal and postnatal develop-
ment. Endocrinology 128, 2336–2344.
Lee, C.Y., Chung, C.S., Simmen, F.A., 1993. Ontogeny of the porcine insulin-like growth factor system.
Mol. Cell. Endocrinol. 63, 71–80.
Energy metabolism in early development of the pig 373

Leenhouvers, J.I., 2001. Biological aspects of genetic differences in piglet survival. Ph.D. dissertation,
Wageningen University, Wageningen, The Netherlands.
Lefaucheur, L., Ecolan, P., Lossec, G., Gabillard, J.C., Butler-Browne, G.S., Herpin, P., 2002. Influence
of early postnatal cold-exposure on myofiber maturation in pig skeletal muscle. J. Muscle Res. Cell.
Motil. 22, 439–452.
Lefaucheur, L., Hoffman, R., Okamura, C., Gerrard, D., Leger, J.J., Rubinstein, N., Kelly, A., 1997.
Transitory expression of alpha cardiac myosin heavy chain in a subpopulation of secondary genera-
tion muscle fibers in the pig. Dev. Dyn. 210, 106–116.
Lefaucheur, L., Vigneron, P., 1986. Post-natal changes in some histochemical and enzymatic character-
istics of three pig muscles. Meat Sci. 16, 199–216.
Léon, A., Schmodt, I., Strullu, F., Fillaut, M., Gauthier, J., Hulin, J.C., Lebreton, Y., Herpin, P., Le
Dividich, J., 1998. Effects of substitution of medium- (MCFA) for long- (LCFA) chain fatty acids in
colostrum on energy metabolism of the newborn pig in relation to envirionmental temperature.
Journée Rech. Porcine en France 30, 275–280.
Lepine, A.J., Boyd, R.D., Whitehead, D.M., 1991. Effect of colostrum intake on hepatic gluconeogene-
sis and fatty acid oxidation in the neonatal pig. J. Anim. Sci. 69, 1966–1974.
Lombardi, A., Damon, M., Vincent, A., Goglia, F., Herpin, P., 2000. Characterisation of oxidative phos-
phorylation in skeletal muscle mitochondria subpopulations in pig: a study using top-down elasticity
analysis. FEBS Lett. 475, 84–88.
Lossec, G., 1998. Muscle energy metabolism during cold-induced shivering in the newborn pig. Ph.D.
Thesis, Paris University, 210 pp.
Lossec, G., Duchamp, C., Lebreton Y., Herpin, P., 1999. Postnatal changes in regional blood flow during
cold-induced shivering in sow-reared piglets. Can. J. Physiol. Pharmacol. 77, 1–8.
Lossec, G., Herpin, P., Le Dividich, J., 1998a. Thermoregulatory responses of the newborn pig during
experimentally induced hypothermia and rewarming. Exp. Physiol. 83, 667–678.
Lossec, G., Lebreton, Y., Hulin, J.C., Fillaut, M., Herpin, P., 1998b. Age-related changes in oxygen and
nutrient uptake by hindquarters in newborn pigs during cold-induced shivering. Exp. Physiol. 83,
793–807.
Louveau, I., Bonneau, M., 2001. Biology and actions of somatotropin in the pig. In: Renaville, R., Burny, A.
(Eds.), Biotechnology and Animal Husbandry, Kluwer Academic Press, Dordrecht, The Netherlands,
pp. 111–131.
Louveau, I., Combes, S., Cochard, A., Bonneau, M., 1996. Developmental changes in insulin-like growth
factor-I (IGF-I) receptor levels and plasma IGF-I concentrations in Large White and Meishan pigs.
Gen. Comp. Endocrinol. 104, 29–36.
Louveau, I., Le Dividich, J., 2002. GH and IGF-I binding in adipose tissue, liver and skeletal muscle in
response to milk intake level in piglets. Gen. Comp. Endocrinol. 126, 310–317.
Marion, J., Le Dividich, J., 1999. Utilization of the energy in sow’s milk by the piglet. In: Cranwell, P.D.
(Ed.), Manipulating Pig Production VII. Australasian Pig Science Association, Werribee, Victoria,
Australia, p. 254.
Mayfield, S.R., Stonestreet, B.S., Brubakk, A.M., Shaul, P.W., Susa, J., Oh, W., 1986. Regional blood
flow in newborn piglets during environmental cold stress. Amer. J. Physiol. 251, G308–G313.
Mayfield, S.R., Stonestreet, B.S., Shaul, P.W., Brubakk, A.M., Susa, J., Oh, W., 1989. Plasma
catecholamines concentrations of newborn piglets in thermoneutral and cold environments. J. Dev.
Physiol. 11, 331–334.
McGarry, J.D., Brown, N.F., 1997. The mitochondrial carnitine palmitoyltransferase system: from con-
cept to molecular analysis. Eur. J. Biochem. 244, 1–14.
Mellor, D.J., Cockburn, F., 1986. A comparison of energy metabolism in the newborn infant, piglet and
lamb. Q. J. Exp. Physiol. 71, 361–371.
Mersmann, H.J., 1974. Metabolic patterns in the neonatal pig. J. Anim. Sci. 38, 1022–1030.
Mersmann, H.J., Goodman, J., Houk, J.M., Anderson, S., 1972. Studies on the biochemistry of mito-
chondria and cell morphology in the neonatal swine hepatocyte. J. Cell. Biol. 53, 335–347.
Morovat, A., Dauncey, M.J., 1995. Regulation of porcine skeletal muscle nuclear 3,5,3′-triiodothyronine
receptor binding capacity by thyroid hormones: modification by energy balance. J. Endocrinol. 144,
233–242.
Mount, L., 1968. The Climatic Physiology in the Pig. Edward Arnold, London.
Mutvei, A., Husman, B., Anderson, G., Nelson, B.D., 1989. Thyroid hormone and not growth hormone
is the principal regulator of mammalian mitochondrial biogenesis. Acta Endocrinol. 121, 223–228.
374 P. Herpin et al.

Newcomb, M.D., Harmon, D.L., Nelssen, J.L., Thulin, A.J., Allee, G.L., 1991. Effect of energy source
fed to sows during late gestation on neonatal blood metabolite homeostasis, energy stores and com-
position. J. Anim. Sci. 69, 230–236.
Nicot, C., Hegardt, F.G., Woldegiorgis, G., Haro, D., Marrero, P.F., 2001. Pig liver carnitine palmitoyl-
transferase I, with low Km for carnitine and high sensitivity to malonyl-CoA inhibition, is a natural
chimera of rat liver and muscle enzymes. Biochemistry 40, 2260–2266.
Noblet, J., Le Dividich, J., 1981. Energy metabolism of the newborn pig during the first 24 h after birth.
Biol. Neonate 40, 175–182.
Odle, J., 1997. New insights into the utilization of medium-chain triglycerides by the neonate: observa-
tions from a piglet model. J. Nutr. 127, 1061–1067.
Odle, J., Lin, X., van Kempen, T.A.T.G., Drackley, J.K., Adams, S.H., 1995. Carnitine palmitoyltrans-
ferase modulation of hepatic fatty acid metabolism and radio-HPLC evidence for low ketogenesis in
neonatal pigs. J. Nutr. 125, 2541–2549.
Pégorier, J.P., Duée, P.H., Simoes-Nunes, C., Peret, J. Girard, J., 1984. Glucose turnover and recycling in
unrestrained and unaesthetized 48-h-old fasting or post-absorptive newborn pigs. Brit. J. Nutr. 52,
277–287.
Pégorier, J.P., Simoes-Nunes, C., Duée, P.H.,. Peret, J. Girard, J., 1985. Effect of intragastric triglycerides
administration on glucose homeostasis in newborn pigs. Amer. J. Physiol. 249, E268–E275.
Père, M.C., 1995. Maternal and fetal blood levels of glucose, lactate, fructose and insulin in the conscious
pig. J. Anim. Sci. 73, 2994–2999.
Persson, B., Gentz, J.C.H., Hakkarainen, J., Kellum, M., 1971. Catecholamine-induced lipolysis and its
relation to oxygen consumption in the newborn piglet. Pediat. Res. 5, 435–445.
Quiniou, N., Dagorn, J., Gaudré, D., 2002. Variation of piglets’ birth weight and consequences on sub-
sequent performance. Livest. Prod. Sci., Special Issue 78, 63–70.
Randall, G.C., 1983. Changes in the concentrations of cortocosteroids in the blood of fetal pigs and their
dams during late gestation and labor. Biol. Reprod. 29, 1077–1084.
Ricquier, D., Bouillaud, F., 2000. Mitochondrial uncoupling proteins: from mitochondria to the regula-
tion of energy balance. J. Physiol. 529, 3–10.
Sarkar, N.K., Kramer, J.K.G., Wolynetz, M.S., Elliot, J.I., 1985. Influence of dietary fat on growth and
body composition of the neonatal pig. Can. J. Anim. Sci. 65, 175–184.
Scanes, C.G., Lazarus, D., Bowen, S., Buonomo, F.C., Gilbreath, R.L., 1987. Postnatal changes in cir-
culating concentrations of growth hormone, somatomedin C and thyroid hormones in pigs. Domest.
Anim. Endocrinol. 4, 252–257.
Schmidt, I., Herpin, P., 1998. Carnitine palmitoyltransferase I (CPT I) activity and its regulation by mal-
onyl-CoA are modulated by age and cold exposure in skeletal muscle mitochondria from newborn
pigs. J. Nutr. 128, 886–893.
Schmidt, I., Herpin, P., 1997. Postnatal changes in mitochondrial protein mass and respiration in skeletal
muscle from the newborn pig. Comp. Biochem. Physiol. B118 639–647.
Schnoebelen-Combes, S., Louveau, I., Postel-Vinay, M.-C., Bonneau, M., 1996. Ontogeny of GH recep-
tor and GH-binding protein in the pig. J. Endocrinol. 148, 249–255.
Slebodzinski, A.B., 1979. Metabolic responses to thyroxine in the newborn pig. Biol. Neonate 36, 198–205.
Slebodzinski, A.B., Nowak, G., Zamyslowska, H., 1981. Sequential observation of changes in thyroxine,
triiodothyronine and reverse triiodothyronine during the postnatal adaptation of the pig. Biol. Neonate
39, 191–199.
Steffen, D.G., Brown, L.J., Mersmann, H.J., 1978. Ontogenic development of swine (Sus domesticus)
adipose tissue lipases. Comp. Biochem. Physiol. B 59, 195–198.
Trayhurn, P., Temple, N.J., Van Aerde, J., 1989. Evidence from immunoblotting studies on uncoupling protein
that brown adipose tissue is not present in the domestic pig. Can. J. Physiol. Pharmacol. 67, 1480–1485.
Thulin, A.J., Allee, G.L., Harmon, D.L., Davis, D.L., 1989. Utero-placental transfer of octanoic, palmitic
and linoleic acids during late gestation in gilts. J. Anim. Sci. 67, 738–745.
Wrutniak-Cabello, C., Casas, F., Cabello, G., 2001. Thyroid hormone action in mitochondria. J. Mol.
Endocrinol. 26, 67–77.
Yu, X.X., Drackley, J.K., Odle, J., 1997. Rates of mitochondrial and peroxisomal β-oxidation of palmi-
tate change during postnatal development and food deprivation in liver, kidney and heart of pigs.
J. Nutr. 127, 1814–1821.
15 Hepatic gluconeogenesis in developing
ruminants

S. S. Donkina and H. Hammonb

aDepartment of Animal Sciences, Purdue University, West Lafayette, IN 47907, USA


bResearchInstitute for Biology of Farm Animals Nutrition Physiology
(Oskar Kellner Institute), 18196 Dummerstorf, Germany.

The transition from preruminating to ruminating status represents one of the most dramatic
changes in glucose metabolism in mammals. Within 5 weeks of birth, ruminants must
undergo the anatomical and physiological adaptations necessary to permit extensive fermen-
tation of plant materials in the rumen and postabsorptive utilization of the end-products.
Several well-characterized metabolic adaptations have been documented that act to spare glucose
oxidation with the onset of rumination; however, the endocrine and molecular factors that
modulate changes in glucose synthesis and metabolism during this transition are not yet fully
characterized. This review focuses on the endocrine and metabolic state of the ruminant fetus
at term, the development of metabolic competence in the neonatal ruminant, and changes that
occur during the transition to ruminating status.

1. GLUCONEOGENIC SUBSTRATES AND METABOLISM


Propionate, lactate, and amino acids furnish most of the carbon used for gluconeogenesis in
fed ruminants and glycerol provides some gluconeogenic carbon during feed restriction
(Huntington, 1990). In neonatal and developing ruminants, milk lactose supplies approxi-
mately 25% of the daily glucose needs (Girard, 1990). In the absence of a functional rumen,
amino acids, lactate, and to a limited extent, glycerol from milk are used for gluconeogenesis.
Development of the fermentation capacity of the rumen is accompanied by changes in the
type of carbohydrates ingested, reductions in the amount of fat in the diet, a decrease in avail-
ability of dietary carbohydrate to the developing ruminant, and an increased supply of
propionate as a gluconeogenic precursor. In both the preruminant and ruminant states the
need for active gluconeogenesis to maintain glucose homeostasis is apparent.
Pyruvate is a common entry point in the gluconeogenic pathway for lactate, alanine, and
other gluconeogenic amino acids. Pyruvate formed from lactate and amino acids is transported
into the mitochondria and carboxylated to oxaloacetate by pyruvate carboxylase (PC) (fig. 1).

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
375 © 2005 Elsevier Limited. All rights reserved.
Fig. 1. Metabolism of propionate, lactate, pyruvate, and alanine to glucose in bovine liver. Substrate abbreviations are given in the key. Reactions catalyzed by the key enzymes discussed
in the text, cytosol phosphoenolpyruvate carboxykinase (PEPCK-C), mitochondrial phosphoenolpyruvate carboxykinase (PEPCK-M), pyruvate carboxylase (PC), and pyruvate kinase (PK),
are indicated by the shaded backgrounds.
Hepatic gluconeogenesis in developing ruminants 377

Propionate, in contrast, is metabolized through part of the TCA cycle to oxaloacetate follow-
ing activation to propionyl-CoA and metabolism through propionyl-CoA carboxylase,
methylmalonyl-CoA racemase, and methylmalonyl-CoA mutase. Oxaloacetate can be metab-
olized to phosphoenolpyruvate (PEP) by phosphoenolpyruvate carboxykinase (PEPCK) or
metabolized in the TCA cycle. In turn, PEP carbon can be metabolized to glucose or recycled
to pyruvate via pyruvate kinase (PK). In order for lactate carbon to be metabolized to glucose,
the flux through PEPCK and PC must exceed the PK flux, whereas net flux of propionate
carbon requires only a greater flux through PEPCK relative to PK. Therefore, an increase in
PEPCK activity in the absence of changes in PC activity would favor the use of propionate
for gluconeogenesis.
The presence of PEPCK activity in the cytosol (PEPCK-C) and mitochondria (PEPCK-M)
is one of the important features of gluconeogenesis that permits compartmentalization of the
pathway and results in the characteristic pattern of regulation and use of lactate and pyruvate.
The distribution of this activity is uniquely species-dependent and most mammals display
both a mitochondrial and a cytosolic form of the PEPCK enzyme. Rodents express primarily
PEPCK-C and both forms are found in liver of the developing chicken, yet only the mitochon-
drial form is found in liver from the adult chicken. There are approximately equal activities
of PEPCK-M and PEPCK-C in the ruminant (Taylor et al., 1971) and human (Hod et al.,
1987) liver. Bovine PEPCK-C and PEPCK-M have been recently cloned and characterized
(Agca et al., 2002) and the ratio of mRNA indicates a 10-fold greater expression of PEPCK-C
than PEPCK-M in lactating cows. Similar data are not yet available for developing bovine.
The stoichiometry of gluconeogenesis dictates that the formation of phosphoenolpyruvate
from propionate, pyruvate, and some amino acids requires the independent synthesis of
NADH in the cytosol for the subsequent reduction of 1,3-diphosphoglycerate in gluconeo-
genesis. It has been proposed that PEPCK-C is required for gluconeogenesis from amino
acids and PEPCK-M is more suited to gluconeogenesis from lactate (Watford et al., 1981).
Pyruvate and amino acids are metabolized to oxaloacetate in mitochondria and are shuttled to
the cytosol as malate from which NADH and oxaloacetate are regenerated followed by PEP
formation that is catalyzed by PEPCK-C. Lactate can also be metabolized to PEP in mito-
chondria of species that possess appreciable PEPCK-M activity and subsequently shuttled to
the cytosol (Holcomb et al., 1995).

2. GLUCOSE RELEASE FROM HEPATOCYTES


Glucose-6-phosphatase is a membrane-bound enzyme that is located on the internal mem-
brane of the endoplasmic reticulum and is involved in the terminal step of gluconeogenesis
as well as glycogenolysis. The enzyme catalyzes the conversion of glucose-6-phosphate to
glucose to enable release from the cell. In nonruminants the enzyme is expressed in liver,
kidney cortex, and jejunum, but only the liver form of the enzyme is upregulated at birth and
to weaning in rodents (Chatelain et al., 1998; Kalhan and Parimi, 2000). The hepatocyte glucose
transporter, GLUT2, and glucose-6-phosphatase act in concert to control the release of glucose
from liver. The symmetry of GLUT2 enables the transport of glucose into or out of the hepato-
cyte and the directionality depends only on the concentration differential between intracellular
free glucose and blood glucose (Burchell, 1994).
Glucose-6-phosphate (G-6-P), formed through gluconeogenesis or glycogenolysis, must be
dephosphorylated through the action of glucose-6-phosphatase (G-6-Pase), an enzyme that is
contained within the endoplasmic reticulum in order to release glucose from the hepatocyte.
An endoplasmic glucose transporter GLUT7 was initially proposed that would facilitate the
378 S. S. Donkin and H. Hammon

transport of G-6-P to the endoplasmic reticulum where G-6-Pase acts to release free glucose
into the cytoplasm (Burchell, 1994), but has since been retracted (Burchell, 1998). It is now
thought that G-6-Pase acts in combination with a specific G-6-P translocase to channel G-6-P
into the endoplasmic reticulum where G-6-Pase is compartmentalized (Van Schaftingen and
Gerin, 2002). Recently a second transporter protein for G-6-P has been identified (Hosokawa
and Thorens, 2002) which complements the activity of the specific G-6-P translocase.
When G-6-P is overexpressed in hepatocytes there is a marked increase in glucose release
and a decline in intracellular G-6-P and glycogen concentrations (Seoane et al., 1997; Aiston
et al., 1999). Measures of G-6-Pase activity at term indicate that the capacity of the enzyme
is fully developed at birth in ruminants (Edwards et al., 1975; Stevenson et al., 1976;
Narkewicz et al., 1993). Unfortunately data are not yet available for ruminants describing
developmental changes in expression of G-6-P translocases.

3. GLUCONEOGENESIS IN FETAL RUMINANTS


The contribution of fetal gluconeogenesis to the glucose needs of the developing ruminant
conceptus is equivocal. A portion of the discrepancies regarding the contribution of gluco-
neogenesis to the glucose economy of the developing fetus lies in recycling errors that are
inherent to measuring glucose entry rates using isotope dilution (Kalhan and Parimi, 2000).
Available data indicate that uterine glucose requirements during the last trimester of preg-
nancy account for 20–70% of the glucose needs of pregnant ewes (Prior, 1982), that fetal
glucose uptake is reduced when ewes are deprived of feed (Tsoulos et al., 1971; Boyd et al.,
1973; Chandler et al., 1985; Leury et al., 1990), and that the rate of fetal glucose utilization
during maternal feed restriction is less affected than glucose removed via the umbilical artery
(Hay et al., 1984). These observations imply that the hypoglycemic ovine fetus is capable of
significant endogenous glucose release and is subject to activation in utero in response to
maternal nutrition.
Amino acids and lactate are the major gluconeogenic substrates in the fetus, and urea excre-
tion rates indicate that 25% of oxygen consumption by fetal ovine liver is due to amino acid
catabolism (Gresham, 1972). Efficient extraction of propionate by maternal liver precludes
appreciable propionate supply to the developing ruminant fetus for gluconeogenesis. In some
experiments, gluconeogenesis from lactate accounts for 22% of lactate turnover and supplies
49% of fetal glucose (Prior, 1980), but in other experiments fetal gluconeogenesis from
lactate was undetectable (Warnes et al., 1977). These contradictory observations reflect vari-
ations in maternal nutrition immediately prior to the experimental period (Girard et al., 1992).
Experimental evidence suggests that the rate of gluconeogenesis is increased in fetal lamb
liver in response to inadequate nutrition of the dam (Leury et al., 1990; Apatu and Barnes,
1991a). Increased fetal urea production during nutritional insufficiency in pregnant ewes is
consistent with an increase in fetal gluconeogenesis from amino acids (Hodgson et al., 1982).
Therefore gluconeogenesis is active and adaptable in fetal ruminants, unlike rodents, and
plays a critical role in the glucose economy of the maternal–fetal unit.
Enzyme activities have been used to characterize developmental changes in liver metabo-
lism and provide an estimate of the maximum flux through a single step in the gluconeogenic
pathway. The activities of glucose-6-phosphatase, fructose-1,6,-bisphosphatase, pyruvate
carboxylase, and PEPCK, key enzymes for gluconeogenesis, are similar in term-fetal, neonatal,
and adult sheep (Warnes et al., 1977; Narkewicz et al., 1993). There appears to be sequential
development of gluconeogenic enzymes in caprine (Dhanotiya and Bhardwaj, 1988) and ovine
fetuses (Stevenson et al., 1976). However, the enzymes of the gluconeogenic path are present
Hepatic gluconeogenesis in developing ruminants 379

in fetal liver by 100−128 days of gestation (Warnes et al., 1977; Prior, 1980). There is greater
activity of alanine aminotransferase in fetal than in neonatal liver, which may reflect a greater
capacity for amino acid metabolism to glucose in utero. Changes in hepatic enzyme activity
during the period of rumen development are modest and changes in glucose metabolism
reflect decreased glycolytic activity in both muscle and liver (Howarth et al., 1968; Pearce
and Unsworth, 1980). The activity of lactate dehydrogenase and alanine aminotransferase are
lower in adult ewes compared to neonatal sheep or 3-month-old lambs (Edwards et al., 1975),
suggesting decreased capacity to metabolize alanine to glucose with development in neonatal
ruminants.

4. NUTRITIONAL CHANGES AT BIRTH


At birth the neonate must cope with the loss of umbilical glucose supply and survive a brief
period of starvation before receiving colostrum and milk. Liver glycogen, at birth, is approx-
imately 4–6% of liver wet weight. This energy reserve is depleted within a few hours
(Hamada and Matsumoto, 1984; Girard, 1990) and supplies glucose for erythrocytes, brain,
and kidney medulla. Glucose supplied by milk lactose accounts for approximately 25% of
glucose utilization of the neonatal lamb (Girard, 1986), therefore gluconeogenesis is necessary
to maintain neonatal glucose homeostasis.
Development of supporting pathways, production of cofactors, and substrate supply may
affect the rates of gluconeogenesis in utero and during postnatal development. For example
the inability to oxidize fatty acids at birth has been characterized in detail and stems from a
lack of activity of fatty acyl-CoA synthases, carnitine palmitoyltransferase I (CPT-I), enoyl-
CoA hydratase, 3-hydroxyacyl-CoA dehydrogenase, and oxoacyl-CoA thiolase (Girard et al.,
1992). The gluconeogenic promoting effects of fatty acids have been recognized for some
time in nonruminants (Williamson et al., 1966). A similar regulation is likely in the bovine
based on data that indicate that specific long-chain fatty acids promote gluconeogenesis in
hepatocytes from ruminating calves (Mashek et al., 2002). Likewise, inhibition of CPT-I activity
decreases gluconeogenesis in sheep hepatocytes (Chow and Jesse, 1992). Information is lacking
on initiation of fatty acid oxidation in neonatal ruminants, but if a parallel can be drawn from
rodent data, the induction of gluconeogenic capacity may be linked to the induction of fatty
acid oxidation.

5. HORMONAL CHANGES: INSULIN, GLUCAGON,


GLUCOCORTICOIDS
Nutrient supply during the prenatal period consists primarily of a carbohydrate-rich energy
supply (glucose, lactate), yet during the neonatal period a switch is made to a high-fat, low-
carbohydrate diet (Aynsley-Green, 1988). Newborns develop marked hypoglycemia after
birth because glucose derived from lactose in colostrum does not meet postnatal glucose
demands (Girard, 1986). Therefore, glycogenolysis and gluconeogenesis increase rapidly
in the liver of newborns; however, there are species differences in prenatal development of glu-
coneogenesis. In the developing rat and pig fetus the gluconeogenic pathway does not mature
in utero (Ballard and Oliver, 1963; Swiatek, 1971), whereas in the bovine fetus there is gluco-
neogenic activity measurable from day 80 of gestation (Prior and Scott, 1977). This might
indicate, as discussed above, that the bovine fetus is less dependent on maternal glucose
supply than the rat and pig fetus; however, newborn calves experience hypoglycemia as do
other species (Aynsley-Green, 1988; Egli and Blum, 1988; Hadorn et al., 1997; Hammon and
380 S. S. Donkin and H. Hammon

Blum, 1998). In neonatal ruminants and nonruminants, glucose is mostly produced by gluco-
neogenesis using amino acids (alanine, glycine, glutamine), glycerol, and lactate. In human
newborns, research using stable isotopes demonstrates that gluconeogenesis from lactate,
glycerol, and alanine occurs at a significant rate within the first 8 h of life and is critical for
neonatal survival (Ferre et al., 1986).
Insulin and glucagon are integral to normal fetal development in ruminants (reviewed in
Blum and Hammon, 1999a). The change in glucose supply that accompanies loss of umbili-
cal nutrient supply is reflected by an increase in glucagon and decreased insulin concentration
during the immediate postnatal period (Girard, 1990). Blood profiles in newborn calves are
characterized by hypoglycemia, high nonesterified fatty acids, and low triglyceride, phospho-
lipids, and cholesterol (Blum and Hammon, 1999b). Neonatal calves respond to nutritional
challenges by increasing glucagon and decreasing insulin in a manner similar to adult animals;
however, the glucose–insulin relationship is less developed in neonates. The lack of glucose
clearance in response to insulin may prevent hypoglycemia and serve to protect the neonatal calf.
Cortisol plays an important role in enhancing fetal capacity for glucose production and
glycogen storage (Fowden, 1995; Barnes, 1997). Changes in plasma insulin and glucagon
may be related to the stress associated with birth and the concomitant rise in serum cortisol,
fetal hypoxia, or both (Girard, 1990). Key genes for gluconeogenesis are also responsive
to thyroid hormones (Park et al., 1997); a rise in thyroxine during the first 24 h of life in
neonatal sheep (Fisher et al., 1977) may play a role in induction of metabolic competence.
Recent data indicate that fetal thyroid hormone production is essential to the development of
gluconeogenesis and is especially critical under adverse conditions such as undernutrition
(Fowden et al., 2001).
Plasma glucagon concentrations rise during the immediate postnatal period due, in part, to
a drop in blood glucose that occurs within a few hours of birth. An infusion of somatostatin
in lambs induces hypoglycemia and infusion of glucagon reverses the effects of somatostatin
(Sperling et al., 1977). The rate of glucose output by fetal, neonatal, and adult ovine liver was
increased similarly during glucagon infusions (Apatu and Barnes, 1991b); however, the effective
dose of glucagon necessary to stimulate gluconeogenesis is greater in fetal liver (Girard and
Sperling, 1983). Postnatal increases in glucagon receptor numbers and full development of
intracellular signal transduction pathways along with a decrease in insulin receptor numbers
favor regulation of gluconeogenesis in the neonate that is more sensitive to changes in glucagon
concentrations (Girard and Sperling, 1983).
Hepatic glucagon receptor numbers are lower in fetal and newborn ruminants than in
adults. The effective dose of glucagon required to stimulate gluconeogenesis in adult sheep
liver is not effective in fetal sheep liver (Girard and Sperling, 1983). In 21-day-old rats the
number of liver glucagon receptors was only 40% of the receptor number for adult liver
(Ganguli et al., 1983). Insulin receptor number and affinity are higher in fetal than in adult
liver in rats and humans (Neufeld et al., 1980). High insulin and low glucagon receptor activity
in utero favors glucose oxidation, whereas the coupling of glucagon receptor to cAMP synthesis
combined with an increase in glucagon receptor numbers in early postnatal life favors
gluconeogenesis (Girard and Sperling, 1983).

6. GLUCONEOGENESIS IN NEONATAL
AND DEVELOPING RUMINANTS
Gluconeogenesis from lactate is similar between fetal and maternal liver in the bovine (Prior
and Scott, 1977). The rates of [2-14C]propionate and [2-14C]lactate incorporation to glucose
Hepatic gluconeogenesis in developing ruminants 381

and glycogen obtained from liver slices from adult sheep and newborn lambs are similar
(Ballard and Oliver, 1963). In contrast, the rate of gluconeogenesis from [2-14C]pyruvate is
greater for liver slices from neonatal lambs compared with adult tissue. The rate of pyruvate
metabolism to glucose appears to peak at about 2–4 weeks of age in lambs (Ballard and
Oliver, 1963). Likewise the rates of metabolism of lactate are markedly reduced in weaned
lambs (Savan et al., 1986) and calves (Donkin and Armentano, 1995).
In developing ruminants there is a marked decline in the capacity to metabolize lactate to
glucose coupled with a reduced sensitivity to the effects of insulin (Donkin and Armentano,
1995). Radioisotope tracer data indicate that there is almost exclusive flux of lactate through
pyruvate carboxylase (PC) to glucose in neonatal calf liver and very little isotope exchange
with carbon of the TCA cycle (Donkin and Armentano, 1994). The substantial loss in lactate
metabolism to glucose during the preruminant to ruminant transition (Donkin and
Armentano, 1995), and similar use of propionate for gluconeogenesis between the two groups
(fig. 2), suggests a loss in capacity to draw lactate into the gluconeogenic pathway. These results
are perplexing in light of the extensive use of lactate for gluconeogenesis in nonruminants,
but agree with the 10-fold lower rate of glucose recycling in vivo in adult versus neonatal (5- or
21-day old) sheep (Muramatsu et al., 1974). These changes suggest developmentally regulated
differences in gluconeogenesis that are unique to lactate.
Lactate is equilibrated rapidly with pyruvate in liver. The rates of [1-14C]lactate and
[1-14C]pyruvate metabolism to glucose are not different for hepatocytes obtained from preru-
minating calves (Donkin and Armentano, 1994). This measurement has not been made
directly in hepatocytes from ruminating calves, but is not likely the limiting step in gluco-
neogenesis from lactate based on a lack of control of gluconeogenesis in response to
alterations in cytosolic redox state (Aiello and Armentano, 1987). As indicated above, pyru-
vate formed from lactate is carboxylated to oxaloacetate by pyruvate carboxylase (PC) and
oxaloacetate is either metabolized to phosphoenolpyruvate (PEP) or metabolized in the TCA
cycle. The similar ratios of [14C]glucose:14CO2 from [2-14C]propionate and carbons 2 and 3 of
lactate support a common oxaloacetate pool for the metabolism of propionate and lactate
in bovine hepatocytes (Donkin and Armentano, 1994). Therefore it is likely that decreased

Fig. 2. Effect of developmental state on gluconeogenesis from propionate and lactate in calves. Hepatocytes
were isolated from preruminating (n = 4) and ruminating calves (n = 3) and cultured for 48 h. The rate of glu-
coneogenesis from [2-14C]propionate or [U-14C]lactate was determined during the last 3 h of incubation.
Adapted from Donkin and Armentano (1995).
382 S. S. Donkin and H. Hammon

flux through PC is the cause of decreased gluconeogenesis from lactate during postnatal
development in calves.
Chronic exposure of neonatal bovine hepatocytes to insulin results in decreased gluconeo-
genesis from lactate (Donkin et al., 1997), which is consistent with data suggesting that a
portion of the reduction in gluconeogenesis from lactate in milk-fed calves may be due to
chronically elevated insulin concentrations (Breier et al., 1988). However, the direct actions
of insulin do not fully explain the reduction in gluconeogenesis from lactate that is observed
in calves during the transition from the preruminant to the ruminant state. Comparing the rate of
gluconeogenesis from lactate relative to propionate metabolism suggests additional changes
in hepatic lactate metabolism. Gluconeogenesis from lactate is reduced to 32% of the rate of
propionate conversion to glucose following chronic exposure to insulin (Donkin et al., 1997),
but the developmental change in lactate metabolism reduces gluconeogenesis from lactate to
only 10% of the rate of gluconeogenesis from propionate (Donkin and Armentano, 1995).
The data described above point to PC as a primary control point for gluconeogenesis in
developing ruminants and is supported by data from adult ruminants suggesting that PC may
be a control point for gluconeogenesis. In cattle and sheep the activity of PC is responsive to
nutritional and physiological states that impose the greatest demands for endogenous glucose
production such as lactation and feed deprivation (Greenfield et al., 2000; Velez and Donkin,
2000). In contrast, the activity of PEPCK is relatively invariant between different nutritional
and physiological states in ruminants, diabetes being the exception (Filsell et al., 1969; Taylor
et al., 1971). When both PEPCK and PC activity are examined in response to physiological
state or nutrient supply, the ratio of their activities suggests that an increase in capacity for
lactate metabolism is primarily responsible for increases in hepatic gluconeogenesis.
The dramatic reduction in basal rate of gluconeogenesis from lactate appears to be a due
to a reduction in PC activity and gene expression. Data, from sheep, examining the relationship
between prenatal development of gluconeogenic enzymes and activities found in maternal
liver fail to reveal any striking differences in activity of PC, PEPCK, or PK (Edwards et al.,
1975; Stevenson et al., 1976). Analysis of PC mRNA in liver biopsy samples from 7 through
84 days of age indicates a decline in expression of PC mRNA (Donkin et al., 1998) and
suggests a decrease in capacity for lactate metabolism. A decline in PC mRNA expression
was observed in both milk-fed calves and ruminanting calves by 84 days of age that mirrors
a reduction in gluconeogenesis from lactate (Donkin et al., 1998). Taken together, these data
suggest a developmental decrease in PC expression that is likely reflected as a decrease in lactate
recycling (Muramatsu et al., 1974) and reduced lactate metabolism to glucose in the weaned
calf (Donkin and Armentano, 1995). Data from lactating cows indicate that PC activity and
mRNA expression are induced when demands for gluconeogenesis are elevated at calving
(Greenfield et al., 2000) and during restricted feed intake (Velez and Donkin, 2000).
The onset of rumen development is marked by the production and absorption of volatile
fatty acids (VFA). Acetate and propionate form the bulk of VFA produced by rumen fermen-
tation. Acute exposure to propionate decreases gluconeogenesis from lactate equally in
hepatocytes from preruminating and ruminating calves (Donkin and Armentano, 1995).
Prolonged exposure of hepatocytes from preruminating calves to valerate (which can be
metabolized to acetate and propionate) had no effect on subsequent capacity for gluconeo-
genesis from propionate (Donkin and Armentano, 1993). However, an intermediate of propionate
metabolism, methyl malonyl-CoA, can directly inhibit lactate metabolism (Blair et al., 1973) and
is thought to be responsible for the acute effects of propionate in limiting gluconeogenesis from
lactate in bovine hepatocytes (Donkin and Armentano, 1994). At present the nature of the
developmental suppression of PC activity and gene expression is unknown.
Hepatic gluconeogenesis in developing ruminants 383

Long-term regulation of gluconeogenesis in nonruminants has been characterized by


changes in the expression of genes encoding glucoregulatory enzymes (Pilkis and Claus,
1991). It is well established that insulin represses and glucagon (or cAMP) and glucocorti-
coids induce the activity of the PEPCK enzyme by directly regulating expression of the
PEPCK gene (reviewed in O’Brien and Granner, 1990). From control strength studies in rats,
gluconeogenesis from lactate is distributed between pyruvate kinase and the reactions involving
PC and PEPCK (Sistare and Haynes, 1985). Glucocorticoids have little effect on flux through the
PK-catalyzed reaction; therefore an increase in gluconeogenesis from lactate in glucocorticoid-
treated rats or hepatocytes is mainly due to the combined increases in flux through reactions
catalyzed by PC and PEPCK (Jones et al., 1993).

7. GLUCOCORTICOIDS AND POSTNATAL DEVELOPMENT


Uncomplicated neonatal growth depends on maturation of vital organs and critical metabolic
pathways including lung and cardiac development, thyroid axis, somatotropic axis, initiation
of thermogenesis, and control of glucose homoeostasis. In altricial species, such as rodents,
a rise in cortisol at birth is necessary to initiate neonatal maturation of many of the critical
metabolic pathways including gluconeogenesis (Dalle et al., 1985; Gluckman et al., 1999). In
ruminants, the concentration of fetal cortisol usually exceeds maternal cortisol concentra-
tions; therefore caution should be exercised when extending data on the effects of
glucocorticoids from rodent studies to the biology of liver metabolism in neonatal ruminants.
The central role of glucocorticoids in regulation of expression of PEPCK, G-6-P, PC, and
CPT-I (Jitrapakdee and Wallace, 1999; Van Schaftingen and Gerin, 2002) is established and
there are indications in nonruminants that these effects are mediated through peroxisome
proliferator-activated receptor γ coactivator-1 (Louet et al., 2002). Cortisol injected into
developing sheep fetuses induced activity of hepatic G-6-Pase, fructose-6-phosphatase, PC,
and PEPCK by 2- to 3-fold (Fowden et al., 1993). Glucocorticoids may also play a more gen-
eral role in switching the fetal physiological state to a postnatal state (Liggins, 1977; Fowden,
1995). For example, gastrointestinal tract developmental, gastrin secretion, and intestinal
absorption of immunoglobulins are stimulated by cortisol in neonatal piglets and play a role
in maturation of the fetal exocrine pancreas of pigs and lambs (Sangild, 2001).
Glucocorticoids are important regulators of the glucose status after birth in the immature
neonatal calf. Cortisol concentrations decreased after birth in neonates (Baumrucker and
Blum, 1994; Hadorn et al., 1997; Hammon and Blum, 1998). Importantly, plasma cortisol
concentrations depend on the level and source of nourishment (milk or colostrum) after birth
(Hammon and Blum, 1998). Calves fed milk replacer from birth were characterized by higher
plasma cortisol concentrations and lower plasma glucagon concentrations than calves fed
colostrum (Hammon and Blum, 1998).
The prepartum cortisol surge may play an important role in initiating the perinatal switch of
the somatotropic axis from the fetal to the postnatal status and function (Breier et al., 2000).
Glucocorticoids stimulate gluconeogenesis in vivo by increasing plasma glucagon concentrations
as well as augmenting the effects of glucagon to stimulate gluconeogenesis (Marco et al., 1973;
Wise et al., 1973; Lecavalier et al., 1990). Furthermore, glucocorticoid treatment induces insulin
resistance in late gestation in sheep (Challis et al., 2001) and postnatally in humans (Weinstein
et al., 1995; Dirlewanger et al., 2000) and in dairy cows (Maciel et al., 2001). The interaction of
glucocorticoids and growth hormone has not been fully characterized for neonatal ruminants, but
postnatal growth is characterized by changes from a substrate-limited prenatal growth to enteral
feeding with the somatotropic axis becoming the dominant endocrine regulatory system.
384 S. S. Donkin and H. Hammon

Glucocorticoids enhance the maturation of the somatotropic axis and the prepartum corti-
sol surge may play an important role in initiating the perinatal switch of the somatotropic axis
from the fetal to the postnatal status and function (Breier et al., 2000; Carroll et al., 2000).
Cortisol acts to stimulates hepatic growth hormone receptor (GHR) numbers and IGF-I mRNA
levels in the sheep fetus (Li et al., 1996). In vivo studies using porcine hepatocytes indicate
that IGF-I mRNA expression is more responsive to GH in the presence of dexamethasone and
thyroxine (Brameld et al., 1999). Therefore, it might be speculated that in precocious species
such as ruminants, elevated cortisol levels at birth serve to enhance postnatal maturation of
the somatotropic axis. The somatotropic axis in neonatal calves is functional, but immature at
birth (Hammon and Blum, 1997) owing partly to reduced GH-binding capacity of the liver in
neonatal calves (Breier et al., 1994). Little is known about the ontogeny of the growth hormone
receptor in the neonatal bovine or its coordination with other hepatic functions including glucose
metabolism. Growth hormone as well as the GH receptor are present in the bovine fetus but
growth hormone does not affect IGF-I production in the liver (Gluckman et al., 1999), perhaps
owing to GHR numbers, activity of receptors, or both (Fowden, 1995; Freemark, 1999).

8. GLUCONEOGENESIS AND REGULATION OF GENE EXPRESSION


Most of the enzymes for gluconeogenesis, including PC, PEPCK-M, fructose 1,6-bisphosphatase,
and G-6-Pase, have substantial activity in near-term fetuses of ruminants (Edwards et al.,
1975; Stevenson et al., 1976; Narkewicz et al., 1993). The classic work of Ballard and Hanson
(1967) established PEPCK-C as the limiting step in development of gluconeogenesis in rats.
These data have been substantiated for rabbit and other species (Girard et al., 1992), but there
is no limitation in development of PEPCK-C activity in ruminant liver (Edwards et al., 1975;
Stevenson et al., 1976; Narkewicz et al., 1993). In rodents the rapid increase in PEPCK-C is
linked to the process of birth rather than fetal age (Girard et al., 1992) and is related to the late
fetal appearance of developmentally regulated transcription factors such as CCAAT/
enhancer-binding protein (Cassuto et al., 1999). Therefore it would follow that these tran-
scription factors or their functional homologs are likely to be present in utero in liver of the
developing ruminants.
The expression of PC is tissue-specific with the highest catalytic activity of the enzyme
found in liver, kidney, adipose tissue, brain, adrenal gland, and lactating mammary tissue.
Changes in PC abundance, through alteration in rate of synthesis, constitute long-term regula-
tion of pyruvate metabolism for gluconeogenesis and lipogenesis (Barritt, 1985). Short-term
allosteric regulation of PC activity by acetyl-CoA is well noted; however, sustained changes in
the activity of the PC enzyme require parallel increases in PC mRNA (Zhang et al., 1993).
Northern analysis of total RNA indicates the presence of a single 4.2 kb mRNA for rat and
human PC that is the product of a single copy gene (Jitrapakdee et al., 1996). However, selec-
tive amplification of the 5′ untranslated region (UTR) of PC cDNA indicates the presence of
five alternative forms of PC cDNA that are generated through differential splicing of RNA
transcripts and use of two tissue-specific promoters (Jitrapakdee et al., 1996). Transcripts
generated from the proximal promoter are restricted to gluconeogenic and lipogenic tissue
whereas those generated from the distal promoter are expressed in several tissues. These
5′ UTR isomers of the PC primary transcript share the same open reading frame and result in
one PC protein.
The liver expresses the C, D, and E forms of PC transcript, although the C and D forms
predominate. During the suckling to weaning transition the abundance of the C isoform
decreases as does PC mRNA and enzyme activity (Jitrapakdee et al., 1998). The fact that
Hepatic gluconeogenesis in developing ruminants 385

the C transcript is more functionally potent (2× greater than D) in translation reactions
suggests that a small increase in the C transcript may result in proportionately greater increase
in PC activity than changes in the D form. Furthermore an increase in C transcript and an
offsetting decrease in the D transcript would result in no net change in PC mRNA abundance
by Northern analysis but would increase PC translation and maximal PC activity.
The entire coding sequence of bovine PC has been cloned (Agca et al., 2000) and the
coding sequence contains 3075 bases with 85% identity to human PC. Furthermore, bovine
PC is expressed as six 5′ UTR variants of different lengths (Agca and Donkin, 2001).
Experiments are ongoing to test the functional significance of bovine PC variants relative to
gluconeogenesis and neonatal development in cattle.
Regulation of PEPCK expression has been extensively studied in liver of rodents as well
as rat and human hepatoma cell lines. Glucagon acting in the presence of dexamethasone is
one of the primary stimulators of PEPCK gene expression. The activity of PEPCK-C is deter-
mined by the rate of transcription of the PEPCK-C gene and the rate of turnover of its mRNA
whereas the activity of PEPCK-M appears to be constitutive (Hanson and Reshef, 1997). The
coding sequence for bovine PEPCK-C and a fragment of bovine PEPCK-M have been cloned
recently (Agca et al., 2002). Unlike PEPCK-C the expression of PEPCK-M mRNA is not
responsive to changes in physiological state (Greenfield et al., 2000; Agca et al., 2002).
Control of PEPCK-C activity is largely exerted through transcription of the gene through
activation of basal, tissue-specific, and hormone-dependent promoter elements within the
5′ region of the PEPCK-C gene (Hanson and Reshef, 1997). Crucial liver control elements
are located within −460 to +73 of the promoter and six primary protein-binding sites have
been characterized by DNAse I footprinting; these six sites contain docking sites for at least
15 separate transcription factors (Hanson and Reshef, 1997). The cAMP response element I
(CRE-I) acting synergistically with protein-binding sites 3 and 4 is primarily responsible for
the cAMP-mediated increase in PEPCK-C transcription (Hanson and Reshef, 1997). Insulin
counteracts the effects of cAMP by repressing the promoter, perhaps by blocking the ability
of glucocorticoids to promote activity of accessory factor-2 (O’Brien and Granner, 1990).
Although the PEPCK-C gene is generally thought to be transcriptionally controlled, there is
regulation through stability of the PEPCK-C mRNA which is mediated through cAMP action
on a 3′ noncoding sequence (Lemaigre and Rousseau, 1994). There is some indication that
PEPCK-C expression may be inhibited directly by glucose as is the case with other insulin-
responsive genes. The lack of appreciable glucokinase activity in ruminant liver leads to
questions regarding a similar control in ruminants.

9. FUTURE PERSPECTIVES
There is no question that gluconeogenesis is critical to the survival and normal development
of fetal, neonatal, and postnatal ruminants. The precocious development of gluconeogenic
machinery in the liver of the ruminant fetus provides a number of advantages for survival at
birth. There are many developmental aspects of gluconeogenesis that have been described in
detail for nonruminants that are applicable to the developing ruminant, but several processes
are species-specific. Information is lacking on the basic biology that accompanies the onset
of metabolic competence in the developing ruminant, including processes that may modulate
gluconeogenesis, and in many cases parallels must be drawn from rodent models. Issues asso-
ciated with the initiation of expression of key genes for gluconeogenesis including PC and
PEPCK in ruminants and the molecular cues that initiate development of gluconeogenesis
remain to be clarified. Several aspects of gluconeogenesis in developing ruminants have been
386 S. S. Donkin and H. Hammon

identified, but a more complete characterization of fetal and neonatal gluconeogenesis is


needed to identify unique regulatory controls including the molecular and biochemical events
that accompany the postnatal reduction in gluconeogenesis from lactate. Conversely the bio-
chemical anomalies identified for ruminants, such as the inherent lack of hepatic G-6-Pase
activity, could provide unique opportunities to study glucose trafficking in liver in order to
better understand metabolic diseases of humans.

REFERENCES
Agca, C., Bidwell, C.A., Donkin, S.S., 2000. Cloning of two alternately spliced transcripts of bovine
hepatic pyruvate carboxylase. FASEB J. 14, A515.
Agca, C., Donkin, S.S., 2001. Pyruvate carboxylase mRNA 5′UTR variants are heterogeneously expressed
in bovine tissues except two variants specific to adipose tissue and liver. J. Dairy Sci. 84, Suppl. 1, 386.
Agca, C., Greenfield, R.B., Hartwell, J.R., Donkin, S.S., 2002. Cloning and characterization of bovine
cytosolic and mitochondrial phosphoenolpyruvate carboxykinase and characterization during the
transition to lactation. Physiol. Genomics 11, 53–63.
Aiello, R.J., Armentano, L.E., 1987. Gluconeogenesis in goat hepatocytes is affected by calcium, ammo-
nia and other key metabolites but not primarily through cytosolic redox state. Comp. Biochem.
Physiol. B. 88, 193–201.
Aiston, S., Trinh, K.Y., Lange, A.J., Newgard, C.B., Agius, L., 1999. Glucose-6-phosphatase overexpres-
sion lowers glucose 6-phosphate and inhibits glycogen synthesis and glycolysis in hepatocytes without
affecting glucokinase translocation: evidence against feedback inhibition of glucokinase. J. Biol.
Chem. 274, 24559–24566.
Apatu, R.S., Barnes, R.J., 1991a. Release of glucose from the liver of fetal and postnatal sheep by portal
vein infusion of catecholamines or glucagon. J. Physiol. 436, 449–468.
Apatu, R.S., Barnes, R.J., 1991b. Blood flow to and the metabolism of glucose and lactate by the liver
in vivo in fetal, newborn and adult sheep. J. Physiol. 436, 431–447.
Aynsley-Green, A., 1988. Metabolic and endocrine interrelationships in the human fetus and neonate: an
overview of the control of the adaptation to postnatal nutrition, In: Lindblad, B.S. (Ed.), Perinatal
Nutrition. Academic Press, New York, pp. 161–191.
Ballard, F.J., Hanson, R.W., 1967. Phosphoenolpyruvate carboxykinase and pyruvate carboxylase in
developing rat liver. Biochem J. 104, 866–871.
Ballard, F.J., Oliver, I.T., 1963. Glycogen metabolism in embryonic chick and neonatal rat liver. Biochim.
Biophys. Acta 71, 578–588.
Barnes, R.J., 1997. Perinatal carbohydrate metabolism and the blood flow of the fetal liver. Equine Vet.
J., Suppl., 24, 26–31.
Barritt, G.J., 1985. Regulation of enzymatic activity. In: Keech, D.B., Wallace, J.C. (Eds.), Pyruvate
Carboxylase. CRC Press, Boca Raton, FL, pp. 179–248.
Baumrucker, C.R., Blum, J.W., 1994. Effects of dietary recombinant human insulin-like growth factor on
concentrations of hormones and growth factors in the blood of newborn calves. J. Endocrinol. 140, 15–21.
Blair, J.B., Cook, D.E., Lardy, H.A., 1973. Interaction of propionate and lactate in perfused rat liver.
J. Biol. Chem. 248, 3608–3615.
Blum, J.W., Hammon, H., 1999a. Pancreatic hormones (insulin and glucagon) in calves: ontogenetic
changes and nutritional effects. In: Pierzynowski, S.G., Zabielski, R. (Eds.), Biology of the Pancreas
in Growing Animals. Elsevier, Amsterdam, pp. 27–43.
Blum, J.W., Hammon, H., 1999b. Endocrine and metabolic aspects in milk-fed calves. Domest. Anim.
Endocrinol. 17, 219–230.
Boyd, R.D., Morris, F.H. Jr., Meschia, G., Makowski, E.L., Battaglia, F.C., 1973. Growth of glucose and
oxygen uptakes by fetuses of fed and starved ewes. Amer. J. Physiol. 225, 897–902.
Brameld, J.M., Gilmour, R.S., Buttery, P.J., 1999. Glucose and amino acids interact with hormones to
control expression of insulin-like growth factor-I and growth hormone receptor mRNA in cultured pig
hepatocytes. J. Nutr. 129, 1298–1306.
Breier, B.H., Ambler, G.R., Sauerwein, H., Surus, A., Gluckman, P.D., 1994. The induction of hepatic
somatotrophic receptors after birth in sheep is dependent on parturition-associated mechanisms.
J. Endocrinol. 141, 101–108.
Hepatic gluconeogenesis in developing ruminants 387

Breier, B.H., Gluckman, P.D., Bass, J.J., 1988. Plasma concentrations of insulin-like growth factor-I and
insulin in the infant calf: ontogeny and influence of altered nutrition. J. Endocrinol. 119, 43–50.
Breier, B.H., Oliver, M.H., Gallaher, B.W., 2000. Regulation of growth and metabolism during postnatal
development. In: Cronjé, P.B. (Ed.), Ruminant Physiology: Digestion, Metabolism, Growth and
Reproduction. CABI Publishing, New York, pp. 187–204.
Burchell, A., 1994. Hepatic microsomal glucose transport. Biochem. Soc. Trans. 22, 658–663.
Burchell, A., 1998. A re−evaluation of GLUT 7. Biochem. J. 331, 973.
Carroll, J.A., Daniel, J.A., Keisler, D.H., Matteri, R.L., 2000. Postnatal function of the somatotrophic
axis in pigs born naturally or by caesarian section. Domest. Anim. Endocrinol. 19, 39–52.
Cassuto, H., Aran, A., Cohen, H., Eisenberger, C.L., Reshef, L., 1999. Repression and activation of
transcription of phosphoenolpyruvate carboxykinase gene during liver development. FEBS Lett. 457,
441–444.
Challis, J.R.G., Sloboda, D., Matthews, S.G., Holloway, A., Alfaidy, N., Patel, F.A., Whittle, W., Fraser, M.,
Moss, T.J.M., Newnham, J., 2001. The fetal placental hypothalamic-pituitary-adrenal (HPA) axis,
parturition and post natal health. Mol. Cell. Endocrinol. 185, 135–144.
Chandler, K.D., Leury, B.J., Bird, A.R., Bell, A.W., 1985. Effects of undernutrition and exercise during
late pregnancy on uterine, fetal and uteroplacental metabolism in the ewe. Brit. J. Nutr. 53, 625–635.
Chatelain, F., Pegorier, J.P., Minassian, C., Bruni, N., Tarpin, S., Girard, J., Mithieux, G., 1998.
Development and regulation of glucose-6-phosphatase gene expression in rat liver, intestine, and
kidney: in vivo and in vitro studies in cultured fetal hepatocytes. Diabetes 47, 882–889.
Chow, J.C., Jesse, B.W., 1992. Interactions between gluconeogenesis and fatty acid oxidation in isolated
sheep hepatocytes. J. Dairy Sci. 75, 2142–2148.
Dalle, M., Pradier, P., Delost, P., 1985. The regulation of glucocorticosteroid secretion during the
perinatal period. Reprod. Nutr. Dev. 25, 977–991.
Dhanotiya, R.S., Bhardwaj, R., 1988. Sequential development of enzymes of gluconeogenesis and
glucose synthesis in fetal goat liver. Biomed. Biochim. Acta 47, 805–808.
Dirlewanger, M., Schneiter, P., Paquot, N., Jequier, E., Rey, V., Tappy, L., 2000. Effects of glucocorticoids
on hepatic sensitivity to insulin and glucagon in man. Clin. Nutr. 19, 29–34.
Donkin, S.S., Armentano, L.E., 1993. Preparation of extended in vitro cultures of bovine hepatocytes that
are hormonally responsive. J. Anim. Sci. 71, 2218–2227.
Donkin, S.S., Armentano, L.E., 1994. Regulation of gluconeogenesis by insulin and glucagon in the
neonatal bovine. Amer. J. Physiol. 266, R1229–R1237.
Donkin, S.S., Armentano, L.E., 1995. Insulin and glucagon regulation of gluconeogenesis in prerumi-
nating and ruminating bovine. J. Anim. Sci. 73, 546–551.
Donkin, S.S., Bertics, S.J., Armentano, L.E., 1997. Chronic and transitional regulation of gluconeo-
genesis and glyconeogenesis by insulin and glucagon in neonatal calf hepatocytes. J. Anim. Sci. 75,
3082–3087.
Donkin, S.S., Black, D.S., Greenfield, R.B., Agca, C., 1998. Effects of age, gluconeogenic substrates, and
diet on hepatic gluconeogenesis and pyruvate carboxylase expression. J. Dairy Sci. 81, Suppl. 1, 137.
Edwards, E.M., Dhand, U.K., Jeacock, M.K., Shepard, D.A.L., 1975. Activities of enzymes concerned
with pyruvate and oxaloacetate metabolism in the heart and liver of developing sheep. Biochim.
Biophys. Acta 399, 217–227.
Egli, C.P., Blum, J.W., 1998. Clinical, haematological, metabolic and endocrine traits during the first three
months of life of suckling Simmentaler calves held in a cow-calf operation. J. Vet. Med. 45, 99–118.
Ferre, P, Decaux, J.F., Issad, T., Girard, J., 1986. Changes in energy metabolism during the suckling and
weaning period in the newborn. Reprod. Nutr. Dev. 26, 619–631.
Filsell, O.H., Jarrett, I.G., Taylor, P.H., Keech, D.B., 1969. Effects of fasting, diabetes and glucocorti-
coids on gluconeogenic enzymes in the sheep. Biochim. Biophys. Acta 184, 54–63.
Fisher, D.A., Dussault, J.H., Sack, J., Chapra, I.J., 1977. Ontogenesis of hypothalamic-pituitary-thyroid
function and metabolism in man, sheep, and rat. Recent Prog. Horm. Res. 33, 59–116.
Fowden, A.L., 1995. Endocrine regulation of fetal growth. Reprod. Fertil. Dev. 7, 351–363.
Fowden, A.L., Mapstone, J., Forhead, A.J., 2001. Regulation of glucogenesis by thyroid hormones in
fetal sheep during late gestation. J. Endocrinol. 170, 461–469.
Fowden, A.L., Mijovic, J., Silver, M., 1993. The effects of cortisol on hepatic and renal gluconeogenic
enzyme activities in the sheep fetus during late gestation. J. Endocrinol. 137, 213–222.
Freemark, M., 1999. The fetal adrenal and the maturation of the growth hormone and prolactin axis.
Endocrinology 140, 1963–1965.
388 S. S. Donkin and H. Hammon

Ganguli, S., Sinha, M.K., Sterman, B., Harris, P., Sperling, M.A., 1983. Ontogeny of hepatic insulin and
glucagon receptors and adenylate cyclase in rabbit. Amer. J. Physiol. 244, E624–E631.
Girard, J., 1986. Gluconeogenesis in late fetal and early neonatal life. Biol. Neonate 50, 237–258.
Girard, J., 1990. Metabolic adaptations to change of nutrition at birth Biol. Neonate 58, Suppl. 1, 3–15.
Girard, J., Sperling, M., 1983. Glucagon in the fetus and the newborn. In: Lefebre, P.J. (Ed.), Glucagon:
Handbook of Experimental Pharmacology. Springer-Verlag, New York, p. 66.
Girard, J., Ferré, P., Pégorier, J.P., Duée, P.H., 1992. Adaptions of glucose and fatty acid metabolism
during perinatal period and suckling-weaning transition. Physiol. Rev. 72, 507–562.
Gluckman, P.D., Sizonenko, S.V., Bassett, N.S., 1999. The transition from fetus to neonate – an endocrine
perspective. Acta Pædiat. 428, 7–11.
Greenfield, R.B., Cecava, M.J., Donkin, S.S., 2000. Changes in mRNA expression for gluconeogenic
enzymes in liver of dairy cattle during the transition to lactation. J. Dairy Sci. 83, 1228–1236.
Gresham, E.L., James, E.J., Raye, J.R., Battaglia, F.C., Makowski, E.L., Meschia, G., 1972. Production
and excretion of urea by the fetal lamb. Pediatrics 50, 372–379.
Hadorn, U., Hammon, H., Bruckmaier, R.M., Blum, J.W., 1997. Delaying colostrum intake by one day
has important effects on metabolic traits and on gastrointestinal and metabolic hormones in neonatal
calves. J. Nutr. 127, 2011–2023.
Hamada, T., Matsumoto, M., 1984. Effects of nutrition and ontogeny on liver cytosolic and mitochondr-
ial phosphoenolpyruvate carboxykinase activity of the rat, hamster, guinea-pig, pig, kid, calf and
chick. Comp. Biochem. Physiol. B. 77, 547–550.
Hammon, H., Blum, J.W., 1997. The somatotropic axis in neonatal calves can be modulated by nutrition,
growth hormone, and long-R3-IGF-I. Amer. J. Physiol. 273, E130–E138.
Hammon, H.M., Blum, J.W., 1998. Metabolic and endocrine traits of neonatal calves are influenced by
feeding colostrum for different durations or only milk replacer. J. Nutr. 128, 624–632.
Hanson, R.W., Reshef, L., 1997. Regulation of phosphoenolpyruvate carboxykinase (GTP) gene expres-
sion. Annu. Rev. Biochem. 66, 581–611.
Hay, W.W. Jr., Sparks, J.W., Wilkening, R.B., Battaglia, F.C., Meschia, G., 1984. Partition of maternal
glucose production between conceptus and maternal tissues in sheep. Amer. J. Physiol. 245,
E347–E350.
Hod, Y., Cook, J.S., Weldon, S.L., Short, J.M., Wynshaw-Boris, A., Hanson, R.W., 1987. Differential
expression of the genes for the mitochondrial and cytosolic forms of phosphoenolpyruvate carboxyk-
inase. Ann. N.Y. Acad. Sci. 478, 31–35.
Hodgson, J.C., Mellor, D.J., Field, A.C., 1982. Foetal and maternal rates of urea production and disposal
in well-nourished and undernourished sheep. Brit. J. Nutr. 48, 49–58.
Holcomb, T., Curthoys, N.P., Gstraunthaler, G., 1995. Subcellular localization of PEPCK and metabolism
of gluconeogenic substrains of renal cell lines. Amer. J. Physiol. 268, C449–C457.
Hosokawa, M., Thorens, B., 2002. Glucose release from GLUT2-null hepatocytes: characterization of
a major and a minor pathway. Amer. J. Physiol. Endocrinol. Metab. 282, E794–E801.
Howarth, R.E., Baldwin, R.L., Ronning, M., 1968. Enzyme activities in liver, muscle, and adipose tissue
of calves and steers. J. Dairy Sci. 51, 1270–1274.
Huntington, G.B., 1990. Energy metabolism in the digestive tract and liver of cattle: influence of
physiological state and nutrition. Reprod. Nutr. Dev. 30, 35–47.
Jitrapakdee, S., Booker, G.W., Cassady, A.I., Wallace, J.C., 1996. Cloning, sequencing and expression of
rat liver pyruvate carboxylase. Biochem. J. 316, 631–637.
Jitrapakdee, S., Wallace, J.W., 1999. Structure and regulation of pyruvate carboxylase. Biochem. J.
340, 1–16.
Jitrapakdee, S., Gong, Q., MacDonald, M.J., Wallace, J.C., 1998. The rat pyruvate carboxylase gene
structure: alternate promoters generate multiple transcripts with the 5’-end heterogeneity. J. Biol.
Chem. 272, 20522–20530.
Jones, C.G., Hothi, S.K., Titheradge, M.A., 1993. Effect of dexamethasone on gluconeogenesis, pyruvate
kinase, pyruvate carboxylase and pyruvate dehydrogenase flux in isolated hepatocytes. Biochem. J.
289, 821–830.
Kalhan, S., Parimi, P., 2000. Gluconeogenesis in the fetus and neonate. Semin. Perinatol. 24, 94–106.
Lecavalier, L., Bolli, G., Gerich, G., 1990. Glucagon-cortisol interactions on glucose turnover and lactate
gluconeogenesis in normal humans. Amer. J. Physiol. 258, E569–E575.
Lemaigre, F.P., Rousseau, G.G., 1994. Transcriptional control of genes that regulate glycolysis and glu-
coneogenesis in adult liver. Biochem. J. 303, 1–14.
Hepatic gluconeogenesis in developing ruminants 389

Leury, B.J., Bird, A.R., Chandler, K.D., Bell A.W., 1990. Glucose partitioning in the pregnant ewe:
effects of undernutrition and exercise. Brit. J. Nutr. 64, 449–462.
Li, J., Owens, J.A., Owens, P.C., Saunders, J.C., Fowden, A.L., Gilmour, R.S., 1996. The ontogeny of
hepatic growth hormone receptor and insulin-like growth factor I gene expression in the sheep fetus
during late gestation: developmental regulation by cortisol. Endocrinology 137, 1650–1657.
Liggins, G.C., 1977. The role of cortisol in preparing the fetus for birth. Reprod. Fertil. Dev. 6, 141–150.
Louet, J.F., Hayhurst, G., Gonzalez, F.J., Girard, J., Decaux, J.F., 2002. The coactivator PGC-1 is
involved in the regulation of the liver carnitine palmitoyltransferase I gene expression by cAMP in
combination with HNF4 alpha and cAMP-response element-binding protein (CREB). J. Biol. Chem.
277, 37991–38000.
Maciel, S.M., Chamberlain, C.S., Wettemann, R.P., Spicer, L.J., 2001. Dexamethasone influences
endocrine and ovarian function in dairy cattle. J. Dairy Sci. 84, 1998–2009.
Marco, J., Calle, C., Román, D., Díaz-Fierros, M., Villanueva, M.L., Valverde, I., 1973. Hyperglucagonism
induced by glucocorticoid treatment in man. N. Engl. J. Med. 288, 128–131.
Mashek, D.G., Bertics, S.J., Grummer, R.R., 2002. Metabolic fate of long-chain unsaturated fatty acids
and their effects on palmitic acid metabolism and gluconeogenesis in bovine hepatocytes. J. Dairy
Sci. 85, 2283–2289.
Muramatsu, M., Sugawara, M., Tsuda, T., 1974. Changes in rates of recycling and turnover of glucose
during development of sheep. Agr. Biol. Chem. 38, 259–267.
Narkewicz, M.R., Carver, T.D., Hay, W.W. Jr., 1993. Induction of cytosolic phosphoenolpyruvate car-
boxykinase in the ovine fetal liver by chronic fetal hypoglycemia and hypoinsulinemia. Pediat. Res.
33, 493–496.
Neufeld, N.D., Scott, M., Kaplan, S.A., 1980. Ontogeny of the mammalian insulin receptor: studies of
human and rat fetal liver plasma membranes. Dev. Biol. 78, 151–160.
O’Brien, R.M., Granner, D.K., 1990. PEPCK gene as model of inhibitory effects of insulin on gene tran-
scription. Diabetes Care 13, 327–339.
Park, E.A., Song, S., Olive, M., Roesler, W.J., 1997. CCAAT-enhancer-binding protein alpha (C/EBP
alpha) is required for the thyroid hormone but not the retinoic acid induction of phosphoenolpyruvate
carboxykinase (PEPCK) gene transcription. Biochem. J. 322, 343–349.
Pearce, J., Unsworth, E.F., 1980. The effects of diet on some hepatic enzyme activities in the pre-ruminant
and ruminating calf. J. Nutr. 110, 255–261.
Pilkis, S.J., Claus, T.H., 1991. Hepatic gluconeogenesis/glycolysis: regulation and structure/function
relationships of substrate cycle enzymes. Annu. Rev. Nutr. 11, 465–515.
Prior, R.L., 1980. Glucose and lactate metabolism in vivo in ovine fetus. Amer. J. Physiol. 239, E208–E214.
Prior, R.L., 1982. Gluconeogenesis in the ruminant fetus: evaluation of conflicting evidence from radio-
tracer and other experimental techniques. Fed. Proc. 41, 117–122.
Prior, R.L., Scott, R.A., 1977. Ontogeny of gluconeogenesis in the bovine fetus: influence of maternal
dietary energy. Dev. Biol. 58, 384–393.
Sangild, P.T., 2001. Transitions in the life of the gut at birth, In: Lindberg, J.E., Ogle, B. (Eds.), Digestive
Physiology of Pigs. CABI Publishing, New York, pp. 3–17.
Savan, P.M., Jeacock, M.K., Shepherd, D.A., 1986. Gluconeogenesis in foetal, suckling and weaned
lambs. J. Agr. Sci. 106, 259–268.
Seoane, J., Trinh, K., O’Doherty, R.M., Gomez-Foix, A.M., Lange, A.J., Newgard, C.B., Guinovart, J.J.,
1997. Metabolic impact of adenovirus-mediated overexpression of the glucose-6-phosphatase
catalytic subunit in hepatocytes. J. Biol. Chem. 272, 26972–26977.
Sistare, F.D., Haynes, R.C., 1985. The interaction between the cytosolic pyridine nucleotide redox potential
and gluconeogenesis from lactate/pyruvate in isolated rat hepatocytes: implications for investigations
of hormone action. J. Biol. Chem. 260, 12748–12753.
Sperling, M.A., Grajwer, L., Leake, R.D., Fisher, D.A., 1977. Effects of somatostatin (SRIF) infusion on glu-
cose homeostasis in newborn lambs: evidence for a significant role of glucagon. Pediat. Res. 11, 962–967.
Stevenson, R.E., Morris, F.H., Jr., Adcock, E.W. III, Howell, R.R., 1976. Development of gluconeogenic
enzymes in fetal sheep liver and kidney. Dev. Biol. 52, 167–172.
Swiatek, K.R., 1971. Development of gluconeogenesis in pig liver slices. Biochim. Biophys. Acta 252,
274–279.
Taylor, P.H., Wallace, J.C., Keech, D.B., 1971. Gluconeogenic enzymes in sheep liver: intracellular local-
ization of pyruvate carboxylase and phosphoenolpyruvate carboxykinase in normal, fasted and
diabetic sheep. Biochim. Biophys. Acta 237, 179–191.
390 S. S. Donkin and H. Hammon

Tsoulos, N.G., Colwill, J.R., Battaglia, F.C., Makowski, E.L., Meschia, G., 1971. Comparison of glucose,
fructose, and O2 uptakes by fetuses of fed and starved ewes. Amer. J. Physiol. 221, 234–237.
Van Schaftingen, E., Gerin, I., 2002. The glucose-6-phosphatase system. Biochem. J. 362, 513–532.
Velez, J.C., Donkin, S.S., 2000. Administration of bST elevates phosphoenolpyruvate carboxykinase
mRNA in lactating dairy cows. J. Anim. Sci. 78, Suppl. 1, 144.
Warnes, D.M., Seamark, R.F., Ballard, F.J., 1977. Metabolism of glucose, fructose and lactate in vivo in
chronically cannulated foetuses and in suckling lambs. Biochem. J. 162, 617–626.
Watford, M., Hod, Y., Chiao, Y.B., Utter, M.F., Hanson, R.W., 1981. The unique role of the kidney in
gluconeogenesis in the chicken: the significance of a cytosolic form of phosphoenolpyruvate carboxyki-
nase. J. Biol. Chem. 256, 10023–10027.
Weinstein, S.P., Paquin, T., Pritsker, A., Haber, R.S., 1995. Glucocorticoid-induced insulin resistance:
dexamethasone inhibits the activation of glucose transport in rat skeletal muscle by both insulin- and
non-insulin-related stimuli. Diabetes 44, 441–445.
Williamson, J.R., Kreisberg, R.A., Felts, P.W., 1966. Mechanism for the stimulation of gluconeogenesis
by fatty acids in perfused rat liver. Proc. Natl. Acad. Sci. USA 56, 247–254.
Wise, J.K., Hendler, R., Felig, P., 1973. Influence of glucocorticoids on glucagon secretion and plasma
amino acid concentrations in man. J. Clin. Invest. 52, 2774–2782.
Zhang, J., Xia, W., Brew, K., Ahmad, F., 1993. Adipose pyruvate carboxylase: amino acid sequence and
domain structure deduced from cDNA sequencing. Proc. Natl. Acad. Sci. USA 90, 1766–1770.
16 Energy metabolism in the developing
rumen epithelium

B. W. Jesse

Department of Animal Science, Rutgers, The State University of New Jersey,


84 Lipman Drive, New Brunswick, NJ 08901-8525, USA

The physical changes occurring during rumen epithelial development have been extensively
characterized. However, relatively little information is available concerning development of
energy metabolism in rumen epithelium. Available data indicate that both ontogenic and
physiological/dietary factors are necessary for complete rumen epithelial metabolic develop-
ment. Changes in the expression of specific genes, e.g. those for ketone body production, in
response to ontogenic and physiological/dietary factors appear to be responsible for the
changes in energy metabolism in developing rumen epithelium. Future research efforts will
need to identify the mechanisms regulating gene expression within the developing rumen
epithelium to obtain a better understanding of this process.

1. INTRODUCTION
Energy metabolism in the rumen epithelium of mature sheep and cattle has been extensively
characterized over the years. The specific oxidizable substrates required for energy produc-
tion in the rumen epithelium in neonates of these species have received some attention, while
the establishment of the rumen fermentation, and the physical changes occurring to the rumen
epithelium during development, have been extensively researched. However, the changes in
energy metabolism that occur during neonatal rumen epithelial metabolic development, and
most importantly the timing and control mechanisms regulating those changes, have received
relatively little attention. This review will provide a historical overview of the state of our
knowledge in this area, and will discuss in more depth recent evidence that examines metabolic
development in the neonatal rumen epithelium. It will become apparent that relatively little is
known concerning the mechanisms driving rumen metabolic development, and that this is the
result of relatively little research having been conducted in this area. The vast bulk of the
research literature examining rumen development has focused on the rumen fermentation
itself, or on physical changes manifested by changes in rumen epithelial morphology and
blood chemistry in the growing ruminant. While the focus of the review will be on rumen

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
391 © 2005 Elsevier Limited. All rights reserved.
392 B. W. Jesse

epithelium, it should be noted that developmental changes will also be occurring in the reticular
and omasal epithelia. This review will conclude with a brief discussion of key areas for future
research into the topic of rumen epithelial metabolic development.

2. Historical Perspective

To understand the changes in energy metabolism occurring in developing rumen epithelium,


it is necessary to first define the points that mark the limits of these changes, that is, the energy
metabolism present at the start (in the neonatal rumen) and at the endpoint (in the mature
rumen) of rumen development. For the purpose of this review this time period will generally
coincide with the 8-week period following birth, the time period in which the bulk of the
physical and metabolic changes occur in the developing rumen epithelium.

2.1. Energy metabolism in mature rumen epithelium

2.1.1. Substrates and their metabolism

The primary compounds that have been investigated as potential energy-yielding substrates
within the mature rumen epithelium are fatty acids (both short- and long-chain), glucose, and
glutamine (Weigand et al., 1975; White and Leng, 1980; Harmon, 1986; Harmon et al., 1991;
Jesse et al., 1992; Britton and Krehbiel, 1993; Remond et al., 1995; Baldwin and McLeod,
2000). Of these, the short-chain or volatile fatty acids (VFA) are quantitatively the most
important energy sources for the ruminal epithelium under most circumstances.
The VFA, predominantly acetate, propionate, and butyrate, are the products of the rumen
fermentation, and are absorbed by the rumen epithelium for release into the portal circulation.
Prior to release into the portal circulation, the VFA may undergo metabolism within the rumen
epithelium. Depending upon the specific VFA considered, a variable amount of metabolism
occurs (Remond et al., 1995). VFA metabolism may include either oxidation, or conversion
into other intermediates (e.g. lactate, β-hydroxybutyrate [BHBA], acetoacetate [AcAc]), for
release into the portal circulation. The activities of numerous enzymes in the glycolytic pathway,
the citric acid cycle, the ketogenic pathway, the acyl-CoA synthetases for activation of VFA,
and the various enzymes involved in VFA uptake have been determined (Young et al., 1969; Bush
and Milligan, 1971; Ash and Baird, 1973; Nocek et al., 1980; Scaife and Tichivangana, 1980;
Bush, 1982; Leighton et al., 1983; Harmon et al., 1991). These assays have generally been
conducted under saturating substrate concentrations to yield maximal activities of the assayed
enzymes. Consequently, relatively little information is available concerning the kinetic prop-
erties of rumen epithelial enzymes. Activation of VFA to the coenzyme A thioester has been
proposed as the key regulatory point for rumen epithelial VFA metabolism (Ash and Baird,
1973). However, others have indicated that knowledge of both the kinetic properties of these
enzymes as well as the tissue substrate and inhibitor concentrations within the rumen epithelium
is needed to fully justify that statement (Britton and Krehbiel, 1993).
While various researchers have reported on the effects of dietary changes (composition,
level of intake) on VFA metabolism and activities of specific enzymes within the rumen
epithelium, no consensus has yet emerged from these studies. Harmon et al. (1991) noted an
overall increase in ruminal epithelial metabolism in cattle fed at twice maintenance require-
ments than in cattle fed at maintenance. These authors also noted some increase in acyl-CoA
synthetase activities of rumen epithelium from cattle fed a high-forage diet. Some researchers
report no change in activity for a number of enzymes in rumen epithelium from cattle fed
Energy metabolism in the developing rumen epithelium 393

a high-grain versus hay diet (e.g. Young et al., 1969). Others have noted differences in
VFA transport across the rumen epithelium as well as changes in some enzyme activities
(propionyl-CoA synthetase, glutamate dehydrogenase, and aspartate aminotransferase) as a
result of changes in ration physical form and level of rumen-degradable nitrogen (Nocek
et al., 1980). A recent paper suggests that changes in rumen epithelial energy metabolism in
response to dietary energy intake and composition is due in part to changes in tissue mass
rather than to changes in metabolism per unit epithelial mass (McLeod and Baldwin, 2000).
This is in agreement with other studies noting increases in rumen epithelial mass and papil-
lae length in response to increased dietary energy intake (Liebich et al., 1987).
Of the three major VFA absorbed by rumen epithelium, the proportion of absorbed acetate
that is metabolized is lower than any of the other VFA (18–30%; Remond et al., 1995). However,
since significantly more acetate is absorbed than propionate and butyrate, the absolute amount
of acetate metabolized can be relatively large. The literature indicates that, compared to
butyrate, relatively little acetate is converted to ketone bodies (BHBA and AcAc; Remond et al.,
1995). Acetate undergoes primarily oxidation to carbon dioxide by the rumen epithelium,
thereby contributing to the energy needs of the rumen epithelium (Britton and Krehbiel, 1993).
The situation with propionate metabolism in the ruminal epithelium is the least clear of the
VFA. Propionate is not used for the synthesis of ketone bodies, but is converted primarily to
lactate, with some complete oxidation to carbon dioxide, some pyruvate formation, and some
transamination of pyruvate to alanine (Remond et al., 1995). Some studies have estimated that
as much as 70% of absorbed propionate is converted to lactate prior to release into the portal
circulation, although more recent data suggest that the proportion is much less than that (30%;
Remond et al., 1995). Propionate oxidation to carbon dioxide by ruminal epithelium is min-
imal at physiological concentrations of propionate, presumably to spare propionate and its
metabolites for hepatic gluconeogenesis (Remond et al., 1995), although at high propionate
concentrations rumen epithelium in vitro can oxidize propionate at relatively high rates
(Harmon et al., 1991). Butyrate has been noted to inhibit propionate activation to propionyl-CoA,
thereby minimizing propionate metabolism and further sparing propionate for release into the
portal circulation (Harmon et al., 1991).
Butyrate has long been known to be the VFA most extensively metabolized by the rumen
epithelium, undergoing both oxidation to carbon dioxide and conversion to BHBA and AcAc
(Bergman, 1990; Remond et al., 1995). Various researchers have estimated that as much as
90% of the absorbed butyrate undergoes metabolism by the ruminal epithelium (Bergman,
1990; Remond et al., 1995). Generally, a higher proportion of butyrate is converted to ketone
bodies, and a lower proportion to carbon dioxide, than occurs with acetate, although the
absolute rates of carbon dioxide production from acetate and butyrate are comparable
(Harmon et al., 1991). This may simply be a reflection of the relative rate of activation of
these two VFA by their respective acyl-CoA synthetases (Harmon et al., 1991). Acetyl-CoA
synthetase activity is significantly lower than either propionyl- or butyryl-CoA synthetase
activities (Harmon et al., 1991). A lower rate of acetate activation may provide sufficient
Ac-CoA for use in the TCA cycle, but not a sufficiently high concentration for use of Ac-CoA
as a ketogenic substrate, at least when acetate is the sole substrate in vitro. The kinetic param-
eters of the ruminal enzymes responsible for use of acetyl-CoA in these pathways (i.e. citrate
synthase and AcAc-CoA thiolase) are not known, however. Butyrate can inhibit acetate and
propionate activation to their respective coenzyme A thioesters, while acetate and propionate
have relatively little effect on butyrate activation (Harmon et al., 1991). These data are con-
sistent with observations of Scaife and Tichivangana (1980), who isolated a partially purified
short-chain acyl-CoA synthetase from sheep rumen epithelium. The kinetic properties of this
394 B. W. Jesse

fraction suggested the existence of two distinct enzyme activities, one specific for butyrate
activation and the other capable of activating acetate, propionate, or butyrate (Scaife and
Tichivangana, 1980).
The long-chain fatty acid palmitate may be oxidized or used in the synthesis of ketone
bodies by rumen epithelium in much the same manner as in liver (Jesse et al., 1992). Isolated
rumen epithelial cells oxidized palmitate at one-quarter the rate of butyrate, and converted
palmitate to ketone bodies at one-half the rate of butyrate (Jesse et al., 1992). Propionate,
butyrate, and ammonia inhibited ketogenesis from palmitate, but only butyrate and ammonia
inhibited palmitate oxidation (Jesse et al., 1992). These data suggest that during feed restric-
tion, or possibly when consuming a high-fat diet, mature rumen epithelium would be capable
of using long-chain fatty acids as a major energy source.
Both glucose and glutamine can undergo oxidation to carbon dioxide, and conversion to
lactate in the case of glucose, or to glutamate and alanine in the case of glutamine, within the
ruminal epithelium (Remond et al., 1995). Glucose may be a major source of the lactate
produced by rumen epithelium (Remond et al., 1995). Rates of glucose oxidation to carbon
dioxide are comparable to lactate production rates from glucose, but glucose oxidation occurs
at a lower rate than either acetate or butyrate oxidation (Harmon et al., 1991). Glutamine
oxidation rates by ruminal epithelium reportedly were 7 times lower than glucose oxidation
(Harmon et al., 1991), suggesting that glutamine is not a major energy source for ruminal
epithelium, in contrast to the importance of glutamine as an energy source to other tissues of
the digestive tract (Britton and Krehbiel, 1993). More recent data indicate that glutamine
in vitro can be oxidized by rumen epithelial cells at rates faster than butyrate, if present at
sufficiently high concentrations (50 mM; Baldwin and McLeod, 2000). However, the gluta-
mine concentration required for half-maximal oxidation rates by rumen epithelial cells
(6 mM; Baldwin and McLeod, 2000) is about 30 times greater than the glutamine concentra-
tion found in vivo (0.20 mM; Alio et al., 2000; Noziere et al., 2000; Hanigan et al., 2001).
This suggests that little glutamine oxidation would be expected to occur in vivo (Baldwin and
McLeod, 2000), as was noted by Harmon et al. (1991).

2.1.2. Substrate uptake

Prior to activation and metabolism within the rumen epithelium, energy substrates must be
transported into the epithelium. For the rumen VFA this presents a unique challenge, as at the
pH typical of rumen fluid (5.6–6.2), VFA exist predominantly in the ionized form. Ionized
VFA would be unable to diffuse through the plasma membranes of the rumen epithelial cells.
Consequently, some mechanism must exist for the movement of VFA across the epithelium.
While early data supported a transcellular rather than a paracellular mechanism, the exact
mechanism was not known (Remond et al., 1995). Early research suggested the importance
of carbonic anhydrase in the absorption of VFA by rumen epithelium (Aafjes, 1967;
Bergman, 1990). The proposed mechanism involved production of HCO 3− and H+ within the
rumen epithelium, movement of the protons and bicarbonate across the rumen mucosa and
into the rumen fluid, and neutralization of the VFA followed by passive diffusion into the
rumen mucosa down a concentration gradient (Bergman, 1990). Recent data suggest the
existence of both a carrier-mediated transport mechanism and a passive diffusion mechanism
(Sehested et al., 1999a), with the mediated transport mechanism coupled with sodium, chloride,
and bicarbonate (Sehested et al., 1999b). These results are summarized in fig. 1. While the situ-
ation for acetate and propionate is unknown, butyrate transport appears to be energy-dependent,
as inhibition of ATP synthesis in rumen epithelium blocks butyrate uptake (Gabel et al., 2001).
Energy metabolism in the developing rumen epithelium 395

Fig. 1. Diagram of volatile fatty acid (VFA) transport into rumen epithelial cells. Protons and bicarbonate
are generated within the rumen epithelial cells by the action of carbonic anhydrase. Circles represent the
presence of specific transport molecules. Based on Sehested et al. (1999a).

This suggests that ATP is either directly involved in butyrate transport, or that energy-
dependent metabolism of butyrate is necessary for butyrate uptake by rumen epithelium.
Until recently, the mechanism of glucose transport into rumen epithelium had not been
examined. The general concept was that glucose utilized by the rumen epithelium was derived
from the blood and absorbed at the serosal side of the epithelium. A recent report indicated
the presence of GLUT5 (the basolateral facilitative glucose transporter) mRNA in sheep
rumen epithelium (Zhao et al., 1998), which would perform the uptake of blood glucose by
rumen epithelium. Surprisingly, mRNA for the Na+-dependent glucose transporter (SGLT1)
was also detected. Functional analysis of 3-O-methylglucose transport by sheep rumen
epithelium in vitro demonstrated the presence of SGLT1, which was subsequently confirmed
by cloning a cDNA from rumen epithelium with 100% identity to the sheep intestinal SGLT1
(Aschenbach et al., 2000b). In vivo experiments also demonstrated the sodium-dependent
absorption of physiological concentrations of glucose by sheep rumen epithelium
(Aschenbach et al., 2000a). The authors suggested that this could be an important route of
glucose absorption in ruminants consuming high-concentrate diets, especially as a mecha-
nism to minimize the effects of rumen acidosis, as previously suggested by Ganter et al.
(1993). This hypothesis was supported by the observation that sheep rumen epithelial uptake
of glucose by SGLT-1 can be stimulated by β2-adrenoceptors, since increased sympathetic
activity has been noted in acidotic ruminants (Aschenbach et al., 2002). Glucose uptake by
mature rumen epithelium is thus more complex than previously believed.
No studies appear to have examined the mechanism for palmitate or glutamine specific uptake
by mature rumen epithelium. Alternatively, glutamine could enter the rumen epithelium as
396 B. W. Jesse

a component of peptides rather than as the amino acid (Webb et al., 1992). A specific peptide
transporter has been detected in ruminal epithelium that could move glutamine-containing
peptides into the ruminal epithelium by an electrogenic mechanism (Chen et al., 1999).
However, these results are not universally accepted (Martens et al., 2001), indicating that the
role of peptide transport in moving glutamine into the ruminal epithelium has yet to be fully
resolved.

2.2. Energy metabolism in neonatal rumen epithelium

2.2.1. Substrates and their metabolism

Much of the early work on rumen epithelial development was concerned with factors that
would promote anatomical development of the rumen. For example, the classical work of
Warner et al. (1956) was the first detailed study to examine the effect of different dietary treat-
ments on both ruminal size and papillary development. This was the first study to postulate
the importance of VFA from the microbial fermentation as inducers of rumen epithelial devel-
opment. Subsequent research (Tamate et al., 1962; Hamada et al., 1976; Klein et al., 1987)
confirmed the importance of VFA for the stimulation of papillary growth, and noted that
increased rumen volume and musculature were dependent on bulk fill of the rumen.
Relatively little research, however, examined the energy metabolism of the developing rumen
epithelium. The first study examining energy metabolism in undeveloped rumen epithelium
noted that, prior to papillary development, metabolism of VFA by rumen epithelium was low
(Sutton et al., 1963). Blood glucose was subsequently identified as the primary energy substrate
of neonatal calf rumen epithelium (Juhasz et al., 1976).
Giesecke et al. (1979) performed the first systematic analysis of the changes in rumen epithe-
lial metabolism that occur during rumen development. Using slices of rumen epithelium
isolated from weaned and unweaned lambs of various ages, these researchers measured oxygen
consumption and ketone body production by the rumen epithelial slices in vitro in the presence
of glucose, lactate, butyrate, or propionate. The importance of a number of observations that
were made by these authors has repeatedly been demonstrated in the intervening years. Oxygen
consumption by rumen epithelium decreased with age independently of dietary changes and
stage of rumen epithelial development. Glucose, lactate, and butyrate stimulated oxygen con-
sumption by rumen epithelial slices from both 2-week-old and 6-month-old lambs. However,
the ability of glucose and lactate to stimulate oxygen consumption in rumen epithelium from
6-month-old lambs was significantly less than in that from 2-week-old lambs, whereas butyrate
stimulated oxygen consumption equally well in ruminal epithelium from lambs of either age.
This was interpreted as a shift in substrate preference from glucose to VFA by the developing
rumen epithelium, and was supported by a decrease in glucose uptake by the rumen epithelium
from the older lambs. These results also implied that the rate of butyrate oxidation was not
dependent on the stage of rumen epithelial development.
Ketogenesis at different developmental stages of the rumen epithelium was also examined.
In rumen epithelial slices from 9–10-week-old lambs either maintained on milk (undeveloped
epithelium) or weaned to solid feed (developed epithelium), total ketone body production
(BHBA + AcAc) was nearly 1.60-fold greater in the developed than the undeveloped
epithelium, and was comparable to ketogenic rates observed in older lambs. There was also
a shift in the BHBA:AcAc ratio from about 2.7 to 6.2 in the undeveloped and developed lamb
rumen epithelium, respectively, indicating a shift in the redox potential of the rumen epithe-
lium with age. These data indicated the importance of solid feed intake in promoting rumen
Energy metabolism in the developing rumen epithelium 397

metabolic development. Glucose addition to these in vitro incubations stimulated ketogenesis


in a synergistic manner. Perhaps the most interesting observation, however, was the change in
ketogenic capacity with age of rumen epithelium from milk-fed lambs. Total ketone body
production (μmoles/(g tissue dry weight × hour)) by rumen epithelial slices from milk-fed
lambs increased from 19.6 at 1 week of age, to 25.5 at 3–4 weeks of age, to 71.9 at 9–10 weeks
of age. This was the first indication that changes in energy metabolism within the rumen
epithelium could occur in the absence of solid feed intake and the consequent microbial
production of VFA.
The observations of Giesecke et al. (1979) concerning ruminal ketogenesis were subse-
quently confirmed and extended to rumen epithelium from milk-fed and normally reared
(milk-fed to 28 days of age; starter and hay available after 10 days of age) calves by Bush
(1988). This author examined rumen epithelium from 3-, 12-, 19-, 30-, and 60-day-old calves,
providing a more complete time-course of the changes in ketogenesis occurring in develop-
ing rumen epithelium. Total ketone body production from butyrate by rumen epithelium from
normally reared calves was detectable at 3 days of age, and slowly increased through 19 days
of age. By 30 days of age ketogenic rates had jumped to about 40% of that observed in mature
rumen epithelium, and by 60 days of age were similar to ketogenic rates in mature rumen
epithelium. Ketogenesis from butyrate also increased with age in rumen epithelium from
milk-fed calves, although the difference in total ketone body production rate by rumen epithe-
lium from milk-fed and conventionally reared calves at 60 days of age was about 4.8-fold
(Bush, 1988), in contrast to the 1.6-fold difference observed by Giesecke et al. (1979).
Acetate conversion rate to ketone bodies was nearly 12.5-fold less than was observed with
butyrate as substrate in rumen epithelium from 60-day-old conventionally reared calves,
again similar to that in rumen epithelium from older animals (Bush, 1988).
Two important observations were made in both of these studies (Giesecke et al., 1979;
Bush, 1988). First, some changes in rumen epithelial energy metabolism occur in an onto-
genic manner even in the absence of solid feed intake and the associated rumen fermentation.
Second, complete rumen epithelial metabolic development requires solid feed intake, and is
mediated presumably by the VFA from the resultant feed fermentation.

2.2.2. Substrate uptake

The underlying mechanisms(s) responsible for the changes in glucose, butyrate, and ketone
body metabolism in the developing rumen epithelium were in general not identified by Bush
(1988) or Giesecke et al. (1979). Altered substrate uptake or metabolism, or a combination of
both, could be responsible for the observed changes in energy metabolism during rumen
epithelial development. Giesecke et al. (1979) did note a nearly 10-fold decrease in glucose
uptake in rumen epithelium from 6-month-old lambs compared to that observed in 2-week-old
lambs. (In contrast, lactate uptake by rumen epithelium from 6-month-old lambs were more
than doubled.) Whether this decrease in glucose uptake was due to a decline in glucose trans-
porter activity, or to a decrease in glucose metabolic capacity, during rumen epithelial
development is not known. Similarly, changes in butyrate metabolism by developing rumen
epithelium could be the result of changes in the activity of the appropriate transporter activity,
or in the activity of butyrate metabolizing enzymes within the ruminal epithelium. Prior to
1992, no reports had been made concerning activity changes of metabolic enzymes in devel-
oping rumen epithelium that would provide an explanation for the major metabolic changes
occurring during rumen epithelial development. There apparently was also no research into
VFA absorption by the developing rumen epithelium. Similarly, although glucose oxidation
398 B. W. Jesse

by mature rumen epithelium had been observed at that time, no research had been conducted
into the mechanism of glucose absorption by either mature or developing rumen epithelium.

3. RECENT DATA ON THE ENERGY METABOLISM OF DEVELOPING


RUMEN EPITHELIUM
Beginning in 1992, a series of papers appeared that attempted to define more finely the timing
of the energy metabolism changes occurring in developing rumen epithelium, as well as to
attempt identification of the mechanisms responsible for those changes.

3.1. Substrate oxidation

In 1992, Baldwin and Jesse reported on the developmental changes in glucose and butyrate
metabolism by rumen epithelial cells isolated from conventionally reared lambs of different
ages (0, 4, 7, 14, 28, 42, and 56 [weaned] days of age). In this study, glucose oxidation (based
on 14CO2 production from [1-14C]glucose by isolated rumen epithelial cells) increased from
birth to 14 days of age, remained elevated until 42 days of age, and decreased by weaning at
56 days to rates lower than those observed at birth, but comparable to those observed in mature
sheep (Baldwin and Jesse, 1992). Maximum glucose oxidation rates coincided with the time
period of allometric rumen growth, suggesting the importance of glucose oxidation for energy
generation during this time of rapid rumen tissue accretion. Surprisingly in this study, butyrate
oxidation rates (based on 14CO2 production from [1-14C]butyrate by isolated rumen epithelial
cells) were maximal at 4 days of age (nearly 7-fold greater than those observed at birth),
clearly indicating the ability of undeveloped rumen epithelium to absorb and metabolize VFA.
Butyrate oxidation rates decreased gradually until weaning at 56 days, and were comparable
to those observed in older lambs. In rumen epithelial cells isolated from 28-day-old and
younger lambs, addition of unlabeled glucose or butyrate decreased 14CO2 production from the
other labeled substrate. The data presented could not distinguish between actual inhibition of
oxidation of the alternative labeled substrate by addition of unlabeled substrate, or simple dilu-
tion of the specific activity within the acetyl-CoA pool from the labeled substrate by the
unlabeled substrate prior to complete oxidation in the citric acid cycle (Baldwin and Jesse,
1992). Either explanation, however, is consistent with the ability of neonatal rumen epithelium
to absorb and oxidize butyrate. What is not clear is why neonatal rumen epithelium should pos-
sess that ability in such a magnitude at a time when little if any butyrate is present within the
rumen. The changes in rumen epithelial metabolism are summarized in fig. 2.

3.2. Ketogenesis
Baldwin and Jesse (1992) also found that ketogenesis from butyrate, as measured by BHBA
production rate, was undetectable at birth, but increased to a low, relatively steady rate

Fig. 2. Summary of the metabolic changes occurring during the development of lamb rumen epithelium.
Based on Baldwin and Jesse (1992).
Energy metabolism in the developing rumen epithelium 399

from 4 through 42 days of age. After 42 days of age ketogenic rates increased markedly (about
8-fold), so that at weaning at 56 days of age butyrate conversion to ketone bodies was occur-
ring at nearly adult rates (Giesecke et al., 1979; Bush, 1988; Harmon et al., 1991; Baldwin
and Jesse, 1992). Based on these findings, lamb rumen epithelial development was suggested
to occur in stages, with the two most prominent stages being the period of rapid rumen growth
and keratinization occurring between about 28 and 42 days of age, followed by the onset of
metabolic maturity as indicated by the onset of high rates of ketogenesis from butyrate
(between 42 and 56 days of age) at weaning (Baldwin and Jesse, 1992). The suggestion was
made that VFA from the rumen fermentation may be acting to promote rumen epithelial meta-
bolic development in the same way that VFA had been noted to promote rumen papillary
development (Warner et al., 1956). However, two confounding variables existed in this study,
namely the change in diet (leading to physiological adaptation) and the increase in age of the
lambs that would be associated with ontogeny of rumen epithelial development (Shirazi-
Beechey et al., 1991a; Baldwin and Jesse, 1992).
To distinguish between these two possibilities, a study was conducted to determine the
ability of VFA to stimulate rumen metabolic development (Lane and Jesse, 1997). Milk-fed
lambs received either continuous intraruminal infusions of a physiological mixture of VFA
(acetate, propionate, butyrate) or saline, or no intraruminal infusions, for 7–10 weeks. No sig-
nificant differences in rumen epithelial parameters were found in this study, but several trends
were noted. Papillae length tended to be longer in the VFA-infused lambs, suggesting that the
VFA were acting to stimulate papillae growth as expected (Warner et al., 1956). Glucose
oxidation tended to be lower, and AcAc production from butyrate higher, in the VFA-infused
lambs than in the saline-infused or uninfused controls. No other metabolic differences were
observed among the three infusion treatments. Both glucose oxidation and BHBA production
from butyrate were similar between the various infusion treatments and conventionally reared
lambs (Baldwin and Jesse, 1992; Lane and Jesse, 1997). The results of VFA infusion on stim-
ulating rumen epithelial development in this study were inconclusive, in view of the lack of
significant treatment differences. The minimal effect of VFA infusion on papillae develop-
ment suggests that insufficient amounts of VFA may have been administered during this
experiment to stimulate maximal rumen metabolic development. The similarity in glucose
oxidation and BHBA production from butyrate between the various infusion treatments and
conventionally reared lambs, however, provided additional support to the concept that onto-
genic factors play a prominent role in rumen epithelial metabolic development.

3.3. Ontogeny of rumen metabolic development

A subsequent study then addressed the role of ontogenic development of rumen metabolism
by separating out the effects of age and diet on rumen epithelial metabolic development (Lane
et al., 2000). Lambs were either maintained on a milk diet before slaughter (0, 4, 7, 14, 28,
42, 49, 56, or 84 days of age), or at 49 days of age were weaned onto solid feed and slaugh-
tered at 84 days of age. Glucose oxidation by rumen epithelial cells isolated from the milk-fed
lambs followed the same general pattern as observed with conventionally reared lambs
(Baldwin and Jesse, 1992; Lane et al., 2000). Glucose oxidation rates by isolated rumen
epithelial cells were not different among the 84-day-old milk-fed lambs, 84-day-old lambs
weaned at 49 days of age (Lane et al., 2000), or conventionally reared lambs weaned at 56 days
of age (Baldwin and Jesse, 1992). These results are similar to those of Giesecke et al. (1979),
who observed no difference in glucose oxidation by rumen epithelial pieces from 8–12-week-old
milk-fed or conventionally reared lambs.
400 B. W. Jesse

In contrast to that observed with normally reared lambs (Baldwin and Jesse, 1992),
butyrate oxidation by isolated rumen epithelial cells was undetectable in lambs 7 days old or
less (Lane et al., 2000). No explanation was given for this difference, especially since
conventionally reared lambs at this age would not yet have begun to consume solid feed and
would not differ physiologically from milk-fed lambs in that study (Lane et al., 2000).
Different breeds of sheep were used in the two studies, however, which may have had some
effect on the results. Subsequent to that age (7 days), little difference in butyrate oxidation
by isolated rumen epithelial cells was observed through 94 days. Similar to that for glucose
oxidation, butyrate oxidation was not different between rumen epithelial cells isolated from
84-day-old milk-fed lambs and 84-day-old lambs weaned at 49 days of age (Lane et al.,
2000). Ruminal butyrate oxidation by the 84-day-old milk-fed lambs and 84-day-old lambs
weaned at 49 days of age was ~75% of the rate observed with conventionally reared lambs
weaned at 56 days of age (Baldwin and Jesse, 1992; Lane et al., 2000). The study of Lane et al.
(2000) confirms the findings of Giesecke et al. (1979), and demonstrates that changes in glucose
and butyrate oxidation in developing rumen epithelium can occur independently of diet.
Prior to 42 days of age, ketogenesis from butyrate by isolated rumen epithelial cells, as
measured by BHBA production, was relatively low in rumen epithelial cells from the milk-fed
lambs, but increased thereafter to rates comparable to conventionally reared lambs (Baldwin
and Jesse, 1992; Lane et al., 2000). Similar ketogenic rates were observed in rumen epithelial
cells from the 84-day-old lambs weaned at 49 days of age. These results are consistent with
those obtained by Giesecke et al. (1979), who observed increasing ketogenic rates from
butyrate by rumen epithelial pieces from 1-week, 3–4-week, and 9–10-week-old lambs. In
contrast to Lane et al. (2000), ketogenesis by the rumen epithelial pieces from 9–10-week-old
milk-fed lambs was about 75% of the rate found in lambs of the same age that had been reared
and weaned conventionally. Bush (1988) also noted increased rates of ketogenesis by rumen
epithelial tissue with age from milk-fed calves. However, in that study ketogenesis by rumen
epithelial tissue from conventionally reared and weaned calves was nearly 8-fold greater than
in milk-fed calves of the same age. Thus, all three of these studies indicate that ketogenic
capacity of rumen epithelium increases with age regardless of the diet consumed, although the
magnitude of the reported increase did differ, perhaps due to species differences (Giesecke
et al., 1979; Bush, 1988; Lane et al., 2000). This again suggests that rumen epithelial metabolic
development can occur in the absence of the rumen fermentation and VFA production, although
dietary responses may modulate that ontogenic process.

3.4. Differential gene expression in rumen metabolic development

The mechanism responsible for the observed increase in ketogenesis by the rumen epithelium
from milk-fed lambs may be an increase in expression of the genes encoding ketogenic
enzymes (Lane et al., 2002). Northern blots of total rumen epithelial RNA isolated from
conventionally reared and milk-fed lambs of different ages were probed with cDNA probes
against AcAc-CoA thiolase and HMG-CoA synthase, the two enzymes that are generally
regarded as regulating ruminal ketogenesis (Leighton et al., 1983). In conventionally reared
lambs the relative abundances of AcAc-CoA thiolase and HMG-CoA synthase mRNA in
rumen epithelium increased gradually (Lane et al., 2002), and generally paralleled the
reported changes in ketogenesis in rumen epithelium from conventionally reared lambs
(Baldwin and Jesse, 1992). Relative abundance of HMG-CoA synthase mRNA followed a
similar pattern to that of AcAc-CoA thiolase mRNA, but exhibited a sharper increase between
42 and 49 days of age. Because this was the time when ketogenesis markedly increased in
Energy metabolism in the developing rumen epithelium 401

rumen epithelium from conventionally reared lambs (Baldwin and Jesse, 1992), these data
suggest that expression of the HMG-CoA synthase gene may be the factor controlling rate of
ketogenesis in the developing rumen epithelium.
In milk-fed lambs, changes with age of relative AcAc-CoA thiolase mRNA abundance
were similar to those observed in conventionally reared lambs. However, relative abundance
of HMG-CoA synthase mRNA in rumen epithelium from milk-fed lambs differed from that
observed in conventionally reared lambs. The relative abundance of HMG-CoA synthase
mRNA in rumen epithelium from milk-fed lambs remained relatively low through 42 days of
age, then exhibited an almost quantum jump between 42 and 49 days of age, remaining rela-
tively high thereafter (Lane et al., 2002). Again, this time corresponds generally with the
onset of marked ketogenesis in rumen epithelium from the milk-fed lambs. These data
support the concept that rumen epithelial metabolic development can occur in the absence of
rumen fermentation (i.e. the dramatic jump in relative abundance of HMG-CoA synthase
mRNA after 42 days of age in milk-fed lambs). Nevertheless, solid feed intake and the con-
comitant production of rumen fermentation products (i.e. VFA) can modulate that process, as
shown by the difference in change in relative abundance of HMG-CoA synthase mRNA in
rumen epithelium between the conventionally reared and the milk-fed lambs. These data are
also consistent with the findings of Giesecke et al. (1979) and Bush (1988).
These results in rumen epithelium are consistent with the findings of other researchers in
the small intestine, where both ontogenic development and dietary induction of various
enzymes have been reported. For example, lactase activity decreases and dipeptidylpeptidase
IV activity increases in the lamb small intestine regardless of dietary treatment (conventional
rearing or maintenance on a milk diet), indicating ontogenic control of these enzymes
(Shirazi-Beechey et al., 1991b). On the other hand, activity of the sodium-dependent glucose
cotransporter in lamb intestine does change in response to dietary treatment (Shirazi-Beechey
et al., 1991a). Harmon et al. (1991) have reported increased acyl-CoA synthase activites for
acetate, propionate, and butyrate in adult bovine rumen epithelium in response to increased
dietary energy intake. The response of ketogenic gene expression to solid feed intake in
conventionally reared lambs may be the result of a mechanism similar to that resulting in
increased acyl-CoA synthetase activity in adult rumen epithelium, acting in conjunction with
ontogenic factors. The unique aspect of epithelial metabolic development in the neonatal rumen,
especially ketogenesis, is the association of both ontogenic and physiological factors that appar-
ently affect metabolic development by altering expression of the genes encoding ketogenic, and
perhaps other metabolic, enzymes within the rumen epithelium (Lane et al., 2002).
A recent report found similar ontogenic and physiological effects on sodium and chloride
transport in developing calf rumen epithelium (Breves et al., 2002). Sodium transport by
rumen epithelium increased with age of the calves, independently of dietary treatment (milk-
fed or weaned onto solid feed). Chloride transport by rumen epithelium also increased with
age of the calf, but exhibited a greater increase in those calves weaned onto solid feed than in
those maintained on a milk diet. Increased sodium and chloride transport by the developing
rumen epithelium could reflect an increase in the VFA absorptive capacity by the rumen
epithelium (Sehested et al., 1999b). These data provide further evidence of the importance of
ontogenic events in the metabolic development of the rumen epithelium.

4. FUTURE PERSPECTIVES
From the above discussion it should be clear that there are many unanswered questions
concerning metabolic development in the neonatal rumen epithelium. Certainly the area of
402 B. W. Jesse

substrate uptake in developing rumen epithelium needs to be addressed. To date, no informa-


tion is available on specific transporters for glucose or VFA in developing rumen epithelium,
and how the activity of those transporters changes during development. The recent discovery
of the SGLT1 glucose transport protein in mature rumen epithelium leads to the question of
when that transporter appears during rumen epithelial development (Zhao et al., 1998;
Aschenbach et al., 2000a,b, 2002), and the role the transporter plays in glucose metabolism
by the developing rumen epithelium. Similarly, given the importance of sodium and chloride
in VFA transport by rumen epithelium (Sehested et al., 1999b), the recent discovery that
sodium and chloride transport increase during rumen epithelial development (Breves et al.,
2002) is suggestive that VFA transport capacity may also increase during development. These
issues should be addressed in the future. Related to VFA uptake is the issue of VFA activa-
tion, which has been suggested to be the rate-limiting factor in VFA metabolism (Ash and
Baird, 1973). No information is available to indicate when the acyl-CoA synthetases appear
during rumen epithelial development. Given the ability of neonatal rumen epithelial cells to
utilize butyrate (Baldwin and Jesse, 1992), the acyl-CoA synthetases are likely to be present
soon after, if not at, birth, but that needs to be determined.
The role of ontogeny in rumen epithelial metabolic development is only now being recog-
nized for the importance it plays in this process. The question then arises as to how that
process is controlled. Further research into this area will certainly require the isolation of
genomic clones encoding proteins that respond to dietary changes, e.g. structural proteins
such as the small proline-rich proteins (Wang et al., 1996), as well as those that exhibit onto-
genic patterns of development, such as HMG-CoA synthase (Lane et al., 2002). A comparison
of the regulatory regions of these genes should provide information about the transcription
factors potentially involved in regulating the expression of these genes. That information in
turn could lead to identification of the signal transduction pathways that ultimately lead to the
activation of these genes. Various reports have noted the importance of agents such as
butyrate, insulin, and epidermal growth factor in stimulating the proliferation of rumen
epithelial cells (Sakata et al., 1980; Galfi et al., 1991; Baldwin, 1999; Galfi and Neogrady,
2001). Since the signal transduction pathways of some of these agents have been identified,
this information should be helpful in establishing the mechanisms regulating gene expression
within the developing ruminal epithelium, and the interplay between physiological/dietary
factors and ontogenic factors that result in complete rumen epithelial metabolic development.
Ultimately a more complete characterization of the processes involved in rumen epithelial
metabolic development should lead to more effective management techniques in rearing
young ruminants.

REFERENCES
Aafjes, J.H., 1967. Carbonic anhydrase in the wall of the forestomachs of cows. Brit. Vet. J. 123, 252–256.
Alio, A., Theurer, C.B., Lozano, O., Huber, J.T., Swingle, R.S., Delgado-Elorduy, A., Cuneo, P.,
Deyoung, D., Webb, K.E. Jr., 2000. Splanchnic nitrogen metabolism by growing beef steers fed diets
containing sorghum grain flaked at different densities. J. Anim. Sci. 78, 1355–1363.
Aschenbach, J.R., Bhatia, S.K., Pfannkuche, H., Gabel, G., 2000a. Glucose is absorbed in a sodium-
dependent manner from forestomach contents of sheep. J. Nutr. 130, 2797–2801.
Aschenbach, J.R., Wehning, H., Kurze, M., Schaberg, E., Nieper, H., Burckhardt, G., Gabel, G., 2000b.
Functional and molecular biological evidence of SGLT-1 in the ruminal epithelium of sheep. Amer.
J. Physiol. Gastrointest. Liver Physiol. 279, G20–G27.
Aschenbach, J.R., Borau, T., Gabel, G., 2002. Glucose uptake via SGLT-1 is stimulated by beta2-adrenocep-
tors in the ruminal epithelium of sheep. J. Nutr. 132, 1254–1257.
Ash, R., Baird, G.D., 1973. Activation of volatile fatty acids in bovine liver and rumen epithelium:
evidence for control by autoregulation. Biochem. J. 136, 311–319.
Energy metabolism in the developing rumen epithelium 403

Baldwin, R.L. VI, 1999. The proliferative actions of insulin, insulin-like growth factor-I, epidermal
growth factor, butyrate and propionate on ruminal epithelial cells in vitro. Small Ruminant Res. 32,
261–268.
Baldwin, R.L. VI, Jesse, B.W., 1992. Developmental changes in glucose and butyrate metabolism by
isolated sheep ruminal cells. J. Nutr. 122, 1149–1153.
Baldwin, R.L. VI, McLeod, K.R., 2000. Effects of diet forage:concentrate ratio and metabolizable energy
intake on isolated rumen epithelial cell metabolism in vitro. J. Anim. Sci. 78, 771–783.
Bergman, E.N., 1990. Energy contributions of volatile fatty acids from the gastrointestinal tract in
various species. Physiol. Rev. 70, 567–590.
Breves, G., Zitnan, R., Schroder, B., Winckler, C., Hagemeister, H., Failing, K., Voigt, J., 2002. Postnatal
development of electrolyte transport in calf rumen as affected by weaning time. Arch. Tierernähr. 56,
371–377.
Britton, R., Krehbiel, C., 1993. Nutrient metabolism by gut tissues. J. Dairy Sci. 76, 2125–2131.
Bush, R.S., 1982. Extraction of enzymes and assessment of metabolism in bovine rumen epithelium.
Can. J. Anim. Sci. 62, 429–438.
Bush, R.S., 1988. Effect of age and diet on in vitro metabolism in rumen epithelium from holstein calves.
Can. J. Anim. Sci. 68, 1245–1251.
Bush, R.S., Milligan, L.P., 1971. Enzymes of ketogenesis in bovine rumen epithelium. Can. J. Anim. Sci.
51, 129–133.
Chen, H., Wong, E.A., Webb, K.E. Jr., 1999. Tissue distribution of a peptide transporter mRNA in sheep,
dairy cows, pigs, and chickens. J. Anim. Sci. 77, 1277–1283.
Gabel, G., Muller, F., Pfannkuche, H., Aschenbach, J.R., 2001. Influence of isoform and DNP on butyrate
transport across the sheep ruminal epithelium. J. Comp. Physiol. B 171, 215–221.
Galfi, P., Neogrady, S., 2001. The pH-dependent inhibitory action of N-butyrate on gastrointestinal
epithelial cell division. Food Res. Int. 34, 581–586.
Galfi, P., Neogrady, S., Sakata, T., 1991. Effects of volatile fatty acids on the epithelial cell proliferation
of the digestive tract and its hormonal mediation. In: Tsuda, T., Sasaki, Y., Kawashima, R. (Eds.),
Physiological Aspects of Digestion And Metabolism In Ruminants: Proceedings of the Seventh
International Symposium on Ruminant Physiology. Academic Press, New York, pp. 49–59.
Ganter, M., Bickhardt, K., Winicker, M., Schwert, B., 1993. Experimental studies of the pathogenesis of
rumen acidosis in sheep. Zbl. Veterinärmedizin A 40, 731–740.
Giesecke, D., Beck, U., Wiesmayr, S., Stangassinger, M., 1979. The effect of rumen epithelial develop-
ment on metabolic activities and ketogenesis by the tissue in vitro. Comp. Biochem. Physiol. B 62,
459–463.
Hamada, T., Maeda, S., Kameoka, K., 1976. Factors influencing growth of rumen, liver, and other organs
in kids weaned from milk replacers to solid foods. J. Dairy Sci. 59, 1110–1118.
Hanigan, M.D., Crompton, L.A., Metcalf, J.A., France, J., 2001. Modelling mammary metabolism in the
dairy cow to predict milk constituent yield, with emphasis on amino acid metabolism and milk pro-
tein production: model construction. J. Theor. Biol. 213, 223–239.
Harmon, D.L., 1986. Influence of dietary energy intake and substrate addition on the in vitro metabolism
of glucose and glutamine in rumen epithelial tissue. Comp. Biochem. Physiol. B 85, 643–647.
Harmon, D.L., Gross, K.L., Krehbiel, C.R., Kreikemeier, K.K., Bauer, M.L., Britton, R.A., 1991.
Influence of dietary forage and energy intake on metabolism and acyl-CoA synthetase activity in
bovine ruminal epithelial tissue. J. Anim. Sci. 69, 4117–4127.
Jesse, B.W., Solomon, R.K., Baldwin, R.L. VI, 1992. Palmitate metabolism by isolated sheep rumen
epithelial cells. J. Anim. Sci. 70, 2235–2242.
Juhasz, B., Szegedi, B., Keresztes, M., 1976. The abomasal digestion in calves during development of the
forestomachs. Acta Vet. Acad. Sci. Hung. 26, 281–295.
Klein, R.D., Kincaid, R.L., Hodgson, A.S., Harrison, J.H., Hillers, J.K., Cronrath, J.D., 1987. Dietary
fiber and early weaning on growth and rumen development of calves. J. Dairy Sci. 70, 2095–2104.
Lane, M.A., Jesse, B.W., 1997. Effect of volatile fatty acid infusion on development of the rumen epithelium
in neonatal sheep. J. Dairy Sci. 80, 740–746.
Lane, M.A., Baldwin, R.L. VI, Jesse, B.W., 2000. Sheep rumen metabolic development in response to
age and dietary treatments. J. Anim. Sci. 78, 1990–1996.
Lane, M.A., Baldwin, R.L. VI, Jesse, B.W., 2002. Developmental changes in ketogenic enzyme gene
expression during sheep rumen development. J. Anim. Sci. 80, 1538–1544.
Leighton, B., Nicholas, A.R., Pogson, C.I., 1983. The pathway of ketogenesis in rumen epithelium of the
sheep. Biochem J. 216, 769–772.
404 B. W. Jesse

Liebich, H.G., Dirksen, G., Arbel, A., Dori, S., Mayer, E., 1987. Feed-dependent changes in the rumen
mucosa of high-producing cows from the dry period to eight weeks post partum. Zbl.
Veterinärmedizin A 34, 661–672.
Martens, H., Kudritzki, J., Wolf, K., Schweigel, M., 2001. No Evidence for active peptide transport in
forestomach epithelia of sheep. J. Anim. Physiol. Anim. Nutr. (Berlin) 85, 314–324.
McLeod, K.R., Baldwin, R.L. VI, 2000. Effects of diet forage:concentrate ratio and metabolizable energy
intake on visceral organ growth and in vitro oxidative capacity of gut tissues in sheep. J. Anim. Sci.
78, 760–770.
Nocek, J.E., Herbein, J.H., Polan, C.E., 1980. Influence of ration physical form, ruminal degradable
nitrogen and age on rumen epithelial propionate and acetate transport and some enzymatic activities.
J. Nutr. 110, 2355–2364.
Noziere, P., Remond, D., Bernard, L., Doreau, M., 2000. Effect of underfeeding on metabolism of
portal-drained viscera in ewes. Brit. J. Nutr. 84, 821–828.
Remond, D., Ortigues, I., Jouany, J.P., 1995. Energy substrates for the rumen epithelium. Proc. Nutr. Soc.
54, 95–105.
Sakata, T., Hikosaka, K., Shiomura, Y., Tamate, H., 1980. Stimulatory effect of insulin on ruminal epithe-
lium cell mitosis in adult sheep. Brit. J. Nutr. 44, 325–331.
Scaife, J.R., Tichivangana, J.Z., 1980. Short chain acyl-CoA synthetases in ovine rumen epithelium.
Biochim. Biophys. Acta 619, 445–450.
Sehested, J., Diernaes, L., Moller, P.D., Skadhauge, E., 1999a. Ruminal transport and metabolism of
short-chain fatty acids (SCFA) in vitro: effect of SCFA chain length and pH. Comp. Biochem.
Physiol. A Mol. Integr. Physiol. 123, 359–368.
Sehested, J., Diernaes, L., Moller, P.D., Skadhauge, E., 1999b. Transport of butyrate across the isolated
bovine rumen epithelium: interaction with sodium, chloride and bicarbonate. Comp. Biochem.
Physiol. A Mol. Integr. Physiol. 123, 399–408.
Shirazi-Beechey, S.P., Hirayama, B.A., Wang, Y., Scott, D., Smith, M.W., Wright, E.M., 1991a.
Ontogenic development of lamb intestinal sodium-glucose co-transporter is regulated by diet.
J. Physiol. 437, 699–708.
Shirazi-Beechey, S.P., Smith, M.W., Wang, Y., James, P.S., 1991b. Postnatal development of lamb intes-
tinal digestive enzymes is not regulated by diet. J. Physiol. 437, 691–698.
Sutton, J.D., McGilliard, A.D., Richard, M., Jacobson, N.L., 1963. Functional development of rumen
mucosa. II. Metabolic activity. J. Dairy Sci. 46, 530–537.
Tamate, H., McGilliard, A.D., Jacobson, N.L., Getty, R., 1962. Effect of various dietaries on the anatom-
ical development of the stomach in the calf. J. Dairy Sci. 45, 408–420.
Wang, L., Baldwin, R.L. VI, Jesse, B.W., 1996. Identification of two cDNA clones encoding small
proline-rich proteins expressed in sheep ruminal epithelium. Biochem. J. 317, 225–233.
Warner, R.G., Flatt, W.P., Loosli, J.K., 1956. Dietary factors influencing the development of the ruminant
stomach. J. Agr. Food Chem. 4, 788–792.
Webb, K.E. Jr., Matthews, J.C., Dirienzo, D.B., 1992. Peptide absorption: a review of current concepts
and future perspectives. J. Anim. Sci. 70, 3248–3257.
Weigand, E., Young, J.W., McGilliard, A.D., 1975. Volatile fatty acid metabolism by rumen mucosa from
cattle fed hay or grain. J. Dairy Sci. 58, 1294–1300.
White, R.G., Leng, R.A., 1980. Glucose metabolism in feeding and postabsorptive lambs and mature
sheep. Comp. Biochem. Physiol. A67, 223–229.
Young, J.W., Thorp, S.L., Delumen, H.Z., 1969. Activity of selected gluconeogenic and lipogenic
enzymes in bovine rumen mucosa, liver and adipose tissue. Biochem. J. 114, 83–88.
Zhao, F.Q., Okine, E.K., Cheeseman, C.I., Shirazi-Beechey, S.P., Kennelly, J.J., 1998. Glucose transporter
gene expression in lactating bovine gastrointestinal tract. J. Anim. Sci. 76, 2921–2929.
17 Splanchnic carbohydrate and energy
metabolism in growing ruminants1

N. B. Kristensena, G. B. Huntingtonb, and D. L. Harmonc

aDepartment of Animal Nutrition and Physiology, Danish Institute of


Agricultural Sciences, DK-8830 Tjele, Denmark
bDepartment of Animal Science, North Carolina State University, Raleigh,

NC 27695-7621, USA
cDepartment of Animal Sciences, University of Kentucky, Lexington,

KY 40546-0215, USA

Ruminal fermentation precludes a simple description of nutrient availability based on nutrient


intake. Thus, we must strive to understand the nutrient needs of the microflora and gut and then
evaluate nutrient availability after these needs have been met. Glucose is extensively metabolized
by gut tissues such that the net supply to the liver is often zero or negative. Despite this extensive
metabolism, small intestinal digestion can significantly increase glucose availability and metab-
olism. Lactate is derived from the diet, from ruminal bacterial metabolism and from endogenous
metabolism. Because of its ubiquitous nature, lactate production from the gastrointestinal tract
and viscera varies widely. However, lactate is a major glucose precursor in ruminants, supplying
9–35% of hepatic glucose carbon. Short-chain fatty acids are the major currency of ruminant
energy metabolism, accounting for 45% of digestible energy intake. Significant quantities of
short-chain fatty acids are metabolized by ruminal epithelium; however, it appears that in the fed
ruminant this epithelial metabolism is limited to butyrate and longer short-chain fatty acids.
Estimates indicate that 5% of ruminally supplied propionate is metabolized by the rumen epithe-
lium and 30% of arterially supplied acetate is metabolized by the portal-drained viscera. These
findings allow estimates of ruminal short-chain fatty acid production to be obtained from portal
appearance of short-chain fatty acids corrected for portal-drained visceral metabolism of arterial
short-chain fatty acids and ruminal epithelial metabolism of butyrate.

1. INTRODUCTION
Compared with other mammals, ruminants could seem less efficient in capturing energy in
the form of body tissue, fetus, or milk. For example, a young pig on a nutritionally adequate

1Approved as publication No. 02-07-97 by the Kentucky Agricultural Experiment Station.

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
405 © 2005 Elsevier Limited. All rights reserved.
406 N. B. Kristensen et al.

diet captures 76% of ME (metabolizable energy) as tissue gain (Dunkin and Campell, 1982)
whereas a growing steer on a forage diet captures 46% or less of ME as tissue gain (Varga
et al., 1990). Does this indicate that ruminants are energetically inefficient? The answer to this
question is not as straightforward as might be indicated by its simplicity. First of all, the relative
growth rates of the pig and the steer will affect the energetic efficiency, i.e. lower relative growth
rate means that a relatively higher proportion of the total energy is used for maintenance.
A quite different aspect is that interchanging the diets would have disastrous consequences
for the performance of the young pig whereas the steer might do well though likely to suffer
from overfeeding. The pig will not be able to obtain sufficient amounts of nutrients from a
forage diet to obtain its potential growth rate.
These findings and the fact that ruminants utilize intravenously infused glucose as effi-
ciently as nonruminants (Reid et al., 1980) indicate that the key to understanding both
possibilities and limitations in ruminant nutrition and efficiency is related to the digestive
strategy of ruminants. The forestomach fermentation in ruminants implicates that there is
only an indirect relationship between the molecular composition of the feed and the actual
nutrients available for absorption. The fermentation has a major influence on digestion and
metabolism of all organic dietary components, i.e. carbohydrate, protein, fat, and vitamins.
Carbohydrates make up the largest fraction of almost any diet for functional ruminants and
the utilization of carbohydrate will therefore be of importance to both efficiency and per-
formance. However, a number of controversies still exist connected to the availability of
carbohydrate (starch) for postruminal digestion and absorption as well as quantitative rela-
tionships between carbohydrate fermentation and end-product (short-chain fatty acid; SCFA)
availability to the animal. The purpose of this chapter is to detail some of the unique aspects
of ruminant energy metabolism. Primarily, we aim to focus on the supply of glucose, lactate,
and SCFA as sources of energy and their availability to body tissues. Only through a thorough
understanding of these interrelationships can we hope to predict and explain growth responses
based on dietary inputs.

2. GLUCOSE
Because of pregastric fermentation much of the dietary carbohydrate is fermented to SCFA. This
fermentation leaves little dietary carbohydrate available for absorption in the small intestine.
Only when high-concentrate diets are fed are significant quantities presented to the small
intestine for absorption (Huntington, 1997). Thus, pregastric fermentation necessitates a con-
tinual need for very high rates of gluconeogenesis (Bergman, 1973) to meet the glucose needs
of the ruminant. The fermentation of dietary carbohydrates necessitates unique adaptations in
ruminant glucose metabolism and many of these adaptations have been detailed in some
excellent reviews (Leng, 1970; Bergman, 1973; Young, 1976); gluconeogenesis is also dis-
cussed in a separate chapter within this book (see Chapter 15 by Donkin and Hammon). We
shall focus on how dietary influences affect the glucose economy of growing ruminants and
on current information on the interorgan metabolism of glucose in ruminants.

3. SOURCES OF GLUCOSE
Blood glucose concentrations are typically 4–6 mM in most mammals; however; ruminant
concentrations tend to be lower, at 2–5 mM (Bergman, 1973). Despite low blood glucose
concentrations and continual gluconeogenesis, ruminant blood glucose concentrations
are very responsive to intestinal carbohydrate digestion and absorption (Larson et al., 1956).
Splanchnic carbohydrate and energy metabolism 407

The major carbohydrate that ruminants consume in early life is lactose from milk. However,
by 1–3 weeks of age ruminal fermentation is active and only through suckling will animals
achieve closure of the esophageal groove and bypass significant quantities of materials to the
abomasum for gastric digestion (Orskov et al., 1970). With the increase in ruminal fermenta-
tion there is a decline in the ability to digest lactose in the small intestine. Intestinal lactase
(St. Jean et al., 1989) and glucose transport (Shirazi-Beechey et al., 1989) activities in the small
intestine decline after weaning.
Postweaning dietary carbohydrates contributing directly to glucose supplies include the
various forms of α-linked glucose available in plants. Russell and Gahr (2000) described the
classification of food carbohydrates as occurring in four forms: (1) free (not associated with
the cellular structure), such as lactose in milk or fructose in honey; (2) intracellular, which
includes soluble sugars and storage polysaccharides such as starch and fructans; (3) cell wall
components including cellulose, hemicellulose, pectins, and gums; and (4) chitin, a component
of the exoskeleton. For the functioning ruminant, only the intracellular storage polysaccharide,
starch, contributes significantly to absorbed glucose. The remaining forms of food carbohydrate
are first fermented to SCFA.
Huntington (1997) summarized numerous digestion experiments with starch intakes ranging
from 1.5 to 10.6 kg/d. In these experiments, ruminal starch digestibility ranged from 94% to
50%. The net result is that starch flow to the small intestine ranged from 90 to over 5000 g/d.
These data demonstrate that starch intake can make a sizable contribution to the glucose needs
of growing ruminants. However, to determine the contribution of starch intake to glucose avail-
ability, the efficiency of small intestinal digestion must be known.

4. IMPACT OF INTESTINAL DIGESTION ON GLUCOSE SUPPLY


Several experiments have used animals fitted with hepatic portal vein and hepatic vein
catheters to measure the quantity of glucose exiting the portal-drained viscera (PDV) and
entering the liver (Huntington et al., 1989). This measurement provides a means of deter-
mining the net glucose contributions to the liver or peripheral tissues and measures the sum
of glucose absorption and metabolism. Across a wide range of experiments encompassing
varied diets, intakes, and physiological states, net glucose absorption is almost always zero or
negative (Reynolds et al., 1994). This is not to say that glucose is not being absorbed, but
rather that very large amounts of glucose from the arterial supply are being metabolized such
that the “net” result from absorption and metabolism is zero or negative. In a study designed
to quantitate intestinal contributions to portal glucose supply, Huntington and Reynolds
(1986) abomasally infused glucose and corn starch into heifers. Overall, they recovered an
average of 65% of the glucose and 35% of the starch as glucose in portal blood. No differ-
ences were observed for the amounts of glucose recovered from animals fed alfalfa hay or a
high-concentrate diet at two intakes, suggesting little effect of adaptation for carbohydrate
assimilation. Kreikemeier et al. (1991) fed steers alfalfa hay to minimize intestinal carbohy-
drate supply and abomasally infused them with glucose, corn starch, or corn dextrins at
20, 40, and 60 g/h. Infusions all lasted 10 h, with samples taken during the final 6 h. Glucose
infusion resulted in 90% recovery of intestinal glucose disappearance in portal blood whereas
only 19% and 32% of the dextrin and starch intestinal disappearance were recovered in portal
blood, respectively.
Factors such as microbial fermentation and gut tissue metabolism must certainly make
a large contribution to small intestinal carbohydrate disappearance and emphasize the need
for measures of tissue metabolism and intestinal disappearance to more accurately describe
408 N. B. Kristensen et al.

processes of digestion and absorption. The very high metabolic activity of the PDV tissues
has been shown to be a major factor in the apparently low net rates of glucose absorption
(Reynolds and Huntington, 1988a). These authors (Reynolds and Huntington, 1988a,b) meas-
ured directly the contribution of stomach and intestinal tissues to nutrient absorption in beef
steers. When steers were fed a concentrate diet, comparatively large amounts of glucose were
absorbed from the intestines; however, the amounts utilized by ruminal and other stomach
tissues were so great that the overall net PDV absorption was negative. Attempts were made
in previous studies to account for this negative net glucose absorption and thereby obtain a
better estimate of net glucose absorption by including control (water) infusions (Kreikemeier
et al., 1991). However, more recent work has shown that increasing the peripheral supply of
carbohydrate, either through intraduodenal or intrajugular infusion of glucose, increases the
metabolism of arterially supplied glucose by the PDV (Balcells et al., 1995), making these
corrections tenuous at best.

5. DIET EFFECTS ON GLUCOSE METABOLISM


In previous sections we have attempted to define relationships of intestinal supply and
glucose availability. However, it needs to be clearly pointed out that the major determinant of
glucose supply is dietary energy intake (Herbein et al., 1978). Experiments assessing whole-
body glucose metabolism have clearly shown that glucose irreversible loss, a measure of the
flow of glucose through the body pool never to return, and thus, at steady state, an indicator
of glucose production, is a function of digestible energy intake. Schmidt and Keith (1983)
tested this hypothesis using steers fed 70% corn vs. 70% alfalfa diets fed at equal energy
intakes. They demonstrated that when steers were fed at equal energy intakes, glucose
irreversible loss was equal. When dry matter intakes were equalized, glucose irreversible loss
was greater for the 70% corn diet because of the greater energy intake with the corn. In a
related study (Russell et al., 1986) it was demonstrated that glucose irreversible loss was
directly related to energy intake independent of body size in steers ranging in weight from
136 to 470 kg.
These relationships depend on the tight control between digestible energy intake and glu-
coneogenesis. Organic matter fermented in the rumen will supply glucose precursors,
primarily propionate, to meet the glucose needs of the host. These relations are borne out in
the work of Van Maanen et al. (1978), who determined ruminal propionate production and
glucose irreversible loss in steers fed forage and grain-based diets with the propionate-
enhancing antibiotic, monensin. Monensin increased ruminal propionate production by 49%
on the forage diet and by 76% on the grain diet. Associated with these increases in propionate
were increases in glucose irreversible loss of 7% and 16% for the forage and grain diets,
respectively. This study shows that increasing propionate supply can increase glucose irreversible
loss, but not in direct proportion. These results were similarly borne out by Seal and Parker
(1994) using intraruminal infusion of propionate in calves. Only at their highest propionate
infusion (1 mol/d) was glucose irreversible loss increased. Interestingly, ruminal propionate
infusion decreased PDV glucose use from 28% to 11% of glucose irreversible loss.
The relationships between dietary energy intake and glucose irreversible loss depend upon
two related assumptions: (1) ruminants have a very tight control of hepatic glucose production,
and (2) digestion and absorption of starch in the small intestine contributes little to glucose
irreversible loss in these studies because they are dependent on glucose derived from the
products of ruminal fermentation. Bauer et al. (1995) infused phlorizin, a potent inhibitor of
SGLT1, into the abomasum of steers and sheep and demonstrated that when glucose active
Splanchnic carbohydrate and energy metabolism 409

transport was inhibited, hepatic glucose production increased resulting in no change in total
splanchnic glucose output. This demonstrates that glucose production is well coordinated
between the PDV and the liver.
A different approach was used in the study by Harmon et al. (2001). They infused a partially
hydrolyzed starch solution either ruminally or abomasally in growing steers. Shifting the site
of starch digestion from the rumen to the small intestine increased glucose utilization by PDV
tissues (132%), PDV glucose flux (310%), and irreversible loss of glucose (59%). Abomasal
infusion resulted in greater total energy availability (28%) from the total splanchnic tissues.
Thus, shifting starch digestion to the small intestine increases PDV glucose uptake and uti-
lization without a corresponding decrease in hepatic glucose production. This shift results in
greater glucose supplies to the periphery. This would seem in contrast to the results of
Herbein et al. (1978), who related glucose irreversible loss solely to energy intake. These rela-
tionships may not hold if significant quantities of starch are digested and absorbed in the
small intestine. Balcells et al. (1995) infused sheep jugularly with glucose and found that glu-
cose irreversible loss increased over 2-fold. Accompanying this increase in systemic glucose
availability was an increased utilization of glucose by the PDV. However, in their experiment,
the fraction of whole-body glucose used by the PDV remained constant (30% of whole-body
glucose irreversible loss) despite the increase in glucose irreversible loss. These results are
in agreement with their later work (Cappelli et al., 1997) where sheep received exogenous
glucose either intrajugularly or intraduodenally. Supplying glucose by either route increased
whole-body glucose irreversible loss and portal glucose utilization, and again, portal glucose
utilization was approximately 30% of glucose irreversible loss.
These results suggest that the fraction of whole-body glucose irreversible loss used by
the PDV is relatively constant. However, both of these studies were relatively short-term,
lasting 6 to 8 h. They do not answer whether or not long-term exposure causes tissues to adapt
and use more or less of the available glucose. In the study by Harmon et al. (2001) they
infused a partially hydrolyzed starch solution either ruminally or abomasally in growing
steers for 7 days. In their study, portal glucose utilization was 23% of whole-body glucose
irreversible loss with the ruminal infusion and this increased to 34% when the carbohydrate
was infused abomasally. Thus, despite a 58% increase in glucose irreversible loss, there
was a concomitant increase in the fraction of glucose metabolized by PDV tissues. It is not
known if this increase in metabolism was the result of tissue adaptation or simply differences
in cattle and sheep. With the ruminal infusion an increase in metabolism could reflect more
energy available as SCFA resulting in less PDV glucose use, as was seen with the ruminal
propionate infusions of Seal and Parker (1994) described above. A decrease in net PDV
glucose use has also been reported for steers fed 450 g/d sodium propionate (Harmon and
Avery, 1987).
McLeod et al. (2001) used the ruminal/postruminal infusion of carbohydrate model
described above (Harmon et al., 2001) to study energy balance in growing steers. They
reported that abomasal infusion of carbohydrate increased retained energy; however, based on
calorimetric data, the energy retained was retained solely as fat. When combined, these results
suggest that an increased availability of glucose increases the energetic efficiency and PDV
metabolism of glucose, but this may also result in greater fat deposition. One could speculate
that increased circulating glucose results in increased insulin and increased fat deposition.
Others have suggested that there are specific effects of glucose on lipogenesis in ruminants.
Pearce and Piperova (1984) compared duodenal infusions of glucose and dextrins in sheep
and found that glucose infusion increased in vitro lipogenesis from acetate nearly 7-fold in
subcutaneous adipose tissue as compared with control (noninfused) sheep.
410 N. B. Kristensen et al.

6. DIETARY AND DIGESTIVE SOURCES OF LACTATE


Lactate entering portal blood of the gastrointestinal tract of a ruminant can come from the
diet, can be a product of rumen fermentation, or can be a product of tissue metabolism.
Dietary sources generally include lactate from fermented feeds, e.g., a product of lactobacilli
in silages. Lactate is produced in fermented feeds by homo- or hetero-fermenting lactobacilli
that vary in substrate (sugar) preferences and isomer of lactic acid produced. Most bacteria
can produce either D(−)- or L(+)-lactic acid by virtue of isomer-specific lactate dehydrogenase
and lactate racemase enzyme activity (Counotte and Prins, 1981; McDonald et al., 1991).
Lactic acid concentrations in most silages prepared by adequate or competent techniques
range from 3 to 12 g/kg DM. Treatments that limit fermentative activity, e.g. treatment with
mineral acids, formic acid, or formalin, or wilting before ensiling, can reduce lactic acid con-
centration by one-half or more. Treatments that induce or enhance fermentative activity in the
silo, e.g. inoculation with bacteria, addition of sugars or propionic acid, decreased particle
size by precision chopping of herbage before ensiling, in general increase lactic acid concen-
tration 1.5–2.0-fold (McDonald et al., 1991; Sheperd et al., 1995; Kung et al., 2000; Kung and
Ranjit, 2001). The isomeric proportions of lactic acid in these feedstuffs have not been studied
extensively; available reports indicate that L(+):D(−) ratios range from 0.3:1 to 1:1 (Schaadt,
1968; Hull, 1996; Kung et al., 2000). McDonald et al. (1991) suggested that as time of ensiling
increases, the L(+):D(−) ratio approaches 1:1 because of racemase activity of lactobacilli.
Lactate is both produced and used by ruminal microbes. Numbers (and activity) of lactate
producers and users respond rapidly to readily fermentable substrate (Counotte and Prins,
1981; Goad et al., 1998), which means that ruminal lactate concentrations usually are very
low (1–3 mM) to nondetectable. Calculations of lactate production in the rumen are in a
similar range, 1–3 mmol/h (Counotte and Prins, 1981). In cases of abrupt changes in intake
of readily available carbohydrates there can be a rapid increase in ruminal lactate concentra-
tions, indicating that production can exceed use or removal from the rumen. For example,
Harmon et al. (1985) dosed beef steers intraruminally with 12 g of glucose per kg of body
weight and measured peak concentrations of L(+)- lactate and D(−)-lactate of 77 and 40 mM,
respectively, 30 h after the dosing. As a result of rapid fermentation of the carbohydrates, the
proportion of L(+):D(−)-lactate may change from predominantly L(+) to predominantly D(−).
The change in isomeric ratio is more a function of increased production than differences in
use rates, because both isomers are used by ruminal microbes at similar rates. The rapid pro-
duction and accumulation causes a ruminal acidosis that is lethal to many ruminal protozoa,
and also causes a systemic acidosis in the host ruminant (Dunlop, 1972; Counotte and Prins,
1981; Goad et al., 1998). Ruminal concentrations and isomeric proportions of lactate are the
product of the effects of ruminal production, use, absorption from the rumen, and passage
with digesta to more distal portions of the gastrointestinal tract.

7. ABSORPTION OF LACTATE FROM THE


GASTROINTESTINAL TRACT
L(+)-lactate (and presumably D(−)-lactate) are transported across cell membranes by a family
of monocarboxylate transporters (Price et al., 1998). These transporters also transport
ketones, pyruvate, and acetate. Because lactate can be a product of tissue metabolism, a sub-
strate for tissue metabolism, and the subject of transport across the plasma membrane of
epithelial cells, it is difficult to discern the relative importance of, or interactions among, these
processes on the rate of lactate appearance in portal blood draining the gastrointestinal tract.
Splanchnic carbohydrate and energy metabolism 411

Further, dietary and endogenous factors that alter blood flow can negate or amplify in vivo
changes in concentration differences in blood supplying and draining the PDV. The few data
available for absorption of D(−)-lactate suggest that factors that promote production of
D(−)-lactate in the rumen also promote its absorption and appearance in hepatic portal blood
(Huntington et al., 1980, 1981; Harmon et al., 1985). For the remainder of this discussion of
lactate absorption and metabolism, “lactate” and “L(+)-lactate” will be used synonymously
unless otherwise indicated.
The studies summarized in table 1 are representative of published literature that quantifies
net flux of lactate across splanchnic tissues. The studies show that lactate absorption in sheep
and cattle ranges from approximately 2 to 200 mmol/h. Increased intake of a given diet
increases net absorption (Reynolds et al., 1991; table 1), as does increased body mass (usually
accompanied by increased intake), albeit at a nonlinear rate (Eisemann et al., 1996; table 1).
The data from Taniguchi et al. (1995; table 1) exemplify the positive relationship between
increased ruminal fermentation and lactate absorption (alfalfa vs. alfalfa and ruminal starch
infusion in table 1), and also indicate that increased intestinal appearance of glucose results
in increased portal appearance of lactate, ostensibly as a result of postruminal gut tissue
metabolism (ruminal vs. abomasal infusion of starch in table 1). The postruminal digestive
tract accounted for about one-third of lactate absorption in beef steers fed alfalfa hay or a
high-concentrate diet (Reynolds and Huntington, 1988b). The lactating dairy cows in the
studies in table 1 had similar daily dry matter intakes (data not shown), but the cows eating
the grass diet absorbed less lactate than the cows eating corn silage and supplement (Reynolds
et al., 1991; De Visser et al., 1997; table 1). McLeod et al., (1997) (table 1) found that infusion
of somatostatin decreased blood flow through PDV of sheep, but increased venoarterial dif-
ference of lactate (data not shown), resulting in increased net absorption of lactate. The study
of Bauer et al. (1995; table 1) included intragastric infusion of phlorizin, which decreased net
absorption of glucose (data not shown) but had no statistically significant effect on lactate
flux. Other examples of lack of effects of metabolic regulators include similar net absorption
of lactate in control beef steers vs. steers fed a β-adrenergic agonist (Eisemann and Huntington,
1993) or control steers vs. hyperinsulemic, euglycemic beef steers receiving intravenous infusion
of insulin and glucose (Eisemann and Huntington, 1994).
Lactate makes a small but measurable contribution to the overall energy supply for rumi-
nants. Lactate accounted for approximately 4.3% of the sum of energy absorbed as SCFA and
lactate by lactating dairy cows consuming all-forage diets (De Visser et al., 1997; table 1), 8%
by lactating dairy cows consuming a 60:40 corn silage:supplement diet (Reynolds et al., 1991;
table 1), 9% by steers consuming all-forage diets (Huntington et al., 1988), and 16% by heifers
consuming a diet containing 780 g corn grain/kg of DM (Huntington and Prior, 1983).

8. HEPATIC METABOLISM OF LACTATE


The metabolic importance of lactate for ruminants centers on its role as a glucose precursor
in the liver; net lactate removal by the liver often exceeds portal supply (table 1) and can the-
oretically account for 9–35% of net hepatic glucose release (data not shown) in studies with
bovines listed in table 1. Studies with infusions of radiolabeled glucose and lactate into lambs
and steers indicate that from 5% to 11% of glucose carbon comes from L(+)-lactate, and less
than 1% comes from D(−)-lactate (Huntington et al., 1980, 1981; Harmon et al., 1983).
Recycling of carbon through lactate and glucose would cause underestimations from isotope
infusions, and calculations from net fluxes likely overestimated the true conversion of lactate
to glucose. In the sheep studies of McLeod et al. (1997; table 1) net lactate removal could
Table 1
Selected studies of L(+)-lactate fluxa across portal-drained viscera (PDV), liver, and total splanchnic (TSP) tissues of sheep and cattle

Net flux, mmol/h

Species BW, kg Diet description PDV Liver TSP Reference

Sheep wethers 36 Alfalfa hay, duodenal starch and casein infusion 4.8 −9.8 −5 McLeod et al. (1997)
Sheep wethers 36 Alfalfa hay, duodenal starch and casein infusion, 6.6 −10.6 −4 McLeod et al. (1997)
somatostatin injection
Sheep wethers 40 Alfalfa hay, starch infusion 2.2 3.2 5.4 Bauer et al. (1995)
Beef heifers 321 Alfalfa:concentrate, low intake 45 −17 28 Reynolds et al. (1991)
Beef heifers 321 Alfalfa:concentrate, high intake 82 −28 54 Reynolds et al. (1991)
Beef steers 236 Bromegrass hay:concentrate 60:40 47 −81 −34 Eisemann et al. (1996)
Beef steers 438 Bromegrass hay:concentrate 60:40 67 −101 −34 Eisemann et al. (1996)
Beef steers 522 Bromegrass hay:concentrate 60:40 63 −85 −22 Eisemann et al. (1996)
Beef steers 253 Alfalfa hay 39 −63 −24 Taniguchi et al. (1995)
Beef steers 253 Alfalfa hay, ruminal starch infusion 50 −77 −27 Taniguchi et al. (1995)
Beef steers 253 Alfalfa hay, abomasal starch infusion 75 −68 −7 Taniguchi et al. (1995)
Lactating dairy cows 645 Corn silage:supplement 60:40 216 −249 −33 Reynolds and Huntington (1988c)
Lactating dairy cows 500 Fresh ryegrass 121 −144 −23 De Visser et al. (1997)
aPositive numbers indicate net absorption or release, negative numbers indicate uptake or removal.
Splanchnic carbohydrate and energy metabolism 413

maximally account for 41–62% of net hepatic glucose production. Lactate contribution is not
calculated for the data of Bauer et al. (1995; table 1) because in some of their treatments they
measured net hepatic output of lactate.
The range of these potential hepatic fluxes and potential contribution to hepatic gluconeo-
genesis attest to the flexibility and versatility of lactate to participate in postabsorptive
metabolism. The complete data of Reynolds et al. (1991; table 1) for beef heifers showed an
interaction between intake level and percentage of dietary concentrate; net hepatic lactate
removal and potential contribution of lactate to gluconeogenesis increased when the heifers’
intake of a high-forage diet increased. However, lactate removal and potential contribution to
gluconeogenesis decreased when the heifers’ intake of a high-concentrate diet increased.
A mesenteric vein infusion of alanine in the same heifers (Reynolds and Tyrrell, 1991) increased
net alanine removal and reduced net lactate removal by the liver, but did not affect net hepatic
glucose output. These results indicate a replacement of lactate by alanine as a glucose
precursor. The complete data of Eisemann et al. (1996; table 1) predict decreased net hepatic
removal or extraction of lactate, and increased net hepatic removal of amino acids to support
increased hepatic glucose production in beef steers as they grow from 235 to 525 kg of body
weight. The somatostatin injection that increased net portal absorption of lactate in sheep also
increased net hepatic removal of lactate and increased glucose output by the liver (McLeod
et al., 1997; table 1). Steers fed a β-adrenergic agonist had an acute surge in lactate removal by
the liver that could account for up to 63% of liver glucose output on the first day of treatment.
Hepatic removal and potential contribution to gluconeogensis subsided after 7 days of treatment
(Eisemann and Huntington, 1993).

9. PERIPHERAL METABOLISM OF LACTATE


Circulating concentrations of L(+)-lactate range from 0.2 to 1.0 mM, and concentrations of
D(−)-lactate are 0.10 to 0.50 of concentrations of L(+)-lactate (Huntington et al., 1980, 1981;
Harmon et al., 1983); these studies are cited in table 1. Whole-body lactate turnover in beef
cattle and sheep ranges from approximately 5 to 10 times net portal absorption (Huntington
et al., 1980, 1981), indicating the importance of the Cori cycle in movement of carbon through
lactate and glucose between the liver and peripheral tissues, mostly muscle. Excitement or
agitation of animals can cause a rapid rise in blood lactate levels as a result of heightened
muscle activity. The major fate of D(−)-lactate is oxidation, which accounted for essentially
all D(−)-lactate turnover in steers (Harmon et al., 1983). In vitro studies with bovine tissues
show significant potential for oxidation of D(−)-lactate, with the greatest activity in kidney
cortex followed by heart and liver, the lowest activity being detected in muscle tissue
(Harmon et al., 1984).
Net flux of L(+)-lactate across hindlimbs of cattle varies in response to physiological state
of the animal and physiological interventions by researchers. As stated previously, lactate
interacts with glucose through the Cori cycle, but lactate also is used as a substrate for lipid
synthesis. Therefore, depending on the contribution of fat to tissue makeup, the hindlimbs
may be net users or net releasers of lactate (Prior et al., 1984; Eisemann et al., 1996). The
acute response of beef steers to an orally administered β-adrenergic agonist was a dramatic
increase in lactate production by hindlimbs which was not evident after 7 days of treatment
(Eisemann and Huntington, 1993). Establishment of hyperinsulemia with euglycemia in steers
enhanced glucose uptake by hindquarters, but did not significantly change lactate flux across
those tissues (Eisemann and Huntington, 1994).
414 N. B. Kristensen et al.

10. SHORT-CHAIN FATTY ACIDS OVERVIEW


Short-chain fatty acids are simple aliphatic carboxylic acids with straight or methyl-branched
hydrocarbon chains of 2 to 5 carbons. The SCFA anions with 2 (acetate), 3 (propionate), and
4 (butyrate) carbons are the most prevalent SCFA in the rumen and colon (Bergman, 1990)
and their production is closely related to the energy metabolism of rumen microbes (Russell
and Wallace, 1988). The existence of acetate in the rumen was observed more than a hundred
years ago; however, not until the 1940s was it discovered that SCFA are absorbed from the
forestomachs and make a significant contribution to ruminant metabolism (Barcroft et al.,
1944). About 67% of ruminal SCFA are absorbed across the rumen epithelium or taken up by
the rumen microbes and about 33% are carried out of the rumen by liquid passage (Peters
et al., 1990). Short-chain fatty acids leaving the rumen with liquid outflow are absorbed mainly
in the omasum and abomasum (Masson and Phillipson, 1952; Rupp et al., 1994).
In ruminants, as in other animals, a mixture of undigested feed and organic matter of
endogenous origin enters the hindgut and is fermented into gasses, SCFA, and microbial
organic matter. Fermentation in the hindgut is of little quantitative nutritional importance to
the animal compared to the forestomach, mostly because microbial protein and other non-
SCFA products of fermentation are not readily absorbed. The SCFA production in the hindgut
can be estimated as 6–13% of the total gut production based on the propionate appearance
across mesenteric drained tissues compared to the total PDV net appearance (Reynolds and
Huntington, 1988b). Studies based on isotopic dilution in the rumen and cecum have yielded
similar relative production rates (12%) between forestomach and hindgut (Siciliano-Jones
and Murphy, 1989). Therefore, forestomach fermentation is quantitatively the most important
fermentation in ruminants, and most focus is given to forestomach physiology. However, it
must be kept in mind that total gut production of SCFA does contain a hindgut component.

11. TRANSPORT BY NONIONIC DIFFUSION


The rumen is lined with a keratinized stratified squamous epithelium. The epithelium is a
heterogeneous structure with a physical barrier formed by keratinized cells facing the lumen.
The chemical barrier of the epithelium is below the keratinized cells. The majority of metabolic
activity is located in the basal cells as indicated by their high concentration of mitochondria
(Steven and Marshall, 1970; Henrikson and Stacy, 1971).
Weak electrolytes, a group to which SCFA belong, can pass biological membranes via non-
ionic diffusion; the resulting unidirectional flux is a function of concentration (activity) and
solubility in the membrane (Rechkemmer, 1991). In accordance with this theory, it has been
shown in vivo (Thorlacius and Lodge, 1973) as well as in vitro (Sehested et al., 1999b) that
the unidirectional flux rate of butyrate across rumen epithelium increases with decreasing pH.
However, the lack of proportionality between concentration of protonized acids and acetate
and propionate fluxes as well as a relatively high permeability of these acids compared
to longer-chain fatty acids has been seen as a challenge for the absorption theory based on
nonionic diffusion. Nevertheless, a generally observed phenomenon is that SCFA have a relatively
high permeability to biological membranes relative to longer-chain fatty acids (Dietschy,
1978). This means that the membranes behave as rather polar structures toward small solutes
such as SCFA. The relative absorption rates of SCFA from experiments with washed reticulo-
rumens show that absorption rates of fatty acids longer than butyrate increase with increased
chain length (pH 7), and that methyl-branched SCFA (isobutyrate and isovalerate) have lower
absorption rates than their corresponding straight-chain fatty acids (Oshio and Tahata, 1984;
Splanchnic carbohydrate and energy metabolism 415

Kristensen et al., 2000a). Although the membranes of the rumen epithelium apparently have
a relatively high permeability to acetate, propionate, and butyrate, it still makes sense to
describe their absorption as regulated by mass action (as long as we consider unidirectional
membrane fluxes of SCFA). Recent work has suggested that anion exchangers may contribute
to apical SCFA fluxes in rumen epithelium even though the quantitative importance is unknown
(Kramer et al., 1996). So far, the data available on SCFA absorption from the forestomach
seem to indicate that absorption of SCFA by diffusion can account for the quantitatively most
important SCFA absorption.

12. CARRIER-MEDIATED TRANSPORT IN RUMEN EPITHELIUM


Rumen epithelium mounted in Ussing chambers has consistently shown a remarkable difference
in the net transport of butyrate compared with the net transport of acetate and propionate (Stevens
and Stettler, 1967; Sehested et al., 1999a). While the rumen epithelium shows a small net secre-
tion (net transport from blood to lumen side of the isolated epithelium) of acetate and propionate
when epithelia are incubated without an electrochemical gradient of SCFA, a relatively large net
absorption of butyrate carbon usually occurs. The secretion of acetate and propionate by the
epithelium at first seems to argue against the concept of nonionic diffusion. However, most esti-
mates of SCFA flux in vitro have been based on 14C-labeled acids, implying that release of any
substance carrying carbon from SCFA will be interpreted as SCFA flux. A small proportion of
acetate and propionate transported across the epithelium will be oxidized under these conditions
and the epithelium has been shown to primarily excrete the CO2 on the luminal side, explaining
at least partly the net excretion of these acids (Sehested et al., 1999a).
The rumen epithelium has long been known to be capable of metabolizing SCFA and, in
particular, to have high affinity and capacity for metabolism of butyrate (Pennington, 1952).
This in fact is the key to explaining the differences in the epithelial transport of butyrate
compared with acetate and propionate. The metabolism of butyrate into acetoacetate and
3-hydroxybutyrate and the subsequent release of these compounds across the basolateral
membrane would be in agreement with the apparent normal metabolic activity of the epithe-
lium and would also explain why [14C]-butyrate was transported differently from acetate and
propionate. It is likely that the products of butyrate metabolism are transported to the serosal
(blood side) buffer carrying the label from butyrate. Keto- and hydroxyacids such as acetoac-
etate, 3-hydroxybutyrate, and lactate are more polar than SCFA because of their hydrophilic,
secondary functional group, and consequently these acids have a lower permeability in bio-
logical membranes. In skeletal muscle a monocarboxylate transporter which co-transports
lactate and protons solves an analogous transport problem for lactate across the cell membrane
(Juel, 1997). The missing piece of the puzzle would therefore be to find monocarboxylate
transporters in the epithelium that enable polarized transport of acetoacetate and 3-hydroxy-
butyrate. Recently, this transporter was shown to be present in rumen epithelium which agrees
with this sequence of events (Müller et al., 2001). It has also been shown that blocking cellular
metabolism abolishes the active component of butyrate absorption in vitro (Gäbel et al., 2001),
confirming that it is the ketone bodies formed from butyrate that are selectively transported
to the serosal side of the epithelium and not butyrate itself.

13. RUMEN EPITHELIAL METABOLISM


One of the central observations on SCFA metabolism in ruminants has been the apparently
extensive metabolism of ruminally produced SCFA by the rumen epithelium. However, this
416 N. B. Kristensen et al.

has been among the most difficult features of SCFA metabolism to understand. Numerous
reviews are available discussing SCFA metabolism (Bergman, 1990; Britton and Krehbiel,
1993; Seal and Reynolds, 1993; Rémond et al., 1995; Kristensen et al., 1998; Seal and Parker,
2000). Recent studies have challenged the view that the rumen epithelium has a dominant role
in the metabolism of acetate and propionate absorbed from the rumen.
The classic attempt to determine the quantitative relationship between SCFA production in
the gut and SCFA absorption was the work by Bergman and Wolff (1971). The production of
SCFA in the rumen based on isotopic dilution was compared with portal appearance of SCFA
corrected for PDV uptake of arterial acetate. It was concluded that large amounts not only of
butyrate, but also of acetate and propionate, were metabolized by gut epithelia. In support of
this conclusion, rumen epithelium also seemed to metabolize a large fraction of SCFA trans-
ported in vitro (Stevens, 1970). Nevertheless, these figures have long been doubted when
considering the large amounts of SCFA apparently being absorbed in high producing rumi-
nants (Sutton, 1985). These figures also lead to the paradoxical conclusion that the rumen
epithelium of a lactating cow should have oxidative needs comparable to the entire fasting
heat metabolism of the animal (Kristensen and Danfær, 2001).
Studies on rumen epithelial metabolism of absorbed SCFA may have overestimated the
metabolism by the epithelium because the actual estimation is the mixed effect of rumen
microbial and rumen epithelial metabolism. Studies on SCFA absorption under washed reticulo-
ruminal conditions that minimize bacterial activity have shown that the portal appearance of
acetate, propionate, and isobutyrate could account for the entire disappearance of these acids
from the rumen when the PDV uptake of arterial acetate is taken into account and 5% of
the propionate is assumed metabolized into lactate by the rumen epithelium (Kristensen
et al., 2000a). Butyrate was also extensively metabolized by the rumen epithelium under
washed reticulo-rumen conditions and no more than 23% of the butyrate disappearance from
the rumen could be accounted for by portal appearance of butyrate. It has previously been
observed that there is increasing portal recovery of butyrate with increasing disappearance
rates of butyrate from the rumen of sheep (Kristensen et al., 1996b, 2000b; Nozière et al.,
2000). This effect is in agreement with a saturable metabolic capacity of the epithelium.
To what extent there is interspecies differences in the metabolic capacity of butyrate in the
rumen epithelium is not yet clear, but in a study with steers, the portal recovery of butyrate
did not increase with increasing ruminal infusion rates of butyrate (Krehbiel et al., 1992).
The recovery was relatively high at all infusion levels in the steers (25%), and was equivalent
to the highest recovery level obtained in the sheep experiments. In sheep, increasing ruminal
butyrate infusion not only leads to increasing portal recovery of butyrate, but also to increas-
ing portal recovery of ruminal valerate (Kristensen et al., 2000b). These results point to
a redefinition of the role of the rumen epithelium in SCFA metabolism and suggest that the
rumen epithelium is not metabolizing large amounts of acetate and propionate as previously
assumed.

14. IS BUTYRATE OXIDIZED TO CARBON DIOXIDE


DURING ABSORPTION?
In vitro studies have shown that rumen epithelium is able to oxidize all of the three quantita-
tively most important SCFA (Baldwin and McLeod, 2000); however, the epithelial production
of 3-hydroxybutyrate and acetoacetate imply that butyrate oxidation is far lower than its
disappearance across the epithelium. Studies comparing net portal appearance of butyrate
and butyrate infusion into the rumen have indicated that major parts of the butyrate were
Splanchnic carbohydrate and energy metabolism 417

oxidized (lost), because net portal appearance of butyrate, 3-hydroxybutyrate, and acetoac-
etate accounted only for 25–45% of ruminal butyrate infusion (Krehbiel et al., 1992;
Kristensen et al., 1996b). However, the PDV has been shown to utilize 3-hydroxybutyrate
from arterial blood equivalent to 32–42% of the whole-body flux in sheep and thereby mask
the true production rate by the gut epithelia (Kristensen et al., 2000c). Intraruminal microbial
pathways might also utilize part of the infused butyrate and thereby contribute to what could
be interpreted as epithelial oxidation. This latter effect has been indicated by relatively high
recoveries (as compared to expected recovery in the fed animal) of butyrate when infused into
animals maintained under total intragastric nutrition (Gross et al., 1990a) or temporarily
washed reticulo-rumen conditions (Kristensen et al., 2000a). In conclusion, the rumen epithe-
lium has oxidative needs and butyrate is likely the most important carbon source. The
majority of the butyrate absorbed is released as butyrate, acetoacetate, and most importantly,
3-hydroxybutyrate to the portal blood.

15. WHY DO EPITHELIA METABOLIZE BUTYRATE?


Butyrate is generally considered a special metabolite for gut epithelial function (Topping and
Clifton, 2001). One way to explain the special behavior of gut epithelia toward butyrate com-
pared with acetate and propionate is that butyrate is important as an energy source for
epithelial cells (Bugaut, 1987). However, the rumen epithelium has a range of other metabo-
lites available, e.g. acetate and propionate absorbed from the rumen as well as arterially
supplied glucose. One might speculate that butyrate’s role as an important substrate for
epithelial energy metabolism might have evolved secondary to the basic need of having
butyrate removed before it enters the blood stream. Butyrate metabolism by rumen and
hindgut epithelia could therefore be seen as a protective mechanism that has two disposal
pathways, oxidation and ketogenesis. It is obvious that butyrate is handled differently from
acetate and propionate by the epithelia (Pennington, 1952), but another question remains to
be answered: is butyrate a unique metabolite? Valerate, for example, is also efficiently metab-
olized by the rumen epithelium (Kristensen et al., 2000a,b) and it has been shown that the
epithelium have the capacity to metabolize medium-chain (Hird et al., 1966) as well as long-
chain fatty acids (Jesse et al., 1992).
Butyrate is an important substrate for gut epithelia compared with acetate and propionate,
but it is apparently not a unique nutrient. Acetate, propionate, and isobutyrate are all metabo-
lites of endogenous pathways in the organism. Acetate has the lowest membrane permeability,
is utilized from peripheral arterial blood in major extrahepatic tissues (Pethick and Lindsay,
1982), and is a universal metabolite in the body in the form of acetyl-CoA. Propionate is the
main donor of 3-carbon units for gluconeogenesis in the ruminant liver and is efficiently taken
up by the liver (Leng and Annison, 1963). The endogenous sources of propionate include
degradation of uneven chained fatty acids and some amino acids (methionine, threonine,
isoleucine, and valine). Isobutyrate (an intermediate from catabolism of valine) appears in
relatively low concentrations in the rumen, but is efficiently taken up by the liver for gluco-
neogenesis (Stangassinger and Giesecke, 1979). These SCFA are not only well tolerated in
hepatic and peripheral tissues, but are key metabolites (especially acetate and propionate) in
these tissues, and this agrees with a limited uptake of these SCFA in the gut epithelia.
Butyrate, valerate, and probably longer, medium-chain fatty acids (MCFA) are less polar
and will have a relatively high permeability in cell membranes. One way of controlling per-
meability is partial oxidation of these SCFA into acetoacetate and 3-hydroxybutyrate in the
gut epithelia. When butyrate appears in the systemic circulation or is added to cell cultures,
418 N. B. Kristensen et al.

it has been shown to have a number of adverse effects: inhibition of growth and induction of
morphological changes in cultured cells of different origins including ruminal epithelial cell
lines (Prasad and Sinha, 1976; Gálfi et al., 1991); being an insulin secretagogue (Manns and
Boda, 1967); inhibition of gastrointestinal motility by stimulation of epithelial receptors
(Crichlow, 1988) and/or via systemic effects (Le Bars et al., 1954); stimulation of rumen
epithelial development (Sander et al., 1959); or killing (2.5 mmol butyrate/kg BW in lambs)
the animal (Manns and Boda, 1967). The epithelia of the gut have apparently evolved to
perform gatekeeping functions by controlling the entry of butyrate and longer-chain fermen-
tation acids into the peripheral circulation. It is tempting to speculate that the side effects of the
gatekeeper function are that these metabolites also become quantitatively important oxidative
substrates.

16. ACYL-CoA SYNTHETASES


Activation of SCFA by an acyl-CoA synthetase (also named CoA ligase or thiokinase) is the
first step in the metabolism of any SCFA in the cells of the gut epithelium, liver, or peripheral
tissues (Groot et al., 1976). The acyl-CoA synthetases are therefore believed to be key
enzymes in different tissues’ selectivity to metabolism of different SCFA. There exist
a number of distinct acyl-CoA synthetases: acetyl-CoA, propionyl-CoA, butyryl-CoA, medium-
chain fatty acid, and long-chain fatty acid-CoA synthetases.
The acetyl-CoA synthetase (EC 6.2.1.1) has a high affinity for acetate, and some affinity
for propionate (Campagnari and Webster, 1963; Groot et al., 1976; Ricks and Cook, 1981b).
However, it is noteworthy that the activity of this enzyme has been found to be low in the
rumen epithelium and liver of ruminants (Cook et al., 1969; Ash and Baird, 1973). These
observations are in line with a limited role of the rumen epithelium and the liver in metabo-
lism of absorbed acetate.
The ruminant liver has a relatively high propionyl-CoA synthetase (EC 6.2.1.17) activity
(Ash and Baird, 1973) and there exist a number of indications that propionyl-CoA synthetase
is a distinct enzyme (Ricks and Cook, 1981a,c). Among the interesting features of this
enzyme is that it is not present in rumen epithelium. This is not the same as denying any
possible activation of propionate in rumen epithelium, which obviously can occur (Weekes,
1974), but it has been shown that the propionyl-CoA synthetase activity in the liver is almost
insensitive to the presence of butyrate whereas the activity in the rumen epithelium is almost
completely inhibited by the presence of butyrate (Ash and Baird, 1973; Harmon et al., 1991).
As is the case with acetyl-CoA synthetase in rumen epithelium, the lack of propionyl-CoA
synthetase activity is in agreement with in vivo observations showing a very limited uptake of
propionate by the rumen epithelium.
As described above, one of the most striking features of rumen epithelial metabolism is a
high affinity and capacity for metabolism of butyrate. This feature is reflected in the butyryl-
CoA synthetase activity of the epithelium (Ash and Baird, 1973). The relative importance of
the liver and the rumen epithelium in the metabolism of propionate and butyrate, respectively,
is directly reflected in the acyl-CoA synthetase activities. Moreover, as butyrate was found to
have an insignificant effect on propionate activation in the liver, propionate had no effect on
butyrate activation in the rumen epithelium, but decreased the butyrate activation in the liver
(Ash and Baird, 1973). A distinct butyryl-CoA synthetase (EC 6.2.1.2) was first purified from
bovine heart mitochondria and this enzyme showed a high affinity for valerate and caproate
(Webster et al., 1965). In ruminants, butyrate affinity is also found in xenobiotic/medium-chain
fatty acid-CoA synthetases. These acyl-CoA synthetases activate a broad spectrum of
Splanchnic carbohydrate and energy metabolism 419

straight-chain fatty acids: butyrate, longer SCFA, branched-chain fatty acids, and a number
of xenobiotic (of foreign origin) carboxylic acids; others include benzoate and phenylacetate
(Aas, 1971; Vessy et al., 1999). Indirect evidence from PDV flux studies indicates cross-
specificity for valerate activation, which agrees with both types of butyrate activating
systems. So far, no specific information seems to be available on the interaction of SCFA
activation with longer-chain fatty acids or xenobiotic compounds absorbed from the rumen
(Cremin et al., 1995); however, the fact that the isolated rumen epithelium or isolated rumen
epithelial cells are able to use a wide range of fatty acids from SCFA to palmitate indicates
the presence of some activity for activating medium- as well as long-chain fatty acids by the
epithelium (Jesse et al., 1992; Hird et al., 1966).

17. HOW IS BUTYRATE METABOLIZED BY GUT EPITHELIA?


Rumen epithelial ketogenesis is remarkable compared to hepatic ketogenesis by virtue of the
fact that rumen epithelial ketogenesis is a main pathway in the fed state, and not a pathway
turned on at fasting or when the organism is facing a high “metabolic drain”. This feature is
obviously connected to the constant fueling of rumen epithelial ketogenesis via butyrate
absorption in combination with an apparent need for removal of butyrate before entering the
blood stream.
The oxidation of butyryl-CoA to acetoacetyl-CoA in rumen epithelium (fig. 1) is, from a
chemical point of view, identical to the initial steps of long-chain fatty acid β-oxidation. (For
a review on this subject, see Eaton et al., 1996.) The first 3-hydroxybutyrate intermediate of
this pathway is the L-(S)-isomer, which is not released to the peripheral circulation. Oxidation
of L-3-hydroxybutyrate-CoA yields acetoacetyl-CoA. Acetoacetyl-CoA is a branching point
between acetyl-CoA formation and ketone release because of the acetoacetyl-CoA thiolase
(EC 2.3.1.9) catalyzed equilibrium between acetoacetyl-CoA and acetyl-CoA (fig. 2). The
equilibrium constant of the reaction (6 × 10−6; Williamson et al., 1968) is strongly favoring
acetyl-CoA and this means that the concentration of acetoacetyl-CoA probably will be rela-
tively low in the mitochondrion. The production of ketone bodies from acetate (Harmon et al.,
1991) or valerate (Weigand et al., 1975) in rumen epithelium confirms that acetoacetyl-CoA
thiolase is present in the rumen epithelium, an observation also confirmed by assays on
epithelial cell extracts (Baird et al., 1970). The main function of acetoacetyl-CoA thiolase is
probably not to mediate ketogenesis from absorbed acetate, although this mediation is possible.

Fig. 1. Initial oxidation of butyryl-CoA to acetoacetyl-CoA in rumen epithelium proceeds via a pathway
similar to the initial steps in β-oxidation. A number of isoenzymes are known for both acyl-CoA dehydroge-
nases (first dehydrogenase of the pathway) and 3-hydroxyacyl-CoA dehydrogenases (Eaton et al., 1996).
However, the isoenzymes with specificity for short-chain acyl-CoA are likely to predominate in the rumen
epithelium. The hydratase in the pathway is likely crotonase (EC 4.2.1.17), also an enzyme with the highest
specificity toward short-chain acyl groups.
420 N. B. Kristensen et al.

Fig. 2. The acetoacetyl-CoA thiolase (EC 2.3.1.9) is catalyzing the reversible thiolytic cleavage of ace-
toacetyl-CoA into two acetyl-CoAs. The equilibrium between acetoacetyl-CoA and acetyl-CoA is favoring
acetyl-CoA and as a result the acetoacetyl-CoA concentration in the mitochondrion will usually be low.

The fact that the rumen epithelium contains low amounts of acetyl-CoA synthetase, and the
low affinity of the butyryl-CoA synthetase for acetate when butyrate is present, points to the
conclusion that the main function of the acetoacetyl-CoA thiolase is feeding acetyl-CoA from
the acetoacetyl-CoA pool to the TCA cycle. Therefore, even though butyrate metabolism in
the epithelium cannot be explained from the point of the energy needs of the epithelium,
butyrate metabolism is ensured to be the main oxidative substrate under in vivo conditions.
Contrary to the consensus about the initial steps of butyrate metabolism, there has been more
discussion of the subsequent metabolism of acetoacetyl-CoA. This compound can be deacylated
directly (acetoacetyl-CoA deacylase; EC 3.1.2.11) or deacylated via the 3-hydroxy-3-methyl-
glutaryl-CoA pathway (3-HMG pathway; 3-hydroxy-3-methylglutaryl-CoA synthetase and
lyase; EC 4.1.3.5 and EC 4.1.3.4); however, other alternative pathways have been suggested
and will be discussed briefly. The presence of the 3-HMG pathway (fig. 3) in rumen epithelium
is supported by the fact that the enzymes of the pathway (3-hydroxy-3-methylglutaryl-CoA
synthetase, and lyase) have been shown to be present in the epithelium in significant amounts
(Baird et al., 1970; Leighton et al., 1983).
However, isotopomer studies have had a dominant role in the arguments about ketogenic
pathways in the epithelium. Hird and Symons (1961) investigated isolated ruminal and
omasal epithelial metabolism of [1-14C]butyrate and [3-14C]butyrate into acetoacetate. The
isotopomers of acetoacetate could be partly identified by measuring the label in position 1
(CO2 from decarboxylation of acetoacetate) and in the label in the acetone fraction after
decarboxylation (interpreted as position 3). When the epithelium was incubated with
[1-14C]butyrate, 80% of the label in acetoacetate was found in position 1 and 20% of the label
was found in position 3. When the substrate was [3-14C]butyrate, 37% of the label in aceto-
acetate was found in position 1 and only 63% in position 3. The probable explanation for the
1 to 3 shifts in labeling is the thiolase-catalyzed equilibrium between acetoacetyl-CoA and
acetyl-CoA (fig. 2). The labeling pattern also gives an indication of the relative importance of
the pathway. The fact that 20% of the label in acetoacetate was found in position 3 could lead
Splanchnic carbohydrate and energy metabolism 421

Fig. 3. The 3-hydroxy-3-methylglutaryl-CoA pathway (3-HMG pathway) ensures that acetoacetyl-CoA,


despite its low concentration, can be “trapped” and deacylated. These steps of ruminal ketogenesis are similar
to hepatic ketogenesis.

to the conclusion that 80% of acetoacetate was generated without degradation to acetyl-CoA.
However, that is a false conclusion because if we assume that the labeling does not cause frac-
tionation, then the acetoacetyl-CoA generated from acetyl-CoA (acetyl-CoA will be labeled
in position 1) will be evenly distributed among carbon 1 and carbon 3 of acetoacetate. This
means that at least 40% of the acetoacetate must have been equilibrating with acetyl-CoA to
explain 20% of the total activity in position 3.
The fact that the large majority of label from [1-14C]butyrate ends up in C1 of acetoacetate
has been used as an argument against the function of the 3-HMG pathway in the epithelium.
However, this argument might not be justified because this pathway will conserve C1 label in
position C1, especially if the thiolase activity is relatively low compared to the flux through
the 3-HMG pathway. The fact that Hird and Symons (1961) found a larger 3 to 1 shift in label-
ing of acetoacetate from [3-14C]butyrate is therefore in agreement with the 3-HMG pathway
not only working on acetoacetyl-CoA derived from acetyl-CoA, but also on acetoacetyl-CoA
from the initial-oxidation steps on butyrate. It seems puzzling that only 37% of the label in
acetoacetate generated from [3-14C]butyrate was found in position 1, especially if the majority
of the acetoacetate production is through the 3-HMG pathway. However, the relative enrichment
of the acetyl-CoA pool and the acetoacetyl-CoA pool will have a major impact on the results.
It is likely that the metabolism of [3-14C]butyrate will be accompanied by a lower specific
activity of the acetyl-CoA pool compared with the [1-14C]butyrate because the [3-14C]butyrate
will be less likely to deliver a labeled acetyl-CoA to the acetyl-CoA pool compared with
[1-14C]butyrate as substrate. The only labeling of the acetyl-CoA pool from [3-14C]butyrate
will be through the thiolase-catalyzed acetyl-CoA/acetoacetyl-CoA equilibrium. This implies
that the 3 to 1 shift observed with the [3-14C] butyrate incubation indicates a far higher impor-
tance of the 3-HMG-CoA pathway than that apparently shown by the 37% of [3-14C]butyrate
found in position 1 of acetoacetate simply because the specific activity of acetyl-CoA will be
lower under these conditions.
Though acetoacetate is the product of rumen epithelial ketogenesis, it is not the primary
circulating ketone in plasma. A large proportion of acetoacetate is reduced to D-3-hydroxy-
butyrate (fig. 4) before leaving the epithelial cells catalyzed by 3-hydroxybutyrate
dehydrogenase (EC 1.1.1.30). Data on rumen epithelial enzyme activity and isotopomer
distribution in acetoacetate suggest that the 3-HMG pathway is as quantitatively important in
this tissue as it is in liver. Earlier denials (Annison et al., 1963) are partly correct in pointing
out that butyrate is not completely degraded to acetyl-CoA before incorporation into ketone
bodies.
422 N. B. Kristensen et al.

Fig. 4. D-3-Hydroxybutyrate is the dominating “ketone” in plasma due to the 3-hydroxybutyrate dehydroge-
nase (EC 1.1.1.30) catalyzed and NADH-dependent reduction of acetoacetate.

18. ALTERNATIVE KETOGENIC PATHWAYS


Although the 3-HMG pathway is active in rumen epithelium, it might not be the only keto-
genic pathway. One obvious alternative would be acetate release by the epithelium.
Endogenous acetate production has been observed in vitro when the epithelium is incubated
without substrate (Sehested et al., 1999a). The hydrolysis of acetyl-CoA into acetate and CoA
is catalyzed by acetyl-CoA deacylase (EC 3.1.2.1; Grigat et al., 1979). We might wonder why
the ruminant produces ketone bodies at all when it seems much simpler just to use acetate as
a carrier of acetyl units. One of the reasons for the production of ketone bodies could be the
cost of reactivation in recipient tissues because they would have to pay double the price for
activation when acetate is the substrate compared with acetoacetate. Nevertheless, endoge-
nous acetate production can be observed in vitro by rumen epithelium and acetate would be
an obvious candidate for interorgan acetyl transfer. However, we have only limited and indi-
rect evidence of endogenous acetate production by rumen epithelium in vivo (Kristensen
et al., 2000a). It is unknown to what extent endogenous acetate from the PDV has a role in
interorgan acetyl exchange (i.e. acetyl carbon originally absorbed in fatty acids other than
acetate itself).
Not only acetyl-CoA, but also acetoacetyl-CoA, might be directly deacylated (acetoacetyl-
CoA deacylase; EC 3.1.2.11), and thereby lead to 3-HMG-CoA-independent acetoacetate
synthesis. The acetoacetyl-CoA deacylase has been found in rumen epithelium though only
at a low activity (Baird et al., 1970). One of reasons why direct deacylation of acetoacetyl-CoA
might be of limited importance is the low acetoacetyl-CoA concentration in the mitochondrion.
The low affinity of the acetoacetyl-CoA deacylase present in rat liver was quantitatively not
important, although it was functional under in vitro conditions with high acetoacetyl-CoA
concentrations (Williamson et al., 1968).
A number of alternative pathways have been suggested to explain various parts of ketone
body formation in rumen epithelium: succinyl-CoA:3-ketoacid CoA-transferase (Bush and
Milligan, 1971); a L-3-hydroxybutyrate pathway not involving acetoacetate formation
(Emmanuel et al., 1982); and a butyrate:acetoacetyl-CoA transferase pathway (Emmanuel
and Milligan, 1983). These pathways all suggest metabolism of butyrate to 3-hydroxybutyrate
as one unbroken C4 unit. The two latter pathways appear to be closely related to cytosolic
pathways in tissues utilizing acetoacetate in de novo synthesis of fatty acids (Robinson and
Williamson, 1980). However, it is difficult to determine the quantitative importance of non-
3-HMG-CoA pathways in the rumen epithelium from the limited data available.
Splanchnic carbohydrate and energy metabolism 423

The situation for the succinyl-CoA:3-ketoacid CoA-transferase (EC 2.8.3.5; SCOT) is


different because this enzyme is indeed anticipated to be a key enzyme in ketone body metab-
olism; however, its function is opposite to the function proposed in the rumen epithelium.
This enzyme is a key enzyme in the activation of ketone bodies in peripheral tissues and the
rare deficiency of this enzyme in human infants leads to severe ketoacidosis (Synderman et al.,
1998). Succinyl-CoA:3-ketoacid transferase has been assumed to contribute to the acetoacetyl-
CoA hydrolysis in rumen epithelium because addition of succinate was followed by increased
disappearance of acetoacetyl-CoA (Bush and Milligan, 1971). If SCOT was important for the
hydrolysis of acetoacetyl-CoA, it would suggest that the concentration of either acetoacetyl-CoA
or succinate was higher in rumen epithelium compared with other tissues. However, owing to
the true reversibility (Stern et al., 1956) of the reaction catalyzed by SCOT (acetoacetyl-CoA +
succinate ↔ acetoacetate + succinyl-CoA), it might be suggested that this activity in the rumen
epithelium was connected to the specific incubation conditions in vitro and not necessarily the
pathway of acetoacetyl-CoA metabolism in vivo.

19. METABOLITE INTERACTIONS IN RUMEN EPITHELIAL


KETOGENESIS
If the rumen epithelium works in its usual position in a ruminant, or is maintained for a short
period under in vitro conditions as epithelial slices or isolated cells, it will have an obligate
requirement for chemical energy to maintain Na+, Ca2+, and K+ ion concentration gradients
and other vital cell functions. Considering a situation with a relatively constant workload of
the epithelium, it would then be expected that tissue supplied with small amounts of butyrate
would oxidize a large fraction to CO2 simply to fulfill the basic needs of ATP and sustain
basic cell functions. This relationship has been confirmed in vitro when different butyrate
concentrations were compared. Increasing the supply of butyrate was followed by the oxida-
tion of a decreasing fraction and an increasing fraction metabolized into ketone bodies (Beck
et al., 1984).
From a whole animal perspective, glucose is antiketogenic (Hamada et al., 1982) and
initially it was surprising that glucose had the opposite effect on rumen epithelial ketogenesis,
i.e. ketogenesis was stimulated by glucose (Stangassinger et al., 1979). A number of gluco-
genic substrates have been shown to impose a similar effect on epithelial metabolism. Some
variability in the response concerning the uptake of butyrate and the proportion of butyrate
oxidized has been observed, but generally a shift toward the more reduced “ketone body”,
3-hydroxybutyrate, compared with acetoacetate has been observed with the addition of a
glucogenic substrate (Goosen, 1976; Beck et al., 1984; Giesecke et al., 1985; Baldwin and
Jesse, 1996). Although the rumen epithelium is able to take up a broad range of metabolites
including glucose, glutamine, and glutamate and oxidize them (Harmon, 1986; Baldwin and
McLeod, 2000), this does not mean that glucose is the oxidative substrate that caused the shift
in ketone body production. In fact, we would surmise from the discussion of butyrate metab-
olism (see above) that the epithelium had a source of acetyl-CoA from butyrate that would be
able to fulfill any oxidative need. The reason might be that epithelium incubated without a
glucogenic source will become depleted of TCA cycle intermediates and subsequently have
difficulty maintaining ATP, NADH, and NADPH potentials. A very elegant example of this
effect is the comparison between metabolite production from butyrate and valerate in rumen
epithelium incubated in vitro (Weigand et al., 1975). When rumen epithelium was incubated
with butyrate, 0.67 of the ketone bodies produced were acetoacetate; however, when incubated
with valerate only 3-hydroxybutyrate was produced. This production was accompanied by
424 N. B. Kristensen et al.

lactate produced from the 3-carbon fraction of valerate. Therefore, there seems to be no
reason to believe that glucose or any other glucogenic substrates play a particular role as reg-
ulators of rumen epithelial ketogenesis, but the results point to the general conclusion that
rumen epithelium has a range of nutritional requirements for proper function.

20. THE IN VIVO/IN VITRO PROPIONATE ENIGMA


In vitro, rumen epithelium metabolizes propionate into lactate (Weigand et al., 1975). In vivo,
however, it has not been possible to demonstrate any major propionate metabolism into
lactate using ruminal infusion of 14C or 13C labeled propionate (Weigand et al., 1972; Weekes
and Webster, 1975; Kristensen et al., 2001). These matters have been even further confused
by the fact that in vivo experiments on portal recovery of ruminal propionate indicate that a
large proportion of propionate was metabolized by the epithelium (Bergman and Wolff,
1971). Because of the large capacity of the liver to metabolize propionate in vivo (Berthelot
et al., 2002) it is difficult to explain why the rumen epithelium should limit the propionate
supply to the liver. The reason for the large activation of propionate under in vitro conditions
is probably the cross-specificity of the butyryl-CoA synthetase. In vivo, propionyl-CoA could
be generated by thiolysis of 3-oxo-valeryl-CoA (from valerate). This latter source might be
the explanation for the high capacity of propionyl-CoA-utilizing pathways in rumen epithe-
lium. The usual metabolism of propionate via propionate carboxylation to methylmalonic
acid followed by the TCA intermediate succinate will lead to the buildup of TCA intermedi-
ates. In the liver the main pathway to export surplus TCA intermediates is gluconeogenesis.
Other tissues use nonessential amino acids (e.g. alanine and glutamine synthesis in muscles
and other tissues) to control excess TCA intermediates. In rumen epithelium, it is apparently
the malic enzyme (EC 1.1.1.40) catalyzed decarboxylation of malate into pyruvate (coupled
to reduction of NADP) and the subsequent reduction of pyruvate to lactate that removes the
surplus of propionyl-CoA from the rumen epithelial cells (Young et al., 1969). By these
mechanisms we are able to explain the differing in vitro and in vivo observations on rumen
epithelial metabolism.

21. FITTING THE CARBON BALANCE


OF FERMENTATION IN THE GUT
Although there is no doubt that SCFA are important in ruminant metabolism, no feed evalu-
ation system has been able integrate knowledge of SCFA production, absorption, and
metabolism in ruminants under production conditions. Simulation models constructed to
describe fermentation and SCFA absorption, as well as other nutrients, need to improve in
order to predict SCFA proportions in the rumen (Baker and Dijkstra, 1999). The problem has
also been what to do with the apparently huge metabolic activity of the rumen epithelium. No
model has been able to incorporate this metabolism, and this review attempts a possible
explanation.
Simulation models of ruminal fermentation and metabolism developed to date have been
constructed and validated mainly against duodenal nutrient flows. The re-evaluation of the role
of the gut epithelia in metabolism of SCFA has enabled an alternative method of model com-
parison. If the rumen and other gut epithelia do not metabolize significant amounts of acetate
and utilize only a small percentage of the propionate flux, then the net rumen production
of these acids would be predictable from PDV fluxes. Major corrections to be considered
are, however, uptake of arterial acetate by PDV tissues and epithelial butyrate metabolism.
Splanchnic carbohydrate and energy metabolism 425

The percentage of arterial acetate uptake by PDV tissues has been shown to be stable (about
32% of arterial flux) when evaluated across different rations and with relatively little post-
prandial variation in meal-fed sheep (Bergman and Wolff, 1971; Kristensen et al., 1996a). As
discussed earlier, the portal recovery of butyrate has been shown to be a more complex func-
tion of its availability, and an increased portal recovery of butyrate with increasing ruminal
production rates might be expected.
In an attempt to compare data from studies on PDV fluxes with model predictions of gas-
trointestinal fermentation, Kristensen and Danfær (unpublished data) compared portal fluxes
of SCFA from 36 studies in which a total of 58 different diets were fed to sheep and cattle
under different physiological conditions (growing/maintenance, nonpregnant/pregnant,
dry/lactating). The model used to predict fermentation and digestion of the diets was Karoline
(Danfær, 1990) (version 8a, a dynamic simulation model of a lactating cow mainly validated
against duodenal flow data; Danfær et al., unpublished). In the model, all diets were com-
pared at a fixed dry matter intake of 20 kg/d and predicted carbon output in moles SCFA
carbon per kg dry matter intake was compared to the observed/recalculated PDV fluxes in the
studies. The experimentally observed PDV fluxes were corrected for acetate uptake in PDV
tissues (assumed 32% of the arterial flux), propionate uptake by epithelial tissues (assuming
that portal flux was equal to 95% of true absorption), and butyrate recovery [assuming portal
recovery of gut butyrate production = 0.35 × P/(P + 0.05) where P = portal net appearance
mmol × h−1 × (kg BW0.75)−1]. The portal recovery of butyrate is deliberately set to a higher
level than those typically found following butyrate infusion into the normally functioning
rumen. This recovery agrees with observations with sheep maintained on intragastric nutri-
tion or short-term washed reticulo-rumen (Gross et al., 1990a,b; Kristensen et al., 2000a).
The calculated SCFA production in the 36 experiments using the correction factors above was
11.9 ± 0.4 moles C in SCFA/kg dry matter intake. The simulated value was 12.3 ± 0.2 moles C
in SCFA/kg dry matter intake and the mean bias was 0.4 moles C/kg DMI [Σ(predicted −
observed)/number of observations; see Kohn et al. (1998)]. However, the root mean square
prediction error (RMSPE) was 2.6 moles C in SCFA/kg DMI [(Σ(predicted − observed)2/
number of observations); see Kohn et al. (1998)]. On average, the model and the corrected
experimental data are in good agreement. However, there is still a need for better models to
predict net SCFA output. The corrected, experimentally determined SCFA production was, on
average, 45 ± 2% of the simulated digestible energy. However, estimates based on intragas-
tric tracer dilution, as discussed above, seem to overestimate the SCFA production and
indirect evidence also supports these figures. In fact, the SCFA production accounting for
45% of digestible energy implies that 65% of total digested carbon is found in fermentation
gases and SCFA. However, if the true relationship between portal absorption and gut produc-
tion of SCFA is similar to the relationship described by Bergman and Wolff (1971), then the
production of SCFA would need 116 ± 5% of the digested carbon to account for SCFA and
fermentation gases. This would seem to be impossible. The good news is that values of portal
absorption of SCFA actually make sense in terms of animal energy metabolism. It must
be emphasized, however, that models of ruminal and hindgut fermentation still have a lot to
gain in terms of precision of SCFA production, especially in the prediction of ruminal SCFA
composition.

22. CONCLUSIONS
Glucose is a major metabolic fuel for ruminant tissues, similar to most mammals. The pre-
gastric fermentation dictates that gluconeogenesis serves to supply the glucose needs under
426 N. B. Kristensen et al.

most feeding situations. Lactate derived from the diet, ruminal bacterial metabolism and from
endogenous metabolism is a major glucose precursor in ruminants, supplying 9–35% of
hepatic glucose carbon, and is a key carbon intermediate in growing ruminants. Ruminants
have been shown experimentally to be capable of contributing significant quantities of glucose
through intestinal digestion and glucose absorption. This additional glucose does impact growth
and retention of body tissues.
The gut epithelia have a central function as gatekeepers for butyrate and longer-chain
SCFA and MCFA. These acids have also become the main energy substrates of gut epithelia.
There is no evidence suggesting that the rumen epithelium should have excessive requirements
for energy metabolism, but rather intraruminal (luminal) isotopic dilution techniques overesti-
mate net SCFA production because of microbial metabolism. Data on portal appearance of
SCFA corrected for PDV metabolism of arterial metabolites is therefore the best direct measure
of SCFA availability in ruminants. The average absorption of SCFA in ruminants is equivalent
to about 45% of the digestible energy intake.

23. FUTURE PERSPECTIVES


Current feeding systems often fail to meet today’s demand for accurate prediction of the
nutrient needs of animals consuming a large menu of feedstuffs under a wide array of
environmental conditions. To accomplish this task we need to understand all phases from
digestion and nutrient assimilation to the subsequent use of nutrients by various tissues. Thus,
successful models in the future will span concepts from commodity to animal product by
incorporating the metabolic transformations in between. This chapter has reviewed recent
findings on the impact of intestinal digestion on glucose availability, the contributions of
lactate to meeting the glucose needs, and how our understanding of SCFA metabolism has
been revised from long-accepted concepts. Developing models for predicting animal biolog-
ical response are dependent on findings such as these to supply quantitative information to
describe animal systems. Data are still greatly lacking for concepts as fundamental as SCFA
production and glucose absorption, and their metabolism at various stages of growth and
production. The near-global de-emphasis on agricultural production research may make these
pieces of the puzzle long in coming.

REFERENCES
Aas, M., 1971. Organ and subcellular distribution of fatty acid activating enzymes in the rat. Biochim.
Biophys. Acta 231, 32–47.
Annison, E.F., Leng, R.A., Lindsay, D.B., White, B.A., 1963. The metabolism of acetic acid, propionic
acid and butyric acid in sheep. Biochem. J. 88, 248–252.
Ash, R., Baird, G.D., 1973. Activation of volatile fatty acids in bovine liver and rumen epithelium:
evidence for control by autoregulation. Biochem. J. 136, 311–319.
Baird, G.D., Hibbitt, K.G., Lee, J., 1970. Enzymes involved in acetoacetate formation in varous bovine
tissues. Biochem. J. 117, 703–709.
Baker, S.K., Dijkstra, J., 1999. Dynamic aspects of the microbial ecosystem of the reticulo-rumen.
In: Jung, H.-J.G., Fahey, G.C. Jr. (Eds.), Nutritional Ecology of Herbivores. American Society of
Animal Science, Savoy, IL, pp. 261–311.
Balcells, J., Seal, C.J., Parker, D.S., 1995. Effect of intravenous glucose infusion on metabolism of portal-
drained viscera in sheep fed a cereal/straw-based diet. J. Anim. Sci. 73, 2146−2155.
Baldwin, R.L., Jesse, B.W., 1996. Propionate modulation of ruminal ketogenesis. J. Anim. Sci. 74,
1694–1700.
Baldwin, R.L., McLeod, K.R., 2000. Effects of diet forage:concentrate ratio and metabolizable energy
intake on isolated rumen epithelial cell metabolism in vitro. J. Anim. Sci. 78, 771–783.
Splanchnic carbohydrate and energy metabolism 427

Barcroft, J., McAnally, R.A., Phillipson, A.T., 1944. Absorption of volatile acids from the alimentary
tract of the sheep and other animals. J. Exp. Biol. 20, 120–129.
Bauer, M.L., Harmon, D.L., McLeod, K.R., Huntington, G.B., 1995. Adaptation to small intestinal starch
assimilation and glucose transport in ruminants. J. Anim. Sci. 73, 1828–1838.
Beck, U., Emmanuel, B., Giesecke, D., 1984. The ketogenic effect of glucose in rumen epithelium of
ovine (Ovis aries) and bovine (Bos taurus) origin. Comp. Biochem. Physiol. B 77, 517–521.
Bergman, E.N., 1973. Glucose metabolism in ruminants as related to hypoglycemia and ketosis. Cornell
Vet. 63, 341–382.
Bergman, E.N., 1990. Energy contributions of volatile fatty acids from the gastrointestinal tract in various
species. Physiol. Rev. 70, 567–590.
Bergman, E.N., Wolff, J.E., 1971. Metabolism of volatile fatty acids by liver and portal-drained viscera
in sheep. Amer. J. Physiol. 221, 586–592.
Berthelot, V., Pierzynowski, S.G., Sauvant, D., Kristensen, N.B., 2002. Hepatic metabolism of propionate
and methylmalonate in growing lambs. Livest. Prod. Sci. 74, 33–43.
Britton, R., Krehbiel, C., 1993. Nutrient metabolism by gut tissues. J. Dairy Sci. 76, 2125–2131.
Bugaut, M., 1987. Occurrence, absorption and metabolism of short chain fatty acids in the digestive tract
of mammals. Comp. Biochem. Physiol. B 86, 439–472.
Bush, R.S., Milligan, L.P., 1971. Enzymes of ketogenesis in bovine rumen epithelium. Can. J. Anim. Sci.
51, 129–133.
Campagnari, F., Webster, L.T. Jr., 1963. Purification and properties of acetyl coenzyme A synthetase from
bovine heart mitochondria. J. Biol. Chem. 238, 1628–1633.
Cappelli, P.F., Seal, C.J., Parker, D.S., 1997. Glucose and [13C]leucine metabolism by the portal-drained
viscera of sheep fed on dried grass with acute intravenous and intraduodenal infusions of glucose.
Brit. J. Nutr. 78, 931–946.
Cook, R.M., Liu, S.-C.C., Quraishi, S., 1969. Utilization of volatile fatty acids in ruminants. III.
Comparison of mitochondrial acyl coenzyme A synthetase activity and substrate specificity in differ-
ent tissues. Biochemistry 8, 2966–2969.
Counotte, G.H., Prins, R.A., 1981. Regulation of lactate metabolism in the rumen. Vet. Res. Commun. 5,
101–115.
Cremin, J.D. Jr., McLeod, K.R., Harmon, D.L., Goetsch, A.L., Bourquin, L.D., Fahey, G.C. Jr., 1995. Portal
and hepatic fluxes in sheep and concentrations in cattle ruminal fluid of 3-(4-hydroxyphenyl)propionic,
benzoic, 3-phenylpropionic, and trans-cinnamic acids. J. Anim. Sci. 73, 1766–1775.
Crichlow, E.C., 1988. Ruminal lactic acidosis: forestomach epithelial receptor activation by undissoci-
ated volatile fatty acids and rumen fluids collected during loss of reticuloruminal motility. Res. Vet.
Sci. 45, 364–368.
Danfær, A., 1990. A Dynamic Model of Nutrient Digestion and Metabolism in Lactating Dairy Cows.
Beretning fra Statens Husdyrbrugsforsøg, Frederiksberg Bogtrykkeri, Denmark.
De Visser, H., Valk, H., Klop, A., Van der Meulen, M.J., Bakker, J.G., Huntington, G.B., 1997. Nutrient
fluxes in splanchnic tissue of dairy cows: influence of grass quality. J. Dairy Sci. 80, 1666–1673.
Dietschy, J.M., 1978. General principles govering movement of lipids across biological membranes. In:
Dietschy, J.M., Gotto, A.M.J., Ontko, J.A. (Eds.), Disturbances in Lipid and Lipoprotein Metabolism.
American Physiological Society, Baltimore, MD, pp. 1–28.
Dunkin, A.C., Campbell, R.G., 1982. Some aspects of the partition of metabolizable energy in the youg
pig. In: Ekern, A., Sundstol, F. (Eds.), Proceedings of the 9th symposium on Energy Metabolism of
Farm Animals, Lillehammer, Norway. EAAP Publ. No. 29, pp. 198–201.
Dunlop, R.H., 1972. Pathogenesis of ruminant lactic acidosis. Adv. Vet. Sci. Comp. Med. 16, 259–302.
Eaton, S., Bartlett, K., Pourfarzam, M., 1996. Mammalian mitochondrial β-oxidation. Biochem. J. 320,
345–357.
Eisemann, J.H., Huntington, G.B., 1993. Effects of dietary clenbuterol on net flux across the portal-
drained viscera, liver and hindquarters of steers (Bos taurus). Comp. Biochem. Physiol. C 104,
401–406.
Eisemann, J.H., Huntington, G.B., 1994. Metabolite flux across portal-drained viscera, liver, and
hindquarters of hyperinsulinemic, euglycemic beef steers. J. Anim. Sci. 72, 2919–2929.
Eisemann, J.H., Huntington, G.B., Catherman, D.R., 1996. Patterns of nutrient interchange and oxygen
use among portal-drained viscera, liver, and hindquarters of beef steers from 235 to 525 kg body
weight. J. Anim. Sci. 74, 1812–1831.
Emmanuel, B., Milligan, L.P., 1983. Butyrate:acetoacetyl-CoA transferase activity in bovine rumen
epithelium. Can. J. Anim. Sci. 63, 355–360.
428 N. B. Kristensen et al.

Emmanuel, B., Stangassinger, M., Giesecke, D., 1982. Production of D(−)-3-hydroxybutyrate by an alter-
nate mechanism in rumen epithelium of ovine (Ovis aries) and bovine (Bos taurus). Comp. Biochem.
Physiol. B 72, 415–419.
Gäbel, G., Müller, F., Pfannkuche, H., Aschenbach, J.R., 2001. Influence of isoform and DNP on butyrate
transport across the sheep ruminal epithelium. J. Comp. Physiol. B 171, 215–221.
Gálfi, P., Neogrády, S., Sakata, T., 1991. Effects of volatile fatty acids on the epithelial cell proliferation
of the digestive tract and its hormonal mediation. In: Tsuda, T., Sasaki, Y., Kawashima, R. (Eds.),
Physiological Aspects of Digestion and Metabolism in Ruminants. Academic Press, San Diego, CA,
pp. 49–59.
Giesecke, D., Beck, U., Emmanuel, B., 1985. Ketogenic regulation by certain metabolites in rumen
epithelium. Comp. Biochem. Physiol. B 81, 863–867.
Goad, D.W., Goad, C.L., Nagaraja, T.G., 1998. Ruminal microbial and fermentative changes associated
with experimentally induced subacute acidosis in steers. J. Anim. Sci. 76, 234–241.
Goosen, P.C.M., 1976. Metabolism in rumen epithelium oxidation of substrates and formation of ketone
bodies by pieces of rumen epithelium. Z. Tierphysiol. Tierernähr. Futtermittelkde. 37, 14–25.
Grigat, K.-P., Koppe, K., Seufert, C.-D., Söling, H.-D., 1979. Acetyl-coenzyme A deacylase activity in
liver is not an artifact: subcellular distribution and substrate specificity of acetyl-coenzyme A deacy-
lase activities in rat liver. Biochem. J. 177, 71–79.
Groot, P.H.E., Scholte, H.R., Hülsmann, W.C., 1976. Fatty acid activation: specificity, localization, and
function. Adv. Lipid Res. 14, 75–126.
Gross, K.L., Harmon, D.L., Avery, T.B., 1990a. Portal-drained visceral flux of nutrients in lambs fed
alfalfa or maintained by total intragastric infusion. J. Anim. Sci. 68, 214–221.
Gross, K.L., Harmon, D.L., Minton, J.E., Avery, T.B., 1990b. Effects of isoenergetic infusions of propi-
onate and glucose on portal-drained visceral nutrient flux and concentrations of hormones in lambs
maintained by total intragastric infusion. J. Anim. Sci. 68, 2566–2574.
Hamada, T., Ishii, T., Taguchi, S., 1982. Blood changes of spontaneously ketotic cows before and four
hours after administration of glucose, xylitol, 1,2-propanediol, or magnesium propionate. J. Dairy
Sci. 65, 1509–1513.
Harmon, D.L., 1986. Influence of dietary energy intake and substrate addition on the in vitro metabolism
of glucose and glutamine in rumen epithelial tissue. Comp. Biochem. Physiol. B 85, 643–647.
Harmon, D.L., Avery, T.B., 1987. Effects of dietary monensin and sodium propionate on net nutrient flux
in steers fed a high-concentrate diet. J. Anim. Sci. 65, 1610–1616.
Harmon, D.L., Britton, R.A., Prior, R.L., 1983. Influence of diet on glucose turnover and rates of gluco-
neogenesis, oxidation and turnover of D-(−)-lactate in the bovine. J. Nutr. 113, 1842–1850.
Harmon, D.L., Britton, R.A., Prior, R.L., 1984. In vitro rates of oxidation and gluconeogenesis from
L(+)- and D(−)lactate in bovine tissues. Comp. Biochem. Physiol. B 77, 365–368.
Harmon, D.L., Britton, R.A., Prior, R.L., Stock, R.A., 1985. Net portal absorption of lactate and volatile
fatty acids in steers experiencing glucose-induced acidosis or fed a 70% concentrate diet ad libitum.
J. Anim. Sci. 60, 560–569.
Harmon, D.L., Gross, K.L., Krehbiel, C.R., Kreikemeier, K.K., Bauer, M.L., Britton, R.A., 1991.
Influence of dietary forage and energy intake on metabolism and acyl-CoA synthetase activity in
bovine ruminal epithelial tissue. J. Anim. Sci. 69, 4117–4127.
Harmon, D.L., Richards, C.J., Swanson, K.C., Howell, J.A., Matthews, J.C., True, A.D., Huntington, G.,
Gahr, S.A., Russell, R.W., 2001. Influence of ruminal and postruminal starch infusion on visceral glu-
cose metabolism in steers. In: Chwalibog, A., Jakobsen, K. (Eds.), Energy Metabolism in Animals:
Proceedings of the 15th Symposium on Energy Metabolism in Animals. Wageningen Press,
Wageningen, The Netherlands, pp. 273–276.
Henrikson, R.C., Stacy, B.D., 1971. The barrier to diffusion across ruminal epithelium: a study by electron
microscopy using horseradish peroxidase, lanthanum and ferritin. J. Ultrastruct. Res. 34, 72–82.
Herbein, J.H., Van Maanen, R.W., McGilliard, A.D., Young, J.W., 1978. Rumen propionate and blood
glucose kinetics in growing cattle fed isoenergetic diets. J. Nutr. 108, 994–1001.
Hird, F.J.R., Jackson, R.B., Weidemann, M.J., 1966. Transport and metabolism of fatty acids by isolated
rumen epithelium. Biochem. J. 98, 394–400.
Hird, F.J.R., Symons, R.H., 1961. The mode of formation of ketone bodies from butyrate by tissue from
the rumen and omasum of the sheep. Biochim. Biophys. Acta 46, 457–467.
Hull, S.R., 1996. Composition of corn steep water during steeping. J Agr. Food Chem. 44, 1857–1863.
Huntington, G.B., 1997. Starch utilization by ruminants: from basics to the bunk. J. Anim. Sci. 75, 852–867.
Splanchnic carbohydrate and energy metabolism 429

Huntington, G.B., Prior, R.L., 1983. Digestion and absorption of nutrients by beef heifers fed a high con-
centrate diet. J. Nutr. 113, 2280–2288.
Huntington, G.B., Reynolds, P.J., 1986. Net absorption of glucose, L-lactate, volatile fatty acids, and
nitrogenous compounds by bovine given abomasal infusions of starch or glucose. J. Dairy Sci. 69,
2428–2436.
Huntington, G.B., Prior, R.L., Britton, R.A., 1980. Glucose and lactate absorption and metabolic interre-
lationships in lambs switched from low to high concentrate diets. J. Nutr. 110, 1904–1913.
Huntington, G.B., Prior, R.L., Britton, R.A., 1981. Glucose and lactate absorption and metabolic interre-
lationships in steers changed from low to high concentrate diets. J. Nutr. 111, 1164–1172.
Huntington, G.B., Reynolds, C.K., Stroud, B.H., 1989. Techniques for measuring blood flow in splanch-
nic tissues of cattle. J. Dairy Sci. 72, 1583–1595.
Huntington, G.B., Varga, G.A., Glenn, B.P., Waldo, D.R., 1988. Net absorption and oxygen consumption
by Holstein steers fed alfalfa or orchardgrass silage at two equalized intakes. J. Anim. Sci. 66,
1292–1302.
Jesse, B.W., Solomon, R.K., Baldwin, R.L., 1992. Palmitate metabolism by isolated sheep rumen epithe-
lial cells. J. Anim. Sci. 70, 2235–2242.
Juel, C., 1997. Lactate-proton cotransport in skeletal muscle. Physiol. Rev. 77, 321–358.
Kohn, R.A., Kalscheur, K.F., Hanigan, M., 1998. Evaluation of models for balancing the protein require-
ments of dairy cows. J. Dairy Sci. 81, 3402–3414.
Kramer, T., Michelberger, T., Gürtler, H., Gäbel, G., 1996. Absorption of short-chain fatty acids across
ruminal epithelium of sheep. J. Comp. Physiol. B 166, 262–269.
Krehbiel, C.R., Harmon, D.L., Schneider, J.E., 1992. Effect of increasing ruminal butyrate on portal and
hepatic nutrient flux in steers. J. Anim. Sci. 70, 904–914.
Kreikemeier, K.K., Harmon, D.L., Brandt, R.T. Jr., Avery, T.B., Johnson, D.E., 1991. Small intestinal
starch digestion in steers: effect of various levels of abomasal glucose, corn starch and corn dextrin
infusion on small intestinal disappearance and net glucose absorption. J. Anim. Sci. 69, 328–338.
Kristensen, N.B., Danfær, A., 2001. The relationship between gastrointestinal production and portal
absorption of short-chain fatty acids in ruminants. In: Chwalibog, A., Jakobsen, K. (Eds.), Energy
Metabolism in Animals: Proceedings of the 15th Symposium on Energy Metabolism in Animals.
Wageningen Press, Wageningen, The Netherlands, pp. 277–280.
Kristensen, N.B., Danfær, A., Agergaard, N., 1996a. Diurnal patterns of ruminal concentrations and
portal appearance rates of short-chain fatty acids in sheep fed a hay or a concentrate/straw diet in two
meals daily. Acta Agr. Scand. Sect. A 46, 227–238.
Kristensen, N.B., Danfær, A., Agergaard, N., 1998. Absorption and metabolism of short-chain fatty acids
in ruminants. Arch. Tierernähr. 51, 165–175.
Kristensen, N.B., Danfær, A., Tetens, V., Agergaard, N., 1996b. Portal recovery of intraruminally infused
short-chain fatty acids in sheep. Acta Agr. Scand. Sect. A 46, 26–38.
Kristensen, N.B., Gäbel, G., Pierzynowski, S.G., Danfær, A., 2000a. Portal recovery of short-chain
fatty acids infused into the temporarily-isolated and washed reticulo-rumen of sheep. Brit. J. Nutr.
84, 477–482.
Kristensen, N.B., Pierzynowski, S.G., Danfær, A., 2000b. Net portal appearance of volatile fatty acids in
sheep intraruminally infused with mixtures of acetate, propionate, isobutyrate, butyrate, and valerate.
J. Anim. Sci. 78, 1372–1379.
Kristensen, N.B., Pierzynowski, S.G., Danfær, A., 2000c. Portal-drained visceral metabolism of
3-hydroxybutyrate in sheep. J. Anim. Sci. 78, 2223–2228.
Kristensen, N.B., Steensen, T.H., Pierzynowski, S.G., Danfær, A., 2001. Metabolism of 2-13C-propionate
in the rumen epithelium of sheep. J. Anim. Sci. 79, Suppl. 1, 287.
Kung, L. Jr., Ranjit, N.K., 2001. The effect of Lactobacillus buchneri and other additives on the
fermentation and aerobic stability of barley silage. J. Dairy Sci. 84, 1149–1155.
Kung, L. Jr., Robinson, J.R., Ranjit, N.K., Chen, J.H., Golt, C.M., Pesek, J.D., 2000. Microbial popula-
tions, fermentation end-products, and aerobic stability of corn silage treated with ammonia or a
propionic acid-based preservative. J. Dairy Sci. 83, 1479–1486.
Larson, H.J., Jacobson, N.L., McGilliard, A.D., Allen, R.S., 1956. Utilization of carbohydrates introduced
directly into the omaso-abomasal area of the stomach of cattle of various ages. J. Anim. Sci. 44, 321–330.
Le Bars, H., Lebrument, J., Nitescu, R., Simonnet, H., 1954. Recherches sur la motricité du rumen chez
les petits ruminants. IV. Action de l’injection intraveineuse d’acides gras à courte chaîne. Bull. Acad.
Vet. 27, 53–67.
430 N. B. Kristensen et al.

Leighton, B., Nicholas, A.R., Pogson, C.I., 1983. The pathway of ketogenesis in rumen epithelium of the
sheep. Biochem. J. 216, 769–772.
Leng, R.A., 1970. Glucose synthesis in ruminants. Adv. Vet. Sci. Comp. Med. 14, 209–260.
Leng, R.A., Annison, E.F., 1963. Metabolism of acetate, propionate and butyrate by sheep-liver slices.
Biochem. J. 86, 319–327.
Manns, J.G., Boda, J.M., 1967. Insulin release by acetate, propionate, butyrate, and glucose in lambs and
adult sheep. Amer. J. Physiol. 212, 747–755.
Masson, M.J., Phillipson, A.T., 1952. The composition of the digesta leaving the abomasum of sheep.
J. Physiol. 116, 98–111.
McDonald, P., Henderson, A.R., Heron, S.J.E., 1991. The Biochemistry of Silage, 2nd Edition
Chalcombe Publications, Bucks, UK.
McLeod, K.R., Baldwin, R.L., Harmon, D.L., Richards, C.J., Rumpler, W.V., 2001. Influence of ruminal
and postruminal starch infusion on energy balance in growing steers. In: Chwalibog, A., Jakobsen, K.
(Eds.), Energy Metabolism in Animals: Proceedings of the 15th symposium on Energy Metabolism
in Animals. Wageningen Press, Wageningen, The Netherlands, pp. 385–388.
McLeod, K.R., Bauer, M.L., Harmon, D.L., Reynolds, C.K., Mitchell, G.E. Jr., 1997. Effects of exoge-
nous somatostatin and cysteamine on net nutrient flux across the portal-drained viscera and liver of
sheep during intraduodenal infusion of starch hydrolysate and casein. J. Anim. Sci. 75, 3026–3037.
Müller, F., Huber, K., Pfannkuche, H., Aschenbach, J.R., Breves, G., Gäbel, G., 2001. Functional char-
acterization of the sheep ruminal monocarboxylate transporter (MCT1) using cultured ruminal
epithelial cells. Eur. J. Physiol. 441, Suppl., R248.
Nozière, P., Martin, C., Rémond, D., Kristensen, N.B., Bernard, R., Doreau, M., 2000. Effect of compo-
sition of ruminally-infused short-chain fatty acids on net fluxes of nutrients across portal-drained
viscera in underfed ewes. Brit. J. Nutr. 83, 521–531.
Orskov, E.R., Benzie, D., Kay, R.N., 1970. The effects of feeding procedure on closure of the
oesophageal groove in young sheep. Brit. J. Nutr. 24, 785–795.
Oshio, S., Tahata, I., 1984. Absorption of dissociated volatile fatty acids through the rumen wall of sheep.
Can. J. Anim. Sci. 64, Suppl., 167–168.
Pearce, J., Piperova, L.S., 1984. The effects of duodenal glucose and dextrin infusion on adipose tissue
metabolism in sheep. Comp. Biochem. Physiol. B 78, 565–567.
Pennington, R.J., 1952. The metabolism of short-chain fatty acids in the sheep 1. Fatty acid utilization
and ketone body production by rumen epithelium and other tissues. Biochem. J. 51, 251–258.
Peters, J.P., Shen, R.Y.W., Robinson, J.A., Chester, S.T., 1990. Disappearance and passage of propionic
acid from the rumen of the beef steer. J. Anim. Sci. 68, 3337–3349.
Pethick, D.W., Lindsay, D.B., 1982. Acetate metabolism in lactating sheep. Brit. J. Nutr. 48, 319–328.
Prasad, K.N., Sinha, P., 1976. Effect of sodium butyrate on mammalian cells in culture: a review. In Vitro
12, 125–132.
Price, N.T., Jackson, V.N., Halestrap, A.P., 1998. Cloning and sequencing of four new mammalian mono-
carboxylate transporter (MCT) homologues confirms the existence of a transporter family with an
ancient past. Biochem. J. 329, 321–328.
Prior, R.L., Huntington, G.B., Reynolds, P.J., 1984. Role of insulin and glucose on metabolite uptake by
the hind half of beef steers. J. Anim. Sci. 58, 1446–1453.
Rechkemmer, G., 1991. Transport of weak electrolytes. In: Schultz, S.T., Field, M., Frizzell, R.A. (Eds.),
Handbook of Physiology: The Gastrointestinal System IV. American Physiological Society, Bethesda,
MD, pp. 371–388.
Reid, J.T., White, O.D., Anrique, R., Fortin, A., 1980. Nutritional energetics of livestock: some present
boundaries of knowledge and future research needs. J. Anim. Sci. 51, 1393–1415.
Rémond, D., Ortiques, I., Jouany, J.P., 1995. Energy substrates for the rumen epithelium. Proc. Nutr. Soc.
54, 95–105.
Reynolds, C.K., Huntington, G.B., 1988a. Partition of portal-drained visceral net flux in beef steers. 1.
Blood flow and net flux of oxygen, glucose and nitrogenous compounds across stomach and post-
stomach tissues. Brit. J. Nutr. 60, 539–551.
Reynolds, C.K., Huntington, G.B., 1988b. Partition of portal-drained visceral net flux in beef steers. 2.
Net flux of volatile fatty acids, D-beta-hydroxybutyrate and L-lactate across stomach and post-
stomach tissues. Brit. J. Nutr. 60, 553–562.
Reynolds, P.J., Huntington, G.B., 1988c. Net portal absorption of volatile fatty acids and l(+)-lactate by
lactating Hostein cows. J. Dairy Sci. 71, 124–133.
Splanchnic carbohydrate and energy metabolism 431

Reynolds, C.K., Harmon, D.L., Cecava, M.J., 1994. Absorption and delivery of nutrients for milk protein
synthesis by portal-drained viscera. J. Dairy Sci. 77, 2787–2808.
Reynolds, C.K., Tyrrell, H.F., 1991. Effects of mesenteric vein L-alanine infusion on liver metabolism in
beef heifers fed on diets differing in forage:concentrate ratio. Brit. J. Nutr. 66, 437–450.
Reynolds, C.K., Tyrrell, H.F., Reynolds, P.J., 1991. Effects of diet forage-to-concentrate ratio and intake
on energy metabolism in growing beef heifers: net nutrient metabolism by visceral tissues. J. Nutr.
121, 1004–1015.
Ricks, C.A., Cook, R.M., 1981a. Partial purification of enzymes of bovine kidney mitochondria activat-
ing volatile fatty acids. J. Dairy Sci. 64, 2344–2349.
Ricks, C.A., Cook, R.M., 1981b. Regulation of volatile fatty acid uptake by mitochondrial acyl-CoA
synthetases of bovine heart. J. Dairy Sci. 64, 2336–2343.
Ricks, C.A., Cook, R.M., 1981c. Regulation of volatile fatty acid uptake by mitochondrial acyl-CoA
synthetases of bovine liver. J. Dairy Sci. 64, 2324–2335.
Robinson, A.M., Williamson, D.H., 1980. Physiological roles of ketone bodies as substrates and signals
in mammalian tissues. Physiol. Rev. 60, 143–187.
Rupp, G.P., Kreikemeier, K.K., Perino, L.J., Ross, G.S., 1994. Measurement of volatile fatty acid disap-
pearance and fluid flux across the abomasum of cattle, using an improved omasal cannulation
technique. Amer. J. Vet. Res. 55, 522–529.
Russell, J.B., Wallace, R.J., 1988. Energy yielding and consuming reactions. In: Hobson, P.N. (Ed.), The
Rumen Microbial Ecosystem. Elsevier Applied Science, London, pp. 185–215.
Russell, R.W., Gahr, S.A., 2000. Glucose Availability and Associated Metabolism: Farm Animal
Metabolism and Nutrition. CABI Publishing, Edinburgh, UK, pp. 121–147.
Russell, R.W., Moss, L., Schmidt, S.P., Young, J.W., 1986. Effects of body size on kinetics of glucose
metabolism and on nitrogen balance in growing cattle. J. Nutr. 116, 2229–2243.
Sander, E.G., Warner, R.G., Harrison, H.N., Loosli, J.K., 1959. The stimulatory effect of sodium butyrate and
sodium propionate on the development of rumen mucosa in the young calf. J. Dairy Sci. 42, 1600–1605.
Schaadt, H., 1968. Effects of maturity, fermentation time and urea treatment on D(−) and L(+) lactate in
corn silage. J. Dairy Sci. 51, 802–805.
Schmidt, S.P., Keith, R.K., 1983. Effects of diet and energy intake on kinetics of glucose metabolism in
steers. J. Nutr. 113, 2155–2163.
Seal, C.J., Parker, D.S., 1994. Effect of intraruminal propionic acid infusion on metabolism of mesen-
teric- and portal-drained viscera in growing steers fed a forage diet. I. Volatile fatty acids, glucose,
and lactate. J. Anim. Sci. 72, 1325–1334.
Seal, C.J., Parker, D.S., 2000. Influence of gastrointestinal metabolism on substrate supply to the liver.
In: Cronjé, P.B. (Ed.), Ruminant Physiology: Digestion, Metabolism, Growth and Reproduction.
CABI Publishing, Wallingford, UK, pp. 131–148.
Seal, C.J., Reynolds, C.K., 1993. Nutritional implication of gastrointestinal and liver metabolism in rumi-
nants. Nutr. Res. Rev. 6, 185–208.
Sehested, J., Diernæs, L., Møller, P.D., Skadhauge, E., 1999a. Ruminal transport and metabolism of
short-chain fatty acids (SCFA) in vitro: effect of SCFA chain length and pH. Comp. Biochem.
Physiol. A 123, 359–368.
Sehested, J., Diernæs, L., Møller, P.D., Skadhauge, E., 1999b. Transport of butyrate across the isolated
bovine rumen epithelium: interaction with sodium, chloride and bicarbonate. Comp. Biochem.
Physiol. A 123, 399–408.
Sheperd, A.C., Maslanka, M., Quinn, D., Kung, L. Jr., 1995. Additives containing bacteria and enzymes
for alfalfa silage. J. Dairy Sci. 78, 565–572.
Shirazi-Beechey, S.P., Kemp, R.B., Dyer, J., Beechey, R.B., 1989. Changes in the functions of the intes-
tinal brush border membrane during the development of the ruminant habit in lambs. Comp.
Biochem. Physiol. B 94, 801–806.
Siciliano-Jones, J., Murphy, M.R., 1989. Production of volatile fatty acids in the rumen and cecum-colon
of steers as affected by forage:concentrate and forage physical form. J. Dairy Sci. 72, 485–492.
Stangassinger, M., Beck, U., Emmanuel, B., 1979. Is glucose ketogenic in rumen epithelium? Ann. Rech.
Vét. 10, 413–416.
Stangassinger, M., Giesecke, D., 1979. Quantitative measurement of gluconeogenesis from isobutyrate
in sheep. Arch. Int. Physiol. Biochim. 87, 265–274.
Stern, J.R., Coon, M.J., Del Campillo, A., Schneider, M.C., 1956. Enzymes of fatty acid metabolism. IV.
Preparation and properties of coenzyme A transferase. J. Biol. Chem. 221, 15–31.
432 N. B. Kristensen et al.

Steven, D.H., Marshall, A.B., 1970. Organization of the rumen epithelium. In: Phillipson, A.T. (Ed.),
Physiology of Digestion and Metabolism in the Ruminant. Oriel Press, Newcastle-upon-Tyne, UK,
pp. 80–100.
Stevens, C.E., 1970. Fatty acid tranport through the rumen epithelium. In: Phillipson, A.T. (Ed.),
Physiology of Digestion and Metabolism in the Ruminant. Oriel Press, Newcastle upon Tyne, UK,
pp. 101–112.
Stevens, C.E., Stettler, B.K., 1967. Evidence for active transport of acetate across bovine rumen epithe-
lium. Amer. J. Physiol. 213, 1335–1339.
St. Jean, G.D., Rings, D.M., Schmall, L.M., Hoffsis, G.F., Hull, B.L., 1989. Jejunal mucosal lactase
activity from birth to 3 weeks in conventionally raised calves. Amer. J. Vet. Res. 50, 1496–1498.
Sutton, J.D., 1985. Digestion and absorption of energy substrates in the lactating cow. J. Dairy Sci. 68,
3376–3393.
Synderman, S.E., Sansariq, C., Middleton, B., 1998. Succinyl-CoA:3-ketoacid CoA-transferase deficiency.
Pediatrics 101, 709–711.
Taniguchi, K., Huntington, G.B., Glenn, B.P., 1995. Net nutrient flux by visceral tissues of beef steers
given abomasal and ruminal infusions of casein and starch. J. Anim. Sci. 73, 236–249.
Thorlacius, S.O., Lodge, G.A., 1973. Absorption of steam-volatile fatty acids from the rumen of the cow
as influenced by diet, buffers and pH. Can. J. Anim. Sci. 53, 279–288.
Topping, D.L., Clifton, P.M., 2001. Short-chain fatty acids and human colonic function: roles of resist-
ant starch and nonstarch polysaccharides. Physiol. Rev. 81, 1031–1064.
Van Maanen, R.W., Herbein, J.H., McGilliard, A.D., Young, J.W., 1978. Effects of monensin on in vivo
rumen propionate production and blood glucose kinetics in cattle. J. Nutr. 108, 1002–1007.
Varga, G.A., Tyrrell, H.F., Huntington, G.B., Waldo, D.R., Glenn, B.P., 1990. Utilization of nitrogen and
energy by Holstein steers fed formaldehyde- and formic acid-treated alfalfa or orchardgrass silage at
two intakes. J. Anim. Sci. 68, 3780–3791.
Vessy, D.A., Kelly, M., Warren, R.S., 1999. Characterization of the CoA ligases of human liver mito-
chondria catalyzing the activation of short- and medium-chain fatty acids and xenobiotic carboxylic
acids. Biochim. Biophys. Acta 1428, 455–462.
Webster, L.T. Jr., Gerowin, L.D., Rakita, L., 1965. Purification and characteristics of a butyryl coenzyme A
synthetase from bovine heart mitochondria. J. Biol. Chem. 240, 29–33.
Weekes, T.E.C., 1974. The in vitro metabolism of propionate and glucose by the rumen epithelium.
Comp. Biochem. Physiol. B 49, 393–406.
Weekes, T.E.C., Webster, A.J.F., 1975. Metabolism of propionate in the tissues of the sheep gut. Brit. J. Nutr.
33, 425–438.
Weigand, E., Young, J.W., McGilliard, A.D., 1972. Extent of propionate metabolism during absorption
from the bovine ruminoreticulum. Biochem. J. 126, 201–209.
Weigand, E., Young, J.W., McGilliard, A.D., 1975. Volatile fatty acid metabolism by rumen mucosa from
cattle fed hay or grain. J. Dairy Sci. 58, 1294–1300.
Williamson, D.H., Bates, M.W., Krebs, H.A., 1968. Activity and intracellular distribution of enzymes of
ketone-body metabolism in rat liver. Biochem. J. 108, 353–361.
Young, J.W., 1976. Gluconeogenesis in cattle: significance and methodology. J. Dairy Sci. 60, 1–15.
Young, J.W., Thorp, S.L., De Lumen, H.Z., 1969. Activity of selected gluconeogenic and lipogenic
enzymes in bovine rumen mucosa, liver and adipose tissue. Biochem. J. 114, 83–88.
PART V
Methodology
18 Methodological approaches
to metabolism research

X. Guan and D. G. Burrin

USDA/ARS Children’s Nutrition Research Center, Department of Pediatrics,


Baylor College of Medicine, Houston, TX 77030, USA

Advances in molecular biology and stable isotope techniques during the last decade have led
to an explosion of research aimed at understanding the biological basis of metabolomics from
the level of systemic physiology, to intermediary metabolism, to molecular regulation of crit-
ical proteins, and on down to genomic expression. We shall highlight principles, approaches,
and applications of these cutting-edge molecular and metabolic techniques.

1. GENETICALLY ENGINEERED ANIMAL MODELS


FOR METABOLISM RESEARCH
With the development of transgenic technologies (gene overexpression, knockout, and condi-
tional expression), one is able to explore physiological roles and metabolic functions of
specific genes and to identify individual proteins involved in the control of specific aspects of
metabolism.

1.1. Transgenic techniques

Conventional transgenic technologies (gene overexpression and knockout) are invaluable for
modeling genetic disorders and addressing developmental questions. However, this “all or
nothing” mode is inflexible and cannot be used to fully answer subtle metabolic questions. In
order to obtain precise information about the roles of a specific gene in a specific cell type at
a critical stage of disease or development, conditional transgenic techniques that allow flexi-
ble spatial and temporal control of gene deletion or expression in transgenic animals must be
used (Ryding et al., 2001). In these systems, the switching “on” or “off” of the expression of
a particular gene is conditional upon exposure to a specific stimulus (Ryding et al., 2001).
Three approaches have been used to inhibit specific gene expression in mammalian systems.
First, the most common approach is specific gene ablation by homologous recombination in
embryonic stem (ES) cells (Bronson and Smithies, 1994) and then reproduction of animals

Biology of Metabolism in Growing Animals


D.G. Burrin and H. Mersmann (Eds.)
435 © 2005 Elsevier Limited. All rights reserved.
436 X. Guan and D. G. Burrin

without expression of the specific gene. Targeted gene disruption by the homologous recom-
bination technique takes advantage of the fact that pluripotent ES cells derived from mouse
blastocysts can be cultured in vitro and remain viable for differentiation after injection into a
different embryo. The most commonly used ES cells are those derived from mice that have an
agouti coat color. These ES cells can be microinjected into embryos obtained from mice that
have a black coat color. Offspring with a high degree of agouti coat color, indicating the trans-
mission of ES cell-derived genes, can then be crossbred to obtain mice with a genetic
background identical to that of the ES cells. Therefore, mice with specific gene modifications
can be obtained by manipulation of the ES cell genome. Modification of specific genes in the
ES cell genome depends on the ability of transfected DNA to recombine with the homologous
gene in the chromosome. Isogenic DNA for the targeting construct is used to maximize
hybridization of the targeting DNA to the proper gene locus in the chromosome. A selectable
gene marker, such as the neomycin-resistant gene, is inserted into an exon to disrupt the coding
sequence of the gene of interest. The chimeric targeting gene construct is used to transfect ES
cells. Homologous recombination of the transfected DNA with chromosomal DNA at the
target locus results in the replacement of a portion of the endogenous gene with the targeting
construct, thus disrupting the coding sequence and inactivation of the endogenous gene. The
use of the selectable gene marker allows the selection for cells that have taken up and expressed
the transfected DNA. Growth of the ES cells in the presence of antibiotic indicates the inte-
gration of the transfected DNA into the ES cell genome. However, there are two limitations in
this approach: the low rate of homologous recombination in mammalian cells and the high
rate of random (nontargeted) integration of the vector DNA. Chimeric nucleases and triplex-
forming oligonucleotides may increase homologous recombination and decrease random
integration in cells (Vasquez et al., 2001).
Secondly, anti-sense DNA or RNA that inhibits gene expression by complementation to
single-stranded mRNA (Izant and Weintraub, 1985), and trans-splicing ribozymes (Kohler
et al., 1999) that catalyze RNA hydrolysis in a sequence-specific manner, have been used
successfully to abolish gene expression in mammalian cells. Anti-sense RNA is useful for
suppressing the expression of specific genes in vivo. The anti-sense plasmid construct can be
introduced into eukaryotic cells by transfection or microinjection. Anti-sense transcripts com-
plementary to 5′ untranslated target gene mRNA specifically suppress gene activity or direct
against the protein-coding domain alone. Recently, trans-splicing ribozymes have been
employed to repair mutant mRNAs in vivo. These trans-splicing ribozymes contain catalytic
sequences derived from a self-splicing group I intron, which have been adapted to a chosen
target mRNA by fusion of a region of extended complementarity to the target RNA and pre-
cise alteration of the guide sequences required for substrate recognition. The improved
trans-splicing ribozymes may be tailored for virtually any target RNA, and provide a new tool
for triggering gene expression in specific cell types.
Thirdly, RNA interference (RNAi), an evolutionarily conserved pathway, uses these small
interfering RNAs to degrade mRNAs before translation. Recently, RNAi has emerged as a
specific and efficient method to silence gene expression in mammalian cells and to probe
gene function on a whole-genome scale either by transfection of short interfering RNAs or by
transcription of short hairpin RNAs (Hammond et al., 2001; Hannon, 2002; McCaffrey et al.,
2002). Short interfering RNAs typically consist of two 21-nucleotide (nt) single-stranded
RNAs that form a 19-bp duplex with 2-nt 3′ overhangs. Its antisense strand is used by an
RNAi silencing complex to guide mRNA cleavage, so promoting mRNA degradation. It is
certain that the ability of RNAi technology to silence specific genes will transform future
studies of cellular systems and biology in mammalian cells (McManus and Sharp, 2002).
Methodological approaches to metabolism research 437

1.2. Applications of genetically engineered animal models

The use of nitric oxide synthase (NOS) gene knockout animals has helped elucidate the roles of
different NOS isoforms in the synthesis and function of nitric oxide. Nitric oxide from neuronal
NOS is a major inhibitory neurotransmitter; nitric oxide from endothelial NOS regulates blood
flow under physiological conditions; and nitric oxide from inducible NOS causes hypotension
during severe inflammatory conditions. Moreover, nitric oxide from each isoform has unique
roles in tissue injury and inflammation. The neuronal NOS-deficient mice develop gastric
dilatation and stasis; the endothelial NOS-deficient mice develop hypotension and lack
vasodilatory responses to injury; and inducible NOS-deficient mice are more susceptible to
inflammatory damage but more resistant to septic shock (Mashimo and Goyal, 1999). This
example clearly demonstrates the enormous potential of genetically engineered mice lacking
specific genes in elucidation of mechanisms specific in physiology and pathology.
Another example is that gene silencing by siRNAs has provided insights into insulin regu-
lation of glucose uptake and glycogen synthesis. The serine/threonine protein kinase Akt has
been proposed to mediate insulin signaling in several processes. However, it is unclear if Akt
is involved in insulin-stimulated glucose uptake, and which isoforms of Akt are responsible
for each insulin action. Recently, experiments with isoform-specific siRNA have revealed that
Akt2, and Akt1 to a lesser extent, has an essential role in insulin-stimulated glucose trans-
porter-4 translocation and 2-deoxyglucose uptake in both Chinese hamster ovary cells and
3T3-L1 adipocytes, while Akt1 and Akt2 contribute equally to insulin-stimulated glycogen
synthesis. These data suggest a prerequisite role of Akt in insulin-stimulated glucose uptake
and distinct functions among Akt isoforms (Katome et al., 2003).

2. GENE EXPRESSION TECHNIQUES


FOR METABOLISM RESEARCH
The transcription of genomic DNA to produce mRNA is the first step in the process of pro-
tein synthesis, and differences in gene expression are indicative of cellular responses to
environmental stimuli and perturbations and are responsible for both morphological and phe-
notypic differences between tissues and stages of development. Knowing when, where, and
to what extent a gene is expressed is central to understanding the activity and biological roles
of its encoded protein. In addition, changes in the multigene patterns of expression can pro-
vide clues about regulatory mechanisms and broader cellular functions and biochemical
pathways (Lockhart and Winzeler, 2000).

2.1. Gene expression techniques


There are many techniques for measuring gene expression. Both conventional methods
(including Northern blots, RNase protection assay, in situ hybridization, and RT-PCR) and
DNA microarrays have been employed to measure expression levels of specific genes, to
characterize global expression profiles, and to screen for differences in mRNA abundance.
These conventional methods are simply used in a more targeted fashion to follow up on the
specific genes, pathways, and mechanisms implicated by the microarrays.

2.1.1. DNA microarray analysis for screening global gene expression profile

DNA microarrays are a miniaturized, ordered arrangement of nucleic acid fragments from indi-
vidual genes located at defined positions on a solid support, enabling the expression analysis of
438 X. Guan and D. G. Burrin

thousands of genes in parallel by specific hybridization (Arcellana-Panlilio and Robbins,


2002). This technology is a powerful tool for rapid, comprehensive, and quantitative analysis
of global gene expression profiles of normal/disease statuses and developmental processes
(Bednar, 2000; Lockhart and Winzeler, 2000). In general, there are two kinds of DNA
microarrays: cDNA arrays and oligonucleotide arrays. For cDNA arrays, the nucleic acid
fragments are spotted robotically onto a glass slide. The cDNA used for spotting are usually
derived by PCR amplification of cDNA libraries. For oligonucleotide arrays, oligonucleotides
are synthesized in situ by photolithography. Gene expression analysis using DNA micro-
arrays is based on the competitive hybridization of differently labeled populations of cDNAs.
Fluorescent dyes, usually Cy3 and Cy5, are used to label cDNA pools reverse transcribed
from different mRNA samples (prepared from tissues or cells). The labeled cDNAs are
applied to the microarray and allowed to simultaneously hybridize under conditions analo-
gous to those established for Southern blotting. After washing off nonspecific hybridization,
the slide is read in a confocal laser scanner that can differentiate between Cy3 and Cy5 signals.
Because hybridization is governed by the recognition rules, the signal intensity at each posi-
tion gives not only a measurement of the number of molecules bound, but also the likely
identity of the molecules. Thus, the relative intensity of Cy5/Cy3 signal for each gene is used
to assess the relative abundance of a specific mRNA. It should be noted that the extent of
hybridization on a DNA microarray is influenced by time, concentration of solution-phase
cDNA probes, and length of the arrayed DNA sequences (Stillman and Tonkinson, 2001).

2.1.2. mRNA quantitative techniques for measuring specific gene expression

2.1.2.1. Northern blotting analysis The Northern blotting analysis separates RNA
species on the basis of size by denaturing gel electrophoresis followed by transfer of the RNA
onto a nylon membrane by capillary, vacuum, pressure, or electrical-assisted blotting. The
RNA is then irreversibly bound to the membrane by exposure to short-wave ultraviolet light
or by heating at 80°C in a vacuum oven. The RNA sequences of interest are detected on the
membrane by hybridization to a specific labeled probe. Probes for Northern blot detection
generally contain full or partial cDNA sequences and may be labeled by enzymatic incorpo-
ration of radiolabeled (32P) nucleotides or with nucleotides conjugated to haptens such as
biotin or digoxigenin. After washing off the unbound and nonspecifically bound probe, the
hybridization signal is generally revealed by autoradiography or immunological detection
after antibody incubation. Autoradiograph band intensities may be quantified by densitometry,
by direct measurement of hybridized radiolabeled probe via storage phosphor imaging, or by
scintillation counting of excised bands (for the technique in detail, see Rapley and Walker,
1998). The band identified by the probe indicates the size of the mRNA, and the intensity of
the band corresponds to the relative abundance.
The Northern blotting analysis can detect the steady-state level of a specific mRNA
sequence in the sample. Association of the mRNA expression and the metabolic/physiological
state provides important clues regarding gene regulation, developmental characterization, and
responsiveness to stimuli. The abundance of mRNA is controlled by three major factors: gene
transcription, mRNA processing and transport, and mRNA stability. More sensitive methods
can be used for the analysis of rare transcripts including RT-PCR and RNase protection assay.
However, the Northern blotting analysis is the only method that can determine mRNA size.
In general, this method is semi-quantitative if a standard is used, and is suitable for deter-
mining relative abundances of mRNA species. To compare the relative abundances, each
sample on a membrane must be hybridized with a probe for the specific mRNA of interest and
Methodological approaches to metabolism research 439

a probe for an endogenous internal control. Constitutively expressed “housekeeping” genes,


such as β-actin, cyclophilin, or glyceraldehyde-3-phosphate dehydrogenase (GAPDH), and a
constant level of 18S rRNA are used as the internal control.
Variations of the Northern blotting technology, such as dot blots, slot blots, and fast blots,
have been developed to simplify blot preparation and improve quantitative analysis. These
techniques involve applying an RNA sample (dot and slot blots) or cell extract (fast blot)
directly to the membrane without prior size fractionation on a gel. It is imperative that the
probe used for dot/slot blot analyses is specific for the target mRNA without cross-hybridization
or nonspecific hybridization to other sequences.

2.1.2.2. RNase protection assay Ribonuclease (RNase) protection assay is a technique


used for detection and quantitative analysis of specific RNAs. In principle, cellular mRNA is
hybridized with a gene-specific labeled single-stranded complementary RNA (labeled with
32P). After the hybridization, all unbound single-stranded RNA molecules are degraded by
single-stranded-specific ribonuclease. The protected double-stranded (i.e. hybridized) frag-
ments are separated by denaturing polyacrylamide gel electrophoresis, detected by exposure
on x-ray film, and quantified by densitometry, or quantified by excising and scintillation
counting the region of the gel that contain the protected fragments. The size of each protected
fragment may be derived from the standard marker and the intensity of the bands directly
corresponds with the absolute concentration of the specific mRNA. Unlike Northern blots, the
size of product by the RNase protection assay does not depend on the size of the target
mRNA, but on the size of the probe used in the assay. This assay can determine absolute
abundance of mRNA at relatively high sensitivity (Reue, 1998).

2.1.2.3. In situ hybridization In situ hybridization (ISH) technique allows specific


nucleic acid sequences to be detected in morphologically preserved chromosomes, cells, or
tissue sections. In combination with immunocytochemistry, this technique can relate micro-
scopic topological information to gene activity at the DNA, mRNA, and protein level.
Localization of gene expression at the mRNA level is particular important to confirm the
identity of cells expressing soluble or secreted proteins. Currently, nonradioactive labeled
cRNA probes have become more feasible for detecting target mRNA in tissue sections. For
example, digoxigenin (DIG)-labeled nucleotides may be incorporated at a defined density
into nucleic acid probes by DNA polymerases, RNA polymerases, or terminal transferase.
Usually, cRNA probes are generated by in vitro transcription from a linearized DNA tem-
plate. Hybridized DIG-labeled probes may be detected with high-affinity anti-DIG antibodies
that are conjugated to alkaline phosphatase (AP) or horseradish peroxidase. The antibodies
conjugated to AP can be visualized with colorimetric or fluorescent AP substrates. The advent
of the tyramide signal amplification (TSA) method has dramatically increased the sensitivity
of nonradioactive ISH detection. Tyramide signal amplification is based on the horseradish
peroxidase-catalyzed deposition of labeled tyramide molecules at sites of probe binding. In
contrast, typical AP substrates precipitate diffusely at sites of AP activity. Dual fluorescent
ISH and immunohistochemistry using TSA has provided a rapid and sensitive method to
compare mRNA and protein localization (Zaidi et al., 2000), which offers the ability to dis-
tinguish between the cells responsible for production of the protein and its target cells.

2.1.2.4. Quantitative RT-PCR The reverse transcription polymerase chain reaction


(RT-PCR) is the most sensitive method for the detection of low abundance of steady-state
mRNA (Wang and Brown, 1999). The RT-PCR is an in vitro method for enzymatically ampli-
fying defined sequences of RNA and permits the analysis of different samples from as little
440 X. Guan and D. G. Burrin

as one cell. There are four types of RT-PCR: relative, competitive, comparative, and real-time
RT-PCR.
The first step in the RT-PCR is the reverse transcription of the RNA template into cDNA, fol-
lowed by its exponential amplification in a PCR. The reverse transcription step can be primed
using specific primers, random hexamers, or oligo-dT primers. In theory, PCR should detect the
cDNA derived from a single mRNA molecule; but in practice, ten or more mRNA copies are
required because of the relative inefficiency of the reverse transcriptase reaction required to con-
vert mRNA to cDNA for subsequent amplification. Typically, a small amount of total RNA (1 μg
or less) is used for reverse transcription, and a fraction (1/20 to 1/50) of the resulting cDNA is
used in the PCR. The cDNA is amplified exponentially via cycles of denaturation, annealing,
and extension. Amplification products initially appear at undetectable levels, then accumulate at
a nonlinear rate within an exponential phase, and eventually reach similar levels irrespective of
initial template concentration. Thus, quantitative comparisons must be made during the expo-
nential phase. One strategy to ensure that PCR products are analyzed within the exponential
phase of the amplification reaction is to examine products at progressive cycles during the reac-
tion. This may be accomplished by the use of real-time quantitative RT-PCR, wherein the whole
reaction is monitored rather than just the end product. Real-time RT-PCR employs a fluorescent
signal to report formation of PCR product as each cycle of the amplification proceeds, coupled
with an automated PCR/fluorescent detection system (Heid et al., 1996). For absolute quantita-
tive analysis of a target mRNA, an internal control template and corresponding control probe
with a unique reporter fluorescent dye is included in each reaction tube (Gibson et al., 1996; for
a review, see Bustin, 2000). It should be noted that real-time RT-PCR quantifies steady-state
mRNA levels, which tells the researcher nothing about either transcription levels or mRNA
stability (Bustin, 2002).

2.1.2.5. The method of choice The choice of mRNA quantitative analysis method is
dependent on the study of interest. (1) The Northern blotting analysis is the first step in the
characterization of mRNA expression as it allows visualization of intact mRNA. That is
the only method providing information about the mRNA size, alternative splicing, and the
integrity of the RNA. That also allows great flexibility, as the probe used for hybridization
does not require preparation with specific cloning vectors or primers. (2) The RNase protec-
tion assay is the most useful for mapping transcript initiation and termination sites and
intron/exon boundaries, and for discriminating between related mRNAs of similar size, which
would migrate at similar positions on a Northern blot. (3) In situ hybridization is the most
complex method of all, but is the only one that allows localization of transcripts to specific
cells within a tissue. (4) In term of sensitive, specific, and reproducible quantification of
mRNA, real-time RT-PCR is the method of choice (Bustin, 2000). The RNase protection
assay and real-time RT-PCR are most readily applied to the analysis of mRNAs that have
been previously characterized and sequenced, as they require production of specific vectors
and primers for probe and control template preparation (Reue, 1998).

2.2. DNA binding assays for assessing DNA–protein interactions

It has been known that, at the simplest level, transcription of genes into mRNAs is governed
by transcription factors, which bind to cis-regulatory regions of the DNA in the vicinity of the
target gene. The current challenges, however, include an understanding of (1) which specific
cis-acting DNA sequence elements and which trans-acting factors (transcription factors) are
required for the expression of a given gene; (2) how a given set of DNA–protein interactions
Methodological approaches to metabolism research 441

regulates the expression of a tissue-specific gene; and (3) how these interactions are integrated
into the overall regulation of gene expression during development (Yang, 1998). The highly
specific interaction between a given transcription factor and its recognized binding sequence
(DNA) forms the basis for the biochemical characterization, and provides insight into the
overall molecular mechanisms controlling gene expression.

2.2.1. Electrophoretic mobility shift assay (EMSA)

This assay consists of three steps. (1) A DNA-binding protein (present in a nuclear extract) is
mixed with a 32P-labeled DNA fragment (probe). (2) The DNA–protein complexes migrate
more slowly than free, unbound DNA during electrophoresis. (3) The two bands containing
radiolabeled DNA are detected by autoradiography or a phosphor screen. In general, each
additional protein binding to a DNA–protein complex alters its electrophoretic mobility and
results in an additional retarded band. The EMSA is sufficiently sensitive for the binding of
a monoclonal antibody to the protein–DNA complex to cause a supershift band, which can
confirm the presence of a particular protein in the complex. Though it provides a quantitative
measurement of the amount of a particular DNA binding activity, the EMSA does not give
a direct readout of the DNA nucleotides that the protein recognizes.

2.2.2. DNase I protection (footprinting) assay

A specific binding protein (in a nuclear extract) binds to a specific region within a singly
end-labeled DNA fragment (probe). After digestion by DNase I, the DNA products are elec-
trophoresed in a denaturing polyacrylamide gel. In the absence of any binding protein, the
bands appear as a ladder without any interruption. However, in the presence of the specific
binding protein, some bands disappear because DNase I cannot digest the region of DNA
bound by the protein. This assay allows the determination of a short stretch of a protein-binding
site within a relatively large DNA fragment. The exact nucleotide sequence in the protected
region can readily be determined by concurrently running sequencing reactions of the same
DNA fragment alongside the DNase I digestion products.

2.2.3. Chromatin immunoprecipitation (ChIP) assay

An issue in gene transcription is the in vivo relevance of transcription factor binding sites that
have been identified in vitro. The ChIP assay is being successfully exploited to confirm
in vivo binding sites of specific transcription factors. In this assay, an antibody to a specific
DNA-binding protein is used to immunoprecipitate cross-linked protein–DNA complexes.
Then, the DNA is experimentally released from the complexes and detected by DNA
footprinting (Lee Kang et al., 2002) or DNA microarray (Weinmann et al., 2002). In combi-
nation with the DNA microarray, the ChIP assay is used to probe the genome-wide pattern
of DNA binding sites for specific transcription factors (Weinmann et al., 2002). Moreover,
this technique can distinguish the direct targets of the transcription factors from indirect
downstream effects (Shannon and Rao, 2002).

2.3. Applications: nutrient regulation of gene expression

Effects of nutrition can be exerted at many stages between transcription of the genetic
sequence and translation of a functional protein. Nutrients can influence gene expression
442 X. Guan and D. G. Burrin

through control of the regulatory signals in the untranslated regions of the gene (Hesketh
et al., 1998). In the research of nutritional control of gene expression, it is important not only
to focus on regulation through gene promoter regions but also to consider the possibility of
post-transcriptional control (Hesketh et al., 1998).
An example of nutrient regulation of gene expression is that the polyunsaturated fatty acid
(PUFA) upregulates the expression of genes encoding proteins involved in fatty acid oxida-
tion while simultaneously downregulating the expression of genes encoding proteins
involved in lipid synthesis. The PUFAs appear to regulate gene transcription by modifying
the DNA binding activity and/or the nuclear abundance of the transcription factors (Clarke,
2001). Furthermore, PUFAs govern the expression of enzymes in lipid oxidation and lipid
synthesis by two independent mechanisms: activating peroxisome proliferator-activated
receptor α (Clarke, 2001) and suppressing sterol regulatory element binding protein-1
(Xu et al., 1999, 2001). Duplus et al. (2000) have postulated multiple mechanisms for fatty
acid control of gene transcription. One of them is that the fatty acid itself or its derivative
acts as a ligand for a transcription factor, which then can bind to DNA at a fatty acid
response element in the fatty acid-responsive gene and activate or repress transcription
(Duplus et al., 2000).
Another example is that amino acid availability regulates the expression of genes encod-
ing proteins in the control of growth (Fafournoux et al., 2000). Limitation of several amino
acids greatly increases the expression of a specific gene encoding the CHOP protein, a stress-
inducible nuclear protein that dimerizes with members of the CCAAT/enhancer-binding
protein (C/EBP) family of transcription factors (Bruhat et al., 1997; Fafournoux et al., 2000).
Elevated abundance of CHOP mRNA results from an increased rate of its transcription and
an increased stability of its mRNA (Bruhat et al., 1997, 1999). The C/EBP family is involved
in the regulation of processes relevant to gene expression, energy metabolism, cellular
proliferation, and differentiation (Roesler, 2001). By forming heterodimers with members of
the C/EBP family, the CHOP protein either as an inhibitor or an activator can influence
expression of cell type-specific genes (Ubeda et al., 1996; Sok et al., 1999). In the promoter
region of the CHOP gene, an amino acid response element (AARE) is found to bind in vitro
the activating transcription factor 2, which is essential for leucine-induced transcriptional
activation of the CHOP gene (Bruhat et al., 2000). Further work will be necessary to char-
acterize the molecular steps by which the cellular amino acid availability can regulate gene
expression, particularly to determine (1) the pattern of the AARE in the regulated genes;
(2) the nature of the protein complexes bound to these elements; (3) the identity of the intra-
cellular metabolites that mediate transcriptional activation by amino acid limitation; and
(4) the signaling pathways involved in the control of translation by amino acids (Fafournoux
et al., 2000). These studies will eventually provide insight into the role of amino acids in
the regulation of cellular functions such as protein synthesis and proteolysis (Bruhat
et al., 2002).

3. PROTEIN ABUNDANCE, ACTIVITY, AND LOCALIZATION


Molecular mechanisms that govern cellular function and metabolism are controlled largely by
the structure and function of genetically encoded products, the proteins. Post-transcriptional
processing of mRNA and co-/post-translational processing of proteins lead to a fair degree of
discordance between the open reading frames predicting protein structure and the actual func-
tional product (Witzmann and Li, 2002). Consequently, it is necessary to quantitatively
Methodological approaches to metabolism research 443

measure differential expression at the protein level. Moreover, many protein-mediated


cellular functions are managed and regulated through mechanisms that do not even involve
quantitative changes in protein expression. Instead, they are the consequences of protein
interactions and chemical modifications of existing proteins (e.g. phosphorylation and glyco-
sylation). It is essential to characterize these changes of the proteins using functional and
structural proteomics. Finally, the localization of gene products, which is often difficult to
predict from the DNA sequence, can be determined experimentally only at the protein level
(Pandey and Mann, 2000).

3.1. Two-dimensional gel electrophoresis for screening protein


expression profiles

Two-dimensional (2D) gel electrophoresis/mass spectrometry can be used to visualize differ-


ential protein expression. In the 2D electrophoresis, proteins are subjected to orthogonal
separation methods, the first based on protein charge via isoelectric focusing and then by
mass in sodium dodecyl sulfate. The final product of the 2D electrophoresis separation is
essentially an in-gel array of proteins, each assuming a coordinate position corresponding to
the unique combination of isoelectric point and mass. Protein expression patterns are visual-
ized by a number of staining methods such as fluorescent staining image analysis. Finally,
the identity of the protein(s) in each spot is characterized by liquid chromatography–mass
spectrometry (fig. 1).

Fig. 1. A schematic showing the two-dimensional gel approach. Cells (or tissue) derived from two different
conditions, A and B, are harvested and the proteins solubilized. The crude protein mixture is then applied
to a “first dimension” gel strip that separates the proteins based on their isoelectric points. After this step, the
strip is subjected to reduction and alkylation and applied to a “second dimension” SDS–PAGE gel where
proteins are denatured and separated on the basis of size. The gels are then fixed and the proteins visualized
by staining methods. After staining, the resulting protein spots are recorded and quantified. The spots of
interest are then excised and subjected to mass spectrometric analysis; Reproduced with permission from
Pandey and Mann (2000).
444 X. Guan and D. G. Burrin

3.2. Protein microarrays for screening protein expression profiles

Protein microarrays are being developed for high-throughput analysis of protein expression.
There are two types of protein microarrays. One type of protein array is termed a protein
function array, and consists of thousands of native proteins, recombinant proteins, or their
domains immobilized in a defined pattern, which can be used to examine protein function
(e.g. enzymatic activity or binding property). This type of array is incubated with a cell lysate
containing putative interaction partners. After washing away unbound material, the bound
proteins are eluted and then identified by mass spectrometry (Pandey and Mann, 2000). The
other type of protein array is termed a protein-detecting array and consists of large numbers
of protein-binding agents, which allows for screening protein expression profiles under vari-
ous physiological stimuli (Kodadek, 2001). For example, based on antibody–antigen
interactions, proteins isolated from cells in a particular physiological state are bound to an
array containing specific antigens or antibodies. The extent of the specific binding is then
detected by the fluorescence assay (Haab et al., 2001) or by the enhanced chemiluminescence
assay (Huang, 2001).

3.3. Western blot analysis for measuring specific protein expression

In Western blotting, a complex protein fraction is separated by electrophoresis and the pro-
teins are transferred to a PVDF or nitrocellulose membrane, which is then hybridized with
a primary antibody and visualized using a secondary antibody conjugated with horseradish
peroxidase or alkaline phosphatase (Rapley and Walker, 1998).

3.4. Enzyme-linked immunosorbent assay (ELISA) for measuring


specific protein activity

In the ELISA, the antigen to be detected, being passively attached to the plastic surface of
microplate wells, binds specifically to an antibody conjugated with an enzyme used for detec-
tion (e.g. horseradish peroxidase or alkaline phosphatase). The antigen–enzyme linked
antibody complex is then reacted with a substrate/chromophore. The rate of color change,
resulting from substrate metabolism by the enzyme, is proportional to the amount of enzyme
in the complex. Many modified ELISAs have been developed to detect and quantify specific
proteins (Rapley and Walker, 1998).

3.5. Confocal laser scanning microscopy (CLSM) for visualizing


specific protein location

The CLSM captures only the light coming immediately from the object point in focus and
obstructs the light coming from out-of-focus areas of the sample. A laser beam is concen-
trated on a very small spot and then scans the sample in the X–Y direction. As a result, the
part corresponding to the eliminated light is darkened in the image, making it possible to
optically slice a thick tissue sample. It detects the fluorescence or transmits light from the
sample, and displays an image on the monitor. The CLSM has high contrast and superior
resolution in the light axis direction because of the use of confocal optics. In combination
with immunohistochemistry, the CLSM provides specific information about protein expres-
sion patterns at the single-cell level and may indicate molecular changes relevant to
metabolism.
Methodological approaches to metabolism research 445

4. DNA AND RNA LABELING TECHNIQUES FOR MEASURING


CELL PROLIFERATION AND APOPTOSIS IN VIVO
4.1. Cell proliferation

4.1.1. Bromodeoxyuridine (BrdU) labeling assay

During cell proliferation, the DNA has to be replicated before the cell is divided into two daugh-
ter cells (Sawada et al., 1995). Because of the positive relation between fractional rate of DNA
synthesis and proportion of new cells by counting (Macallan et al., 1998), the measurement of
DNA synthesis is very attractive for assessing cell proliferation. Therefore, cell proliferation has
been assayed by measuring incorporation of radiolabeled nucleosides (e.g. [3H]thymidine) into
DNA. The amount of [3H]thymidine incorporated into the DNA is quantified by liquid scintil-
lation counting. To avoid radioactivity hazards, a method of nonradioactive labeling of the DNA
with 5-bromo-2-deoxyuridine (BrdU, a thymidine analogue) has been developed for measur-
ing cell proliferation. It has been shown that the BrdU, like thymidine, is incorporated into
cellular DNA. The incorporated BrdU during DNA synthesis could be detected by an enzyme
immunoassay using monoclonal antibodies directed against BrdU, and used to quantify cell
proliferation (Maghni et al., 1999).

4.1.2. Stable isotopic tracer incorporation methods

DNA synthesis and breakdown have been measured by labeling DNA with pyrimidine
deoxyribonucleosides (e.g. [3H]thymidine or BrdU); these techniques can be confounded by
physiological factors other than the rates of cell proliferation and death per se (Hellerstein,
1999) and cannot be used safely in humans (Neese et al., 2002). Macallan et al. (1998) and
Martini et al. (2002) have developed a stable isotopic tracer incorporation method for meas-
uring DNA synthesis by labeling the deoxyribose moiety of purine deoxyribonucleotides
through the de novo nucleotide synthesis pathway using [2H]glucose or [U-13C6]glucose or
2H O (Macallan et al., 1998; Martini et al., 2002; Neese et al., 2002). It allows measurement
2
of stable isotope incorporation into DNA and calculation of cell proliferation and death rates
in vivo in humans and animals (Hellerstein et al., 1999; Neese et al., 2001, 2002). This method
counts cell divisions by measuring the proportion of labeled DNA strands present assuming
that each cell division in the presence of label generates two labeled DNA strands (one in each
daughter cell) (Hellerstein et al., 1999).
Compared to BrdU or [3H]thymidine labeling techniques, there are three differences
(Macallan et al., 1998): (1) This method labels deoxyribonucleotides in DNA through the
de novo nucleotide synthesis pathway instead of the nucleoside salvage pathway. The pathways
for labeling of DNA are illustrated in fig. 2. The efficiency of de novo contribution to purine
nucleosides is predictable and high in dividing cells. The activity of the de novo pathway for
purine nucleosides is relatively unaffected by extracellular nucleoside concentrations and
derives almost entirely from extracellular glucose, so that the precursor–product relationship
can be used in a predictable way across cell types (Macallan et al., 1998; Hellerstein, 1999).
(2) This method measures labeling in purine deoxyribonucleosides instead of pyrimidines
(e.g. from [3H]thymidine or BrdU) (Macallan et al., 1998). (3) BrdU is a pyrimidine nucleoside
that is used by the nucleoside salvage pathway and incorporated into DNA as a thymidine ana-
logue. The efficiency of the pyrimidine nucleoside salvage pathway is variable and influenced
by availability of extracellular nucleosides (Hellerstein, 1999). Moreover, BrdU does not
truly quantify mitotic events, but rather labels descendants of dividing cells (Hellerstein, 1999).
446 X. Guan and D. G. Burrin

Fig. 2. Labeling pathways for measuring DNA synthesis and thus cell proliferation (adapted from Neese
et al., 2002). GNG, gluconeogenesis; G6P, glucose-6-phosphate; R5P, ribose 5-phosphate; PRPP, phosphori-
bose pyrophosphate; NDP, nucleoside diphosphate; DNNS, de novo nucleotide synthesis pathway; DNPS,
de novo purine/pyrimidine synthesis pathway; RR, ribonucleotide reductase; dNTP, deoxyribonucleoside
triphosphate; dN, deoxyribonucleosides; dT, thymidine deoxyribonucleoside; BrdU, bromodeoxyuridine.

In particular, the incorporation of 2H2O into the deoxyribose moiety in newly synthesized DNA
allows safe, convenient, reproducible, and inexpensive measurement of in vivo proliferation
rates of slow-turnover cells in humans (Neese et al., 2002).
The fractional production rate of dividing cells (k, per day) can be calculated based on the
precursor–product relationship provided that blood glucose reproducibly provides about 65%
of the deoxyribose present in purine nucleotides recovered from DNA of various dividing
cells (Macallan et al., 1998; Hellerstein et al., 1999). Therefore,

k = −In (1 − (IEd2-dA/IEd2-glucose × 0.65))/t

where IEd2-dA and IEd2-glucose stand for isotopic enrichment of [2H2]deoxyadenosine and
[2H2]glucose, respectively
The absolute production rate of the specific-type cells can be derived by multiplying the k
with their pool size (cells/μl). The half-life (survival time) is indicated by dividing 0.63 by k.

4.2. Cell apoptosis

4.2.1. TUNEL method


The fragmentation of nuclear DNA is one of the endpoints in apoptotic pathways. DNA frag-
mentation can be determined by electrophoresis. However, an in situ labeling DNA method has
been developed to quantify the DNA fragmentation on the basis of the terminal deoxy-
nucleotidyl transferase enzyme reaction after adding deoxynucleotides labeled with biotin or
digoxigenin to free 3′-ends of DNA fragments. Therefore, the formation of a DNA strand break
early in apoptosis is detected by enzymatic labeling of the 3′-OH termini with modified
nucleotides, which is visualized with streptavidin or anti-digoxigenin antibodies. This method
is called terminal deoxynucleotidyl transferase nucleic acid end labeling (TUNEL). However,
it must be noted that the TUNEL will also stain necrotic cells due to extensive DNA degrada-
tion (Walker and Quirke, 2001), and thus is a marker of the apoptotic process rather than a
critical component of the death process itself. False positive staining in the TUNEL method to
Methodological approaches to metabolism research 447

detect apoptosis in the liver and intestine is caused by endogenous endonucleases and is inhib-
ited by diethyl pyrocarbonate (Stahelin et al., 1998).

4.2.2. Caspase method

The cysteine–aspartic acid specific proteases (caspases) are activated in response to different
inducers of apoptosis. The process of their activation is considered to be the key event of
apoptosis (Shi, 2002). Caspases recognize a four-amino-acid sequence on their substrate pro-
teins and target the carboxyl end of aspartic acid within the sequence. Several methods have
been developed to detect the activation of caspases. After the pro-caspases are cleaved, their
products can be revealed electrophoretically and identified on immunoblots using caspase-
specific antibodies. For example, immunostaining of active caspase-3 is a reliable indicator
of apoptotic rate (Marshman et al., 2001). Moreover, activation of caspases in situ can be
measured by immunocytochemical detection of the epitope that is characteristic of the active
form of caspases or by immunocytochemical identification of the specific cleavage products.
In addition, fluorochrome-labeled inhibitors or substrates of caspases have also been used for
measuring the activation of caspases with fluorescence microscopy and flow or laser scanning
cytometry (Darzynkiewicz et al., 2002; Smolewski et al., 2002).

5. LASER CAPTURE MICRODISSECTION TO PROCURE


PURE CELLS IN VIVO FOR MOLECULAR ANALYSIS
Laser capture microdissection (LCM) is a powerful method to procure pure populations of
targeted cells from specific microscopic regions of heterogeneous tissue sections (Emmert-
Buck et al., 1996; Bonner et al., 1997). In this technique, a transparent thermoplastic film
(ethylene vinyl acetate polymer) is applied to the surface of a tissue section mounted on a
glass slide. While the film is activated through pulsing a laser beam, it becomes focally adhe-
sive and fuses to the cells of interest. When the film is removed from the tissue section, the
selected cells remain adherent to the film. The film is then placed directly into the isolation
buffer in a microfuge tube for the DNA, RNA, or protein analysis. For the LCM, individual
cells can be identified based on histological morphology, immunophenotype, function-related
antigen expression (Fend et al., 1999), or electronic images from serial sections (Wong et al.,
2000). This technique allows in vivo analysis of tissue-, cell-, and function-specific molecu-
lar analysis. In combination with high-density oligonucleotide microarray, LCM-procured
cells have been used to obtain gene expression profiles from a discrete cell population (Luzzi
et al., 2001). This technique can be further coupled with real-time quantitative RT-PCR to
quantify mRNA abundance (Betsuyaku et al., 2001), proteomic-based approaches (e.g.
2D-PAGE) to analyze protein expression (Craven and Banks, 2002; Craven et al., 2002), and
biochemical assays to measure cellular metabolite concentrations and enzyme activities
(Simone et al., 2000a,b; Stappenbeck et al., 2002).

6. STABLE ISOTOPIC TRACER TECHNIQUES FOR MEASURING


PROTEIN SYNTHESIS AND BREAKDOWN IN VIVO
Stable isotopically labeled tracer techniques have been used in the research of protein (amino
acid, AA), lipid, and carbohydrate metabolism. The principles and practice of stable isotope
tracer methodology have been introduced in detail (Wolfe, 1992). New developments and
techniques will be highlighted in this section.
448 X. Guan and D. G. Burrin

6.1. Whole-body protein kinetics

Assessment of whole-body protein turnover relies on the measurement of the dilution of


tracer amino acids in plasma or whole blood, i.e. the rate of appearance (the flux) of the tracee
amino acid assuming that the blood pool and tissue free amino acid pools are homogeneously
mixed. Provided that a carbon-labeled indispensable amino acid is infused intravenously at a
constant rate until isotopic equilibrium is attained in the plasma, the amino acid kinetic rates
can be converted to whole-body protein kinetics rates using the average fractional contents of
individual amino acids in body protein (Waterlow and Stephen, 1967). The equation can be
expressed as follows:

Whole-body flux of AA (Q) = Intake of dietary AA + Release of AA from protein breakdown


= Utilization of AA for protein synthesis (nonoxidative dis-
posal) + Oxidation of AA

Whole-body flux of AA (Q) can be estimated by the isotopic enrichment (IE) of plasma
AA at isotopic equilibrium, i.e.

Q = I × [(IEi/IEp) – 1]

where I is the infusion rate of the tracer AA, and IEi and IEp represent the isotopic enrichment
of the labeled AA in infusate and that of plasma AA at plateau, respectively. Both protein syn-
thesis and protein degradation can be solved from the equation. Because protein synthesis,
breakdown, and amino acid oxidation are intracellular events, it is necessary to measure the
isotopic enrichment of the intracellular free amino acid pool rather than its plasma isotopic
enrichment to calculate these kinetics.
There are three technique issues. The first one is how to assess the isotopic enrichment of
the intracellular true precursor (i.e. amino acyl-tRNA) for protein synthesis (see section
6.2.1). The second one is how to assess the rate of oxidation. When a primed constant infu-
sion of 13C-labeled tracer (e.g. 1-13C-leucine or 1-13C-phenylalanine) is employed to estimate
rates of whole-body protein synthesis, one has to determine the rate of oxidation of the tracee
by measuring the appearance rate of 13CO2. However, the labeling position of 13C tracer
affects recovery of the 13CO2. For example, the recovery of the 2-13C label in breath CO2 is
58% relative to the 1-13C label, suggesting that a significant percentage (~42%) is retained in
the body although a majority of the 2-13C label of leucine is recovered in the breath CO2,
presumably by transferring to other compounds via the tricarboxylic acid cycle (Toth et al.,
2001). Not all of the 1-13C liberated from oxidative disposal appears in the breath CO2. Ring-
2H -phenylalanine (Phe) and 1-13C-tyrosine (Tyr) are infused simultaneously to estimate
5
phenylalanine irreversible hydroxylation (Clarke and Bier, 1982). The hydroxylation rate of
Phe into Tyr can be derived from the equation (Short et al., 1999):

QPhe − Tyr = QTyr × IEd4−Tyr/IEd5− Phe

where QTyr is whole-body flux of plasma Tyr that is estimated from 1-13C-Tyr infusion, IEd4-Tyr
and IEd5−Phe are the isotopic enrichments of plasma L-ring-2H4-Tyr and L-ring-2H5-Phe, respec-
tively. This approach has been employed to estimate the in vivo hydroxylation rate of Phe to Tyr
in patients with phenlketonuria (van Spronsen et al., 1998). Other combinations (e.g. 15N-Phe
and ring-2H4-Tyr or ring-2H5-Phe and ring-2H2-Tyr) have also been used to estimate whole-body
hydroxylation (Meek et al., 1998; Short et al., 1999). The advantage of the Phe hydroxylation
model is the rapid assessment of whole-body protein turnover from plasma samples alone with-
out measurement of breath 13CO2 production (Clarke and Bier, 1982; Thompson et al., 1989).
Methodological approaches to metabolism research 449

Fig. 3. A three-compartment model for deriving whole-body protein turnover using 15N,13C-Leu tracer
(adapted from Gowrie et al., 1999). The model describes the kinetics events during an intravenous infusion of
15N,13C-Leu tracer. In brief, infused 15N,13C-Leu tracer enters the plasma free Leu pool (compartment 1) and

further enters the intracellular AA free pool (compartment 2) where it may be irreversibly deaminated (indi-
cated by irreversible loss of 15NH2, k02) or incorporated into the intracellular protein pool (compartment 3,
indicated by the fractional transfer rate from compartment 2 to 3, i.e. k32). The isotopic enrichment of the
intracellular free Leu pool can be diluted by unlabeled tracee Leu from release of protein breakdown (indi-
cated by the fractional transfer rate from compartment 3 to 2, i.e. k23).

The third technique issue is how to model experimental data. Recently, a three-
compartment model has been developed to assess whole-body protein synthesis and break-
down with a 15N,13C-Leu tracer (fig. 3) (Gowrie et al., 1999). This three-compartment model
that represents 15N,13C-Leu tracer kinetics can be described by a set of differential equations.
Fractional rate constants (fractional rates of protein synthesis and breakdown indicated by k32
and k23, respectively) can be solved using the SAAM II program.

6.2. Precursor method for measuring fractional synthesis rate of tissue protein

6.2.1. Constant infusion method

The fractional synthesis rate (FSR) of protein has been evaluated by a direct precursor–
product relationship. The constant infusion method has been used to measure both whole-
body protein turnover and tissue protein synthesis. This method involves the infusion of a
tracer amino acid at a constant rate until steady-state isotopic labeling of the precursor amino
acyl-tRNA pool is reached. Specifically, when a precursor tracer (e.g. a labeled amino acid)
is provided as a primed constant infusion into a system, the isotopic enrichment of a homo-
geneous product pool will increase as a monoexponential function of time (IEt), i.e. IEt =
IEp (1 − e−kt), where IEp is the enrichment of the precursor pool. Therefore,

FSR = k = [(IEt2 − IEt1)/(t2 − t1)]/IEp

The FSR is determined by dividing the initial rate of change in the product isotopic enrich-
ment by the precursor isotopic enrichment at the steady state (Patterson, 1997). For example,
the FSR of human small intestinal mucosal protein is calculated by a primed constant infu-
sion of 1-13C-leucine using this equation, in which IEt2 is the isotopic enrichment of mucosal
protein-bound leucine (from the mucosal biopsy at time 2), IEt1 is the isotopic enrichment of
mucosal protein-bound leucine (from the mucosal biopsy at time 1) or the isotopic enrich-
ment of plasma protein-bound leucine at time 1, and IEp is the isotopic enrichment of the
precursor pool (e.g. tissue-free fluid 13C-leucine or plasma 13C-ketoisocaproate (Bouteloup-
Demange et al., 1998; Charlton et al., 2000). In order to avoid multiple tissue samples, an
450 X. Guan and D. G. Burrin

overlapping (i.e. staggered) infusion of multiple stable amino acid isotopomers has been
developed to measure in vivo FSR when only a single tissue sample can be obtained (Dudley
et al., 1998). Labeled Phe has been also used to assess the FSR of intestinal mucosal and
muscle proteins (Stoll et al., 1997; Biolo et al., 1999). Furthermore, the isotopic enrichment
of plasma VLDL ApoB-100-bound Phe has been used to represent that of the intracellular
free Phe to calculate the FSR of hepatic protein synthesis (Stoll et al., 1997).
The accuracy of this precursor method depends on the measurement of the isotopic enrich-
ment of the intracellular true precursor pool. By definition, the tissue tRNA-bound amino acid
is the immediate precursor used for protein synthesis (Watt et al., 1991). However, it is diffi-
cult to measure the isotopic enrichment of tRNA-bound amino acid, specifically when the
precursor pool is not accessible. There are four alternative solutions to measuring the isotopic
enrichment of the intracellular true precursor pool. The first is to measure isotopic enrichment
of tissue free amino acids. During a constant infusion of labeled Leu, there is a quite close
isotopic equilibrium between muscle-free and tRNA-bound leucine pools (Watt et al., 1991;
Reeds and Davis, 1999). The isotopic enrichment of tissue fluid Leu in human skeletal muscle
has been proved a valid surrogate measurement of the isotopic enrichment for intracellular
leucyl-tRNA (Ljungqvist et al., 1997). The isotopic enrichment of tissue free Leu has also
been used to estimate the FSR of intestinal mucosal protein (Stoll et al., 2000).
The second approach is to measure the isotopic enrichment of a plasma metabolite that is
exclusively derived from the intracellular metabolism of the precursor, e.g. measuring the iso-
topic enrichment of plasma α-ketoisocaproate (KIC) as an index of the isotopic enrichment
of the intracellular free leucine to calculate the FSR of muscle and hepatic proteins (Mansoor
et al., 1997). The KIC is formed intracellularly from leucine and is released, in part, into the
systemic circulation. Thus, the isotopic enrichment of plasma KIC can be used to represent
the isotopic enrichment of the intracellular free leucine pool (Matthews et al., 1982).
However, isotopic enrichments of plasma KIC and leucine have been shown to be consis-
tently higher than those of tissue leucyl-tRNA and tissue fluid leucine (Chinkes et al., 1996a).
Therefore, using the isotopic enrichment of plasma KIC as a surrogate measurement of the iso-
topic enrichment for leucyl-tRNA will underestimate the FSR of muscle protein, whereas the
isotopic enrichment of tissue fluid leucine is a valid surrogate measurement (Watt et al., 1991;
Ljungqvist et al., 1997). In a reversal of this approach, constant infusion of α-[1-13C]KIC is
more accurate than labeled leucine to determine the FSR of muscle protein (Chinkes et al.,
1996a). When labeled KIC is infused, the isotopic enrichment of intramuscular free Leu is the
same level as that of arterial Leu (Chinkes et al., 1996a).
The third approach is to use the isotopic enrichment of newly synthesized protein-bound
amino acid to represent the isotopic enrichment of the true precursor, e.g. using the isotopic
enrichment of very-low-density lipoprotein apolioprotein B (VLDL ApoB)-100-bound
amino acid as an index of the isotopic enrichment of the hepatic amino acid pool (Reeds
et al., 1992). The VLDL ApoB-100 is made in the liver and has a very short half-life in the
circulation. The isotopic enrichment of VLDL ApoB-100-bound amino acid rapidly rises to
the same level as that of the precursor pool in the liver. Because of the heterogeneous com-
position of the hepatic intracellular precursor pool, the isotopic enrichment of VLDL
ApoB-100-bound amino acid may provide a more valid measurement of the isotopic enrich-
ment of the hepatic protein synthetic precursor than the hepatic free amino acid pool does
(Stoll et al., 1997, 1999b).
However, there are discrepancies in the literature. Isotopic enrichments of different pre-
cursors for liver protein synthesis have been compared with that of amino acyl-tRNA using
1-13C-Leu and 15N-Phe as tracers in miniature swine (Ahlman et al., 2001). It is shown in fig. 4
Methodological approaches to metabolism research 451

Fig. 4. Ratios of other precursors to amino acyl-tRNA in porcine liver (data from Ahlman et al., 2001).

that isotopic enrichment ratios of 13C-Leu and 15N-Phe in liver tissue fluid and 13C-KIC in
plasma to those of respective amino acyl-tRNA are close to 1.0, indicating that isotopic
enrichments of tissue fluid amino acid and plasma 13C-KIC are the best predictors of the iso-
topic enrichments of tissue amino acyl-tRNA in the liver and skeletal muscle under different
physiological conditions (Barazzoni et al., 1999; Ahlman et al., 2001). In contrast, isotopic
enrichments of plasma Leu and Phe are substantially higher, whereas that of plasma VLDL
ApoB-100-bound amino acid is lower than that of the respective amino acyl-tRNA (Ahlman
et al., 2001). Consequently, the FSR of the liver protein derived from isotopic enrichments of
plasma 13C-Leu or plasma VLDL ApoB-100-bound amino acid would be underestimated or
overestimated, respectively (Ahlman et al., 2001). Recently, the FSR of slow-turnover protein
has been assessed by orally or intravenously administrating 2H2O to label nonessential amino
acids (Hellerstein et al., 2002; Previs, 2002). This method takes advantage of the fact that
through transamination reactions the α-hydrogen of nonessential amino acids (e.g. alanine
and glutamine) equilibrates rapidly and completely with the 2H of body water. Thus, the FSR
of tissue protein can be estimated by measuring the incorporation of 2H-alanine and/or
2H-glutamine into protein. However, at this time, there has been no demonstration of the

equivalence of the isotopic enrichment of plasma alanine and glutamine and that of their
tissue free pools.
The fourth approach is to derive the isotopic enrichment of the intracellular true precursor
pool from mass isotopomer distribution analysis (see section 9). In conclusion, using the con-
stant infusion method for assessing FSR of tissue protein, the best estimate of the isotopic
enrichment of intracellular true precursor (amino acyl-tRNA) pool seems to be the tissue free
amino acid in muscle (Davis and Reeds, 2001).

6.2.2. Flooding dose method

To avoid the problem in measuring isotopic enrichment of the intracellular true precursor
(amino acyl-tRNA), the flooding dose method has been developed for measuring tissue
protein synthesis. This approach involves giving a bolus injection of labeled amino acid with
452 X. Guan and D. G. Burrin

a large bolus dose of unlabeled amino acid (e.g. 5 to 10 times the endogenous flux of the
tracee) to rapidly create a similar isotopic enrichment in the extra- and intracellular compart-
ments. The FSR of liver protein is determined by flooding dose method with 2H5-Phe using
the equation (Barle et al., 1999):

FSR = IEp × 100/AUC

where IEp is the isotopic enrichment of liver protein-bound Phe at the time of the biopsy and
AUC is the area under the curve for the isotopic enrichment of plasma free Phe versus time.
The flooding dose method assumes rapid equilibration of the isotopic labeling among the
amino acyl-tRNA pool, the tissue free amino acid pool, and the blood free amino acid pool.
This assumption has been recently validated by the fact that ratios of specific radioactivity are
close to 1.0 for the tissue free Phe pool versus the phenylalanyl-tRNA pool in either skeletal
muscle or liver (Davis et al., 1999). Under different nutritional and hormonal conditions, the
isotopic enrichments of the tissue free Phe pool may be considered satisfactory for assessing
the FSR of skeletal muscle and liver proteins when a flooding dose of Phe is administered
(Davis et al., 1999). The short period of measurement with this method is especially valuable,
as it allows the determination of acute changes in tissue protein synthesis within
30 min (Garlick et al., 1994). However, it has been observed that large doses of leucine might
stimulate protein synthesis in muscle tissue (Ballmer et al., 1990). The FSR of muscle pro-
tein by the flooding dose method is higher than that measured by the constant infusion
method when 13C-leucine is used (Garlick et al., 1994; Rennie et al., 1994). To study luminal
versus basolateral modulation of protein metabolism in small intestinal mucosa, a local
(luminal) flooding dose method has been used to determine the fractional rate of protein
synthesis in intestinal mucosa (Adegoke et al., 1999a,b).

6.3. Tracee release method for measuring fractional breakdown


rate of tissue protein

To measure the fractional breakdown rate (FBR) of muscle protein, the tracee release method
has been developed on the basis of the precursor–product principle (Zhang et al., 1996). This
method involves infusing isotope tracer (e.g. ring-2H5-Phe or ring-13C6-Phe) until isotopic
equilibrium is reached. The assessment of the rate of protein breakdown is achieved by meas-
uring isotopic enrichment decay curves of the arterial and tissue free amino acid pools after
the tracer infusion is stopped. Because there is no de novo synthesis of Phe, its appearance in
the tissue free amino acid pool is solely attributed to transport from blood and release by prote-
olysis. At isotopic equilibrium, the isotopic enrichment in the tissue free amino acid pool is
always lower than that in the arterial blood because the former is diluted by intracellular unla-
beled amino acid released from protein breakdown. Once the isotopic infusion is stopped, the
enrichment decay in the tissue free amino acid pool depends on the isotopic enrichment decay in
the arterial blood, which provides tracer and a part of the tracee (i.e. Phe), and on the protein
breakdown, which provides another part of the tracee. The calculation of FBR is based on the
rate at which tracee is released from protein breakdown to dilute the isotopic enrichment of the
tissue free amino acid pool using a modified precursor–product equation (Zhang et al., 1996), i.e.

FBR =
[ IEm (t2 ) − IEm (t1 )] × (Qm / T )
⎡ t2 t2 ⎤
⎢ P ∫ IEa (t )dt − (1 + P) ∫ IEm (t )dt ⎥
⎢⎣ t1 t1 ⎥⎦
Methodological approaches to metabolism research 453

where P = IEm/(IEa − IEm) represents the ratio of fractional tracee from artery versus the frac-
tional tracee from protein breakdown, IEa and IEm being the isotopic enrichments at plateau
in the artery pool and muscle intracellular free pool, respectively. The IEm(t2) − IEm(t1) is the
change of the isotopic enrichment in muscle intracellular free pool from time 1 (t1) to time 2
(t2) after stopping the isotopic infusion,
t2 t2
∫ IEa (t )dt and ∫ IEm (t )dt
t1 t1

are areas under the decay curves of the isotopic enrichments in arterial and muscle intracel-
lular free pools, respectively, from t1 to t2. Qm/T is the ratio of the intracellular free tracee
mass versus protein-bound tracee mass in the muscle. The FBR assessed by the tracee release
method is in agreement with that derived from the arterio-venous tracer balance method (see
section 6.4). The tracee release method is the complement of the tracer incorporation method.
These two methods can be combined to measure both muscle protein synthesis and break-
down in one infusion study and can be applied to other tissues (e.g. skin) if a few biopsies can
be obtained (Zhang et al., 1996; Volpi et al., 2000). The tracee release method has been
recently improved by using a pulse tracer injection (Zhang et al., 2002). This new approach
does not require an isotopic steady state, and it can be completed within an hour and using
one or two muscle biopsies.

6.4. Arterio-venous tracer amino acid balance method for measuring tissue
amino acid transport, protein synthesis, and breakdown

Since phenylalanine is neither synthesized nor degraded by muscle tissue, the measured
removal of tracer and dilution of its isotopic enrichment across the hindlimb can be used to
estimate rates of phenylalanine incorporation into and release from tissue protein. This meas-
urement, coupled with an estimate of tissue blood flow, can provide a readily nondestructive
method for estimation of protein turnover in specific muscle beds in vivo. Measurements can
be made repeatedly over time in a single experiment, allowing the study of acute regulation
of protein turnover (Barrett et al., 1987). This conventional arterio-venous tracer amino acid
balance approach has been improved by measuring the isotopic enrichment of the intracellu-
lar free amino acid pool using muscle biopsy to calculate the relative proportions of
intracellular amino acid derived directly from the blood (labeled) or from tissue protein
breakdown (unlabeled) (Biolo et al., 1995a). Thus, tissue (e.g. muscle) protein synthesis and
breakdown and transmembrane transport of the amino acid can be determined simultane-
ously. The arterio-venous tracer amino acid balance method (the A-V method) can be
described in a three-compartment model. This model is based on an anatomic compartmen-
tation of an indispensable amino acid (e.g. Leu or Phe) into three compartments: the arterial,
the intracellular free, and the venous compartments (fig. 5).
In this compartmental model, no interstitial free pool is assumed, i.e. isotopic tracers are
assumed to enter their intracellular free compartments at the arterial values and leave from
their intracellular free compartments at the venous values. It is also assumed that there is no
recycling of isotopic tracers released from protein breakdown into the intracellular free
compartment. Isotopic enrichments of the intracellular free compartment (in the tissue fluid)
may also be represented by measurement of the isotopic enrichments of other compounds. For
example, the isotopic enrichment of the local venous plasma 13C-KIC is used to represent that
of the intracellular free 13C-Leu. The isotopic enrichments of liver-synthesized protein VLDL
454 X. Guan and D. G. Burrin

Fig. 5. A three-compartment model of amino acid kinetics across the porcine mammary gland during
lactation (adapted from Guan et al., 2002). Free amino acid compartments in artery (A), main mammary vein (V),
and mammary gland (MG) are connected by arrows indicating unidirectional fluxes of free amino acids
between each compartment. Amino acids enter the MG via the mammary artery (Fa,o) and leave the MG via
the main mammary vein (Fo,v). Other fluxes are designated as follows: Fv,a, direct flow of amino acids from
artery to vein without entering the intracellular pool (by the arterial shunt); Fmg,a and Fv,mg, inward and
outward transmembrane transport of amino acids from artery to the MG and from the MG to vein, respectively;
Fmg,o, the rate of intracellular amino acid appearance from endogenous sources (i.e. release from protein
breakdown (PB) and de novo synthesis (DS), if any); and Fo,mg, the rate of the intracellular amino’ acids
disappearance (i.e. the rate of utilization of intracellular amino acids for protein synthesis (PS), oxidation
(OX), and other metabolic fates (OM), if any).

ApoB-100-bound or mammary-synthesized casein-bound amino acids are also used to repre-


sent those of the intracellular precursor for hepatic or mammary tissue protein synthesis (Reeds
et al., 1992; Bequette et al., 2000; Guan et al., 2002). Below, we use this three-compartment
model to derive the rates of protein synthesis and breakdown and the transmembrane transport
of amino acids across an organ of interest.

6.4.1. Protein breakdown

Protein breakdown (PB) can be derived from appearance rate (Ra). Since Ra = PB + arterial
influx, thus PB = Ra − arterial influx, i.e. PB = (Ca × IEa × BF)/IEi – Ca × BF, where IEa and IEi
are isotopic enrichments of Phe in artery and tissue fluid (intracellular free pool), respectively,
Ca is the arterial concentration of Phe, and BF is blood flow rate across the organ. Therefore,

Protein breakdown (PB) = Ca × BF × [(IEa/IEi) − 1] (1a)

Assuming that the isotopic enrichment of Phe in vein (IEv) can represent IEi, therefore,

Protein breakdown (PB) = Ca × BF × [(IEa/IEv) − 1] (1b)

In fact, eq. (1) can also be derived from irreversible loss of tracee (IL). Since IL − PB = Net
mass balance (NB), thus, PB = IL − NB, i.e. PB = Tracer uptake/IEi − NB, therefore, PB =
[(Ca × IEa – Cv × IEv) × BF]/IEi − (Ca − Cv) × BF; when simplified, its equation is identical
to eq. (1a) . It is noted that IEv usually overestimates IEi, thus protein breakdown from eq. (1b)
may be underestimated. Alternatively, PB can be derived from unidirectional influx (UI).
Since NB = UI − PB, thus PB = UI − NB, where UI = Tracer fractional extraction
Methodological approaches to metabolism research 455

rate × Arterial influx = [(Ca × IEa – Cv × IEv)/(Ca × IEa)] × (Ca × BF), and NB = (Ca − Cv) ×
BF, through rearrangement of the equation, therefore,

Protein breakdown (PB) = Cv × BF × [1 − (IEv/IEa)] (2)

6.4.2. Protein synthesis

Protein synthesis (PS) can be derived from irreversible loss of tracee (IL). Since IL = PS +
Oxidation (or hydroxylation of Phe to Tyr in liver or kidney), thus

Protein synthesis (PS) = Irreversible loss − Oxidation (or Hydroxylation) (3)

where irreversible loss (IL) = Tracer uptake/Ei, i.e.

Irreversible loss (IL) = (Ca × IEa – Cv × IEv) × BF/IEi (4)

In eq. (4), IEv may be used to replace IEi assuming that the isotopic enrichment of Phe in vein is
proximate to that of its intracellular free pool. If 1-13C-Leu is infused, the isotopic enrichment of
venous plasma 13C-KIC may represent IEi. In fact, IL can be derived from the difference between
net mass balance and protein breakdown. Since NB = IL − PB, i.e. IL = NB + PB, thus

Irreversible loss (IL) = (Ca − Cv) × BF + PB

Oxidation rate across the organ can be determined by labeled CO2 production. If 1-13C-Leu
is infused, oxidation of tracee = 13CO2 production/IEi, i.e.

Oxidation = (CCO2v × IECO2v− CCO2a × IECO2a) × BF/IEi

The isotopic enrichment of the intracellular free Leu may be indicated by the isotopic enrich-
ment of venous plasma KIC. If different 13C labeling position or multiple 13C labeling of
leucine is infused, oxidation estimated from eq. (5) should be adjusted by a correction factor.
Instead of estimating oxidation, hydroxylation of phenylalanine to tyrosine across the organ

Table 1
Applications of the arterio-venous stable isotopic tracer amino acid balance method

Regional bed Inflow Outflow Blood flow Reference

Kidney Femoral Renal vein The Fick method Moller et al. (2000)
artery (paraaminohippurate) Tessari et al. (1996)
Liver Hepatic artery Hepatic vein Doppler flow probe Tessari et al. (1996;
Portal vein The Fick method Halseth et al. (1997)
(indocyanine green)
Mammary Carotid artery Mammary The Fick method Guan et al. (2002)
gland vein (internal Phe + Tyr)
Placenta Maternal Umbilical The Fick method Paolini et al. (2001)
femoral artery vein (3H2O dilution)
Portal-drained Carotid artery Portal vein Transit-time Guan et al. (2003)
viscera ultrasound flow meter
Skeletal muscle Femoral Femoral The Fick method Meek et al. (1998)
artery vein (indocyanine green) Tessari et al. (1996)
Splanchnic bed Femoral Hepatic vein The Fick method Meek et al. (1998)
artery (indocyanine green) Moller et al. (2000)
Tessari et al. (1996)
456 X. Guan and D. G. Burrin

can be determined (Moller et al., 2000). Note that if this hydroxylation is used to correct irre-
versible loss to obtain protein synthesis on the basis of eq. (3), oxidation should not be
subtracted any more because oxidation is part of hydroxylation.
This A-V method has been widely used to estimate amino acid kinetics across a particular
organ (table 1). For example, in the fasting state, healthy human skeletal muscle is in a catabolic
state to provide amino acids for protein synthesis required by the splanchnic bed. Net amino
acid balance between splanchnic and skeletal muscle beds is achieved through differential reg-
ulations of protein metabolism in these tissues by insulin (Meek et al., 1998). In the fasting state
of the dog, the splanchnic bed contributes about 40% to the whole-body protein breakdown and
the gut and liver each contribute about 50% to the splanchnic bed (Halseth et al., 1997),
indicating the equal significance of gut and hepatic proteolysis to whole-body proteolysis.
It is important to accurately measure regional blood flow rate when using this A-V method.
Three methods are used for measuring regional blood flow rate: fluorescent microsphere,
external unmetabolized marker (e.g. indocyanine green and paraaminohippurate), and ultra-
sonic flow probe. If there are multiple entrance or exit vessels, blood flow rate may be
appropriately measured by the Fick method based on the conservation of mass. For example,
blood flow rates across the splanchnic bed and mammary gland have been estimated by the
Fick method (Meek et al., 1998; Guan et al., 2002). It has been shown that mammary blood
flow rates estimated by the ultrasonic method are comparable to those estimated by the Fick
method (Trottier et al., 1997; Renaudeau et al., 2002).

6.4.3. Transmembrane transport of amino acids

The same three-compartment model has been employed to assess amino acid inward and out-
ward transport (Biolo et al., 1992, 1995a). Using the porcine mammary gland as an example, the
calculations based on references are (Reeds et al., 1992; Bequette et al., 2000; Guan et al., 2002):

Fa,o = Ca × BF

Fo,v = Cv × BF

Net mass balance = (Ca − Cv) ⋅ BF

Based on the net mass balance of AA across the mammary gland (MG),

Fa,o = Fmg,a + Fv,a

Fo,v = Fv,mg + Fv,a

Thus,

(Ca − Cv) × BF = Fmg,a − Fv,mg

Based on the tracer balance of AA across the MG,

(Ca × IEa − Cv × IEv) × BF = Fmg,a × IEa − Fv,mg × IEi

Therefore,

Fmg,a ={[(IEi − IEv)/(IEa − IEi)] × Cv + Ca} × BF (6)

Fv,mg ={[(IEi − IEv)/(IEa − IEi)] × Cv + Cv} × BF (7)

The only source of tracer appearing in the mammary intracellular free AA compartment is
transported inward from plasma. Thus, the isotopic enrichment of tracer AA in the intracellular
Methodological approaches to metabolism research 457

free AA is diluted by endogenous sources (Fmg,o) (e.g. protein breakdown and de novo syn-
thesis, if any). Therefore,

Ra × IEi = Fmg,a × IEa ,

i.e. Ra = (Fmg,a × IEa)/IEi (8)

where Ra is the sum of inward transmembrane transport (Fmg,a) and the appearance rate of
intracellular free AA from the endogenous sources (Fmg,o):

Ra = Fmg,a + Fmg,o

Thus,
Fmg,o = Fmg,a × (IEa/IEi − 1) (9)

At steady state, the total fluxes into the mammary intracellular free AA compartment are
equal to the total fluxes out of this compartment, i.e.

Fmg,a + Fmg,o = Fv,mg + Fo,mg

Thus,

Fo,mg = Fmg,o + NB (10)

The disappearance rate (Fo,mg, i.e. utilization rate of the intracellular free AA) of intracellu-
lar free AA could also be directly calculated as the tracer balance divided by the precursor
enrichment (IEi):

Fo,mg = (Ca × IEa − Cv × IEv) × BF/IEi (11)

It has been shown using this three-compartment model that increased net protein synthesis
in human muscle (Volpi et al., 1998) and the porcine mammary gland (Guan et al., 2002) by
intake of dietary indispensable amino acids is attributed to increased inward transmembrane
transport of these amino acids into the respective organs, and that the net flow of amino acids
from muscle to the gut in the fasting state is attributed to differences in their transmembrane
transport rates (Biolo et al., 1995b).

6.4.4. In vivo nitric oxide synthase activity

This arterio-venous tracer amino acid balance method can be used to assess in vivo nitric
oxide synthase (NOS) activity across an organ. Guanidine-15N2-arginine is converted to
ureido-15N-citrulline and 15NO through NOS reaction (Palmer et al., 1988), and used to quan-
tify in vivo NOS activity across organs (Bruins et al., 2002). If guanidine-15N2-arginine and
ureido-13C-5,5,2H2-citrulline are infused, unidirectional flux of guanidine-15N2-arginine to
ureido-15N-citrulline (QArg→Cit) can be estimated on the basis of ureido-15N-citrulline tracer
balance (see fig. 6), i.e. based on ureido-15N-citrulline tracer balance:

QIN + QNOS = QOUT + QM

where

QIN = Ca × IE15N-Cit,a × BF

QNOS = QArg→Cit × IE15N-Arg,i

QOUT = Cv × IE15N-Cit,v × BF
458 X. Guan and D. G. Burrin

Fig. 6. A compartmental model of arginine and citrulline kinetics across the portal-drained viscera (PDV).
QIN, QOUT, QM, and QArg→Cit represent unidirectional fluxes of ureido-15N-citrulline tracer entered from the
artery, exited to the vein, metabolized in the organ, and de novo synthesized from arginine through the NOS
reaction, respectively. Dotted lines and solid lines indicate fluxes of tracer and tracee, respectively.

QM = Irreversible loss × IE15N-Cit,i


= [(Ca × IEd-Cit,a × BF − Cv × IEd-Cit,v × BF)/IEd-Cit,i] × IE15N-Cit,i

where BF is the blood flow across the organ; Ca and Cv are concentrations of free citrulline in
artery and vein, respectively; IE15N-Cit,a (IEd-Cit,a), IE15N-Cit,i (IEd-Cit,i), and IE15N-Cit,v (IEd-Cit,v) are
isotopic enrichments of ureido-15N-citrulline (ureido-13C-5,5,2H2-citrulline) in artery, the intra-
cellular compartment, and vein, respectively; and IE15N-Arg,i and IE15N-Arg,v are isotopic
enrichments of guanidine-15N2-arginine in artery and vein, respectively. Assuming that
isotopic enrichment of the tracer in vein is proximate to that in the intracellular compartment,
i.e., IE15N-Arg,i = k1 × IE15N-Arg,v; IE 5N-Cit,i = k2 × IE15N-Cit,v; IE d-Cit,i = k3 × IE d-Cit,v; and k1 ≈ k2 ≈
k3 ≈ 1, therefore,
QArg→Cit = [(IE d-Cit,a/IEd-Cit,v) × (IE15N-Cit,v/IE15N-Arg,v) − (IE15N-Cit,a /IE15N-Arg,v)] × Ca × BF (2)
This unidirectional flux (QArg→Cit) indicates in vivo NOS activity across the organ (e.g. PDV).
The principle of this method has been also applicable to assessment of the local conversion (e.g.
hydroxylation of phenylalanine to tyrosine across the liver or kidney) (Moller et al., 2000).

6.4.5. First-pass utilization

In amino acid tracer kinetic studies, ingested amino acid is taken up during its initial transit
through the splanchnic bed and thus not all absorbed amino acids enter the systemic com-
partment. The amount of enterally delivered tracer (or tracee) sequestered by the splanchnic
bed can be estimated by simultaneously administrating a labeled tracer intravenously (iv) and
intraduodenally (id) (or intragastrically, ig) (Matthews et al., 1993a,b). To assess protein
metabolism in the splanchnic bed, the infusion of tracer amino acid into the gastrointestinal
tract should ideally avoid gastric emptying in the postabsorptive state. An intraduodenal
administration of AA tracers is recommended to obtain plasma isotopic enrichments at steady
state (Crenn et al., 2000). This method allows measurement of splanchnic extraction and
Methodological approaches to metabolism research 459

indirect assessment of splanchnic protein metabolism under fasting and feeding conditions
(Crenn et al., 2000). The splanchnic extraction coefficient of leucine is defined (Stoll et al.,
1997, 1999b; Basile-Filho et al., 1998, 1999b) as:
Splanchnic extraction = [Qid − Qiv]/Qid = 1 − Qiv/Qid
in which Qid or Qiv (whole-body flux) = Rate of tracer infusion · [(IEi/IEp) − 1]. If whole-body
flux is not corrected by the amount of tracer infused, then splanchnic extraction is simplified
as (Crenn et al., 2000):
Splanchnic extraction = 1 − [(IEp,id/Iid) / (IEp,iv/Iiv)]
where IEp,id and IEp,iv are isotopic enrichments (tracer-to-tracee ratio, TTR) of plasma amino
acids at plateau after a stable isotopic tracer (e.g. 2H3-Leu) is infused intraduodenally and
intravenously, respectively. Iid and Iiv are infusion rates of the tracer via intraduodenal and
intravenous routes, respectively. Fractional splanchnic oxidation (of whole-body oxidation)
and fractional splanchnic hydroxylation of Phe to Tyr (of whole-body hydroxylation) have
been estimated using this method (Basile-Filho et al., 1998). Because the rate of intragastri-
cally infused tracer (13C-Phe) appearing in the nonsplanchnic pool is the product of the tracer
infusion rate (Iig) and nonsplanchnic extraction (Fnsp,extraction = 1 − splanchnic extraction),
which is handled in first pass in a manner similar to that for intravenously infused tracer, then
the fractional nonsplanchnic oxidation (Fnsp,oxidation, of whole-body total oxidation) can be
calculated as follows (Basile-Filho et al., 1998):
Fnsp,oxidation = Fnsp,extraction × Riv,oxidation/Rig,oxidation
and Fsp, oxidation = 1 − Fnsp,oxidation

where Fsp,oxidation is the fractional splanchnic oxidation (of whole-body oxidation), and Riv,oxidation
and Rig,oxidation are whole-body oxidation rates calculated by intravenously and intragastrically
infused tracer, respectively. Similarly, fractional splanchnic hydroxylation (of whole-body
hydroxylation) can be obtained with an intravenous infusion of ring-2H4-Tyr (see details in
Basile-Filho et al., 1998).
We have combined this method and the portal tracer amino acid balance to further assess
amino acid metabolism in the gut and liver (Stoll et al., 1997; van Goudoever et al., 2000).
Hepatic extraction is defined by the difference between splanchnic extraction (derived from
above) and portal extraction (derived from the portal tracer amino acid balance). We have
found that splanchnic extraction of Phe is attributed to fractional extraction of the gut and the
liver by 75% and 25%, respectively (Stoll et al., 1997), and one-third of the dietary amino
acids is metabolized in the gut (Stoll et al., 1998, 1999a; van Goudoever et al., 2000). It is
important to assess amino acid metabolism and protein turnover in the splanchnic bed
(including the gut and liver) in order to predict the post-splanchnic availability of absorbed
dietary amino acids and to understand tissue protein metabolism under different nutritional,
physiological, and pathological conditions.

7. STABLE ISOTOPIC TRACER TECHNIQUES FOR STUDYING


LIPID METABOLISM
Regulation of lipid metabolism is not only related to growth and fattening of animals, but also
to the development of cardiovascular disease, insulin resistance, diabetes, and obesity in
humans. The measurement of dynamic fluxes of lipids (biosynthesis, oxidation, and lipoly-
sis) poses difficult challenges. Two fundamental advances have recently been made for
460 X. Guan and D. G. Burrin

measuring lipid biosynthesis, namely deuterium-labeled water incorporation method and


mass isotopomer distribution analysis (MIDA). These techniques have resolved the central
methodological problem in measuring the isotopic enrichment of the intracellular true
precursor pool. In the 2H2O incorporation method, rates of deuterium incorporation into fatty
acids and cholesterol are used to assess de novo lipogenesis and cholesterol synthesis, respec-
tively (Diraison et al., 1997; Guo et al., 2000; McDevitt et al., 2001; Bassilian et al., 2002),
in which 2H2O (tracer) equilibrates well among the intracellular precursors NADPH
and water (Di Buono et al., 2000). Besides, the MIDA is also used to assess the FSR of the
VLDL by measuring incorporation of repeating subunits of acetyl-CoA into the newly
synthesized triglyceride (TG) after a constant infusion of [1-13C]acetate (Chinkes et al.,
1996b). Labeled glycerol is also used in the study of lipid metabolism (Siler et al., 1998;
Lemieux et al., 1999).

7.1. Whole-body lipolysis

Whole-body lipolysis can be determined by the dilution method (e.g. at the infusion of
[1,2,3,4-13C4]palmitate and [2H5]glycerol or [2-13C]glycerol) (Horowitz et al., 1999; Siler et al.,
1999; Wang et al., 2000; Bergeron et al., 2001). With intravenous infusion of [2H5]glycerol,
appearance rate (Ra) of plasma glycerol represents the rate of glycerol released into plasma
from hormone-sensitive lipase hydrolysis of adipose tissue and intramuscular TG and the rate of
glycerol released into plasma during lipoprotein lipase hydrolysis of VLDL-TG (Mittendorfer
et al., 2001). However, it does not include the rate of glycerol released during lipolysis of
intra-abdominal adipose tissue TG, which is cleared by the liver (Mittendorfer et al., 2001).
Moreover, lipolysis is underestimated by the extent to which glycerol released by lipolysis
does not enter the systemic circulation, as occurs when lipolysis takes places in the nonhepatic
tissue of the splanchnic bed (Landau, 1999a). Thus, the glycerol Ra is used to calculate the
lower limit for whole-body lipolysis (Aarsland et al., 1996). The rate of appearance of fatty
acids (FA) in plasma (Ra) is determined by the equation:

Ra = I × [(IEi/IEp) − 1]

where I is the infusion rate of fatty acid tracer, and IEi and IEp are isotopic enrichments of the
fatty acid in the infusate and in the plasma at plateau. However, FA may be re-esterified to
TG in most tissues.

7.2. Fatty acid kinetics

The constant infusion of U-13C-labeled fatty acids is used to determine the effects of hyper-
glycemia–hyperinsulinemia on whole-body, splanchnic, and leg fatty acid metabolism in
humans (Sidossis et al., 1999). It has been demonstrated that an increase in glucose avail-
ability inhibits fatty acid oxidation across the leg and the splanchnic region under the constant
availability of fatty acids (Sidossis et al., 1998, 1999). The fatty acid kinetic parameters for
the leg and the splanchnic region are derived in the same manner as in section 6.4. In brief,
Net rate of NEFA uptake or release = (Ca − Chv) × hepatic (or leg) plasma flow

where Ca, Chv, and Cfv are arterial, hepatic venous, and femoral venous concentrations of non-
esterified fatty acids (NEFA), respectively.

Fractional extraction of labeled NEFA = [(IEa × Ca − IEhv × Chv)]/(IEa × Ca);


Methodological approaches to metabolism research 461

or = [(IEa × Ca − IEfv × Cfv)]/(IEa × Ca)

where IEa, IEhv, and IEfv are the isotopic enrichments of NEFA in the artery, hepatic vein, and
femoral vein, respectively.

Absolute rate of uptake of NEFA = Fractional extraction × Ca


× Hepatic (or femoral) plasma flow.

Uptake of NEFA that is released as CO2 (%) = Regional 13CO2 production/


regional uptake of labeled NEFA
= (IECO2,hv × CCO2,hv − IECO2,a × CCO2,a)/
(IEa × Ca − IEhv × Chv)

where IECO2,a and IECO2,hv are the isotopic enrichments of CO2 in the artery and hepatic vein,
respectively, and CCO2,a and CCO2,hv are the concentrations of CO2 in the artery and hepatic
vein, respectively.

Absolute rate of oxidation of NEFA = Absolute rate of uptake of NEFA


× % of NEFA uptake that is released as CO2/
the acetate correction factor.

The acetate correction factor accounts for label fixation that might occur at any step between
the entrance of labeled acetyl-CoA into the tricarboxylic acid cycle until the recovery of label
CO2 in breath (Sidossis et al., 1995a). Because label fixation occurs not only via the bicar-
bonate pool, but also via isotopic exchange reactions in the tricarboxylic acid cycle (Sidossis
et al., 1995b), bicarbonate cannot fully correct the label fixation.

7.3. Muscle triglyceride synthesis

On the basis of the precursor–product relationship and the assumption that the intramuscular
NEFA are the synthetic precursors during the infusion of [U-13C]palmitate (Guo and Jensen,
1998):

Fractional synthesis rate (FSR) of intramuscular TG = (IEt2,TG-palmitate– IEt1,TG-palmitate)/


[Averaged IENEFA-palmitate × Time]

where the numerator is the increment in 13C enrichment of muscle TG palmitate during
a 2–4 h interval, and the denominator is the average 13C enrichment of intramuscular non-
esterified palmitate over the same time interval. This measurement is across a particular
muscle bed.

7.4. Hepatic de novo lipogenesis

The rate at which de novo synthesized palmitate is secreted as VLDL-TG is assessed with a
constant infusion of [1,2-13C]acetate using the MIDA. To calculate the fractional synthesis
rate of VLDL-bound palmitate (FSR), the following formula is used (Aarsland et al., 1996):

FSR = [(IE(t2) – IE(t1))/(t2 − t1)]/[8p(1−p)7]

where t2 and t1 are the times when samples are taken, IE(t) is the doubly labeled enrichment at
time t, and p is the MIDA-derived enrichment of the intrahepatic precursor pool (hepatic
acetyl-CoA) for fatty acid synthesis. Here, the factor of 8 accounts for the fact that it requires
462 X. Guan and D. G. Burrin

eight acetate molecules to form one palmitate molecule (i.e. the principal product of
mammalian de novo fatty acid synthesis). The factor of (1−p)7 accounts for the probability
that seven unlabeled acetate molecules will be incorporated into a palmitate molecule
(Chinkes et al., 1996b). The FSR is defined as the fraction of plasma VLDL-bound palmitate
pool, per unit of time, which is newly synthesized. The absolute synthesis rate is then calcu-
lated by multiplying the FSR and the pool size of VLDL-bound palmitate.

8. MASS ISOTOPOMER DISTRIBUTION ANALYSIS


Mass isotopomer distribution analysis (MIDA) is a new technique for quantifying synthesis
rates of polymeric biomolecules from 15N-, 13C-, or 2H-labeled monomeric units in the pres-
ence of unlabeled polymer. Mass isotopomer distribution is analyzed according to a
combinatorial probability model. The isotopomers of a given type of molecule are the vari-
ous combinations of positions of labeled atoms. For example, when the 12C isotope can be
replaced by 13C independently at each position in glucose, there are 64 (26 = 64) different iso-
topomers. The MIDA allows the isotopic enrichment of the monomeric precursor to be
derived indirectly from the isotopic enrichment of the polymer (product). This derived pre-
cursor enrichment presumably represents the steady-state enrichment of the precursor. The
MIDA has been used to measure fractional rates of cholesterol biosynthesis (Lindenthal et al.,
2002), gluconeogenesis (Trimmer et al., 2002), lipogenesis, and protein synthesis.
Monomer subunits are randomly selected from the precursor pool and incorporated into a
polymer (the product). Theoretical distribution of newly formed product molecules can be pre-
dicted by binomial or multinomial expansion. The probabilities of incorporating a given number
of labeled precursors into the product are determined by the isotopic enrichment of the precursor
pool on the basis of a multinomial distribution (Hellerstein and Neese, 1999):

z z!
d ( z, σ , p ) = U (1 − p)( z −σ ) Pσ
σ = 0 ( z − σ )!σ !

where σ is the number of labeled subunits present in the variable moiety of the polymer, z is
the maximum number of monomer subunits that can be labeled in the variable moiety of the
polymer, and p is the fraction of ΔA∞x /ΔA∞y isotopically labeled subunits in the subunit precursor
pool. The value of p is calculated from the best-fit polynomial regression equation of p against
the ratio of in an appropriate reference table (Hellerstein and Neese, 1999). Here, ΔA∞x and ΔA∞y
are defined as the change in fractional abundance (i.e. excess mass isotopomer abundance) in
the newly synthesized or isotopically perturbed polymers only (e.g. ratio of doubly to singly
labeled product, or triply to doubly labeled product). The precursor enrichment (p) is deter-
mined from the measured ratio of ΔA∞x /ΔA∞y using this equation. Based on the
precursor–product relationship, the fractional synthesis (f, the proportion of newly synthesized
molecules present in the mixture) can be calculated using ΔA∞x at the value of p (derived from
the ratio of ΔA∞x /ΔA∞y ), i.e.

f = ΔAx (mixture)/ΔA∞x

where ΔAx (mixture) is the change of the fractional abundance of a mass isotopomer Mx in the
mixture (measured), and ΔA∞x is the enrichment at plateau, i.e. the precursor enrichment in a
one-source biosynthetic system (calculated from the regression equation of ΔA∞x against p, rep-
resenting the asymptotic value of ΔA∞x ). When all the ions in the mass isotopomer spectrum are
not monitored, a correction equation is used for calculating f (Papageorgopoulos et al., 1999).
Methodological approaches to metabolism research 463

Therefore, the fractional synthetic rate constant (ks) is calculated as follows (Papageorgopoulos
et al., 1999):

ks = −In(1 − f)/t

8.1. Protein synthesis

It is difficult to assess protein synthesis using the conventional precursor–product method


because it not easy to accurately measure the isotopic enrichment of the precursor pool (intra-
cellular amino acyl-tRNA). This difficulty has been conquered with the MIDA. Using the
MIDA to assess protein synthesis has become technically feasible and practical in vivo using
proteolytically derived peptides (Papageorgopoulos et al., 1999). To obtain mass isotopomer
distribution, a small peptide that contains repeats of a selected amino acid is generated from a
whole molecule of protein. The kinetics of the peptide component presumably represents the
kinetics of the intact protein. For example, [5,5,5-2H3]leucine is intravenously infused into rats,
and then a leucine-rich peptide is isolated and purified from trypsin-digested rat serum albumin.
Theoretic abundances and excess abundances of mass isotopomers are calculated and measured.
Biosynthetic rates of rat serum albumin are estimated by the MIDA, which are similar to previ-
ously published values (Papageorgopoulos et al., 1999). Based on the exchange of 2H2O with
α-hydrogen of non-essential amino acids (e.g. alanine and glutamine), the MIDA can be used
for measurement of synthesis rates of slow-turnover proteins (Hellerstein et al., 2002).

8.2. Lipogenesis

If [1-13C]acetate is infused in vivo, VLDL-bound palmitate enrichment can be measured by


the tracer-to-tracee ratio (TTR). The precursor (acetyl-CoA) enrichment p is derived from the
MIDA (Chinkes et al., 1996b):

p = [2 × TTR(M + 2)/TTR (M + 1)]/[(n − 1) + 2 × TTR(M + 2)/TTR(M + 1)]

This precursor enrichment is expressed in terms of acetate units, and is converted to units of
single labeled palmitate (IEp) using the binomial equation:

IEp = np(1−p)n−1

The FSR of VLDL-palmitate is calculated on the basis of the precursor-product relationship


as follows:

FSR = [(IE(t2) − IE(t1))/(t2 − t1)]/[np (1 − p)n−1]

where IE(t) is the singly labeled product enrichment at time t, i.e. singly labeled VLDL-palmitate
(M + 2) enrichment. If doubly labeled acetate ([1,2-13C]acetate) is infused rather than singly
labeled acetate, palmitate will appear at the peaks M + 2 and M + 4 rather than M + 1 and M + 2.
In the calculation of the precursor enrichment, TTR(M + 4)/TTR(M + 2) is used in place of
TTR(M + 2)/TTR(M + 1). Recently, 2H2O has been used to label the glycerol moiety of triglyc-
eride to simultaneously measure in vivo TG synthesis and de novo lipogenesis in adipose tissue
(Antelo et al., 2002; Turner et al., 2002).

8.3. Gluconeogenesis (GNG)

The rate of glucose production is the sum of rates of glycogenolysis and gluconeogenesis.
The rate of glycogenolysis is the rate at which glucose is formed from glycogen, which can
464 X. Guan and D. G. Burrin

be determined by the decline in liver glycogen content measured by 13C nuclear magnetic res-
onance spectroscopy (Rothman et al., 1991). The rate of gluconeogenesis is the rate of
glucose synthesis via glucose-6-phosphate from gluconeogenic precursors (e.g. lactate, ala-
nine, pyruvate, and glycerol). Gluconeogenesis can be determined directly by two
approaches. The first one is to assess the fractional contribution of gluconeogenesis by meas-
uring the ratio of the 2H enrichment of the hydrogen bound to C-5 and to that of C-2 of blood
glucose at steady state after oral intake of 2H2O (Landau et al., 1996; Chandramouli et al., 1997;
Petersen et al., 1999). That is because a hydrogen atom from body water is bound to C-5 of
every molecule of glucose formed via gluconeogenesis and none via glycogenolysis, while a
hydrogen atom from body water is added at C-2 of glucose formed via both gluconeogenesis
and glycogenolysis; the ratio of enrichment at C-5 to that at C-2 also provides a measure
of that fraction (Chandramouli et al., 1997). The rate of gluconeogenesis is calculated by
multiplying that ratio by the rate of glucose production, i.e. the rate of appearance of glucose.
Gluconeogenesis determined by the ratio of the 2H enrichments will be overestimated by the
degree of cycling between glucose-6-phosphate and triose phosphate, and/or loss of label via
transaldolase exchange reactions that are part of the pentose cycle (Ackermans et al., 2001),
for the contribution of the cycling between glucose-6-phosphate and triose phosphate results
in an increase in the labeling of C-5 and, thus, in an overestimation of gluconeogenesis, i.e.
the conversion of glycogen to triose phosphates (then used for glucose synthesis) is included
in the estimate of the contribution of gluconeogenesis rather than glycogenolysis
(Chandramouli et al., 1997).
The second approach is to assess the fractional contribution of gluconeogenesis using the
MIDA. Glucose can be considered as a dimer made of two triose subunits. The MIDA of glu-
cose labeled from [2-13C]glycerol, [U-13C3]glycerol, [3-13C]lactate, or [U-13C3]lactate can be
used for estimating the contribution of gluconeogenesis to glucose production (Neese et al.,
1995). The MIDA of glucose is more precise with uniformly labeled than singly labeled 13C
substrates (Previs et al., 1995). In the latter case, ratios of glucose molecules labeled with two
13C atoms (M ) versus with one 13C atom (M ) are very sensitive to a small error in the fairly
2 1
high background correction at M2. Moreover, the contribution of gluconeogenesis to glucose
production is artifactually underestimated by loss of [2-13C]glycerol carbon via the pentose
cycle when [2-13C]glycerol is infused (Previs et al., 1995; Kurland et al., 2000). It is also pos-
sible that a proportion of glucose is formed from glycerol and from amino acids not converted
to glucose via pyruvate (Landau, 1999c). Thus, [U-13C3]lactate appears to be a suitable tracer
for the MIDA of gluconeogenesis in vivo (Previs et al., 1995), especially for tracing low or
moderate rates of gluconeogenesis (Previs et al., 1998).
The MIDA of plasma glucose and lactate can be carried out during an infusion of [U-13C6]glu-
cose. During an infusion of [U-13C6]glucose (M6 glucose), glycolysis leads to the production
of labeled lactate (m3 lactate). When 13C carbon atoms are recycled in gluconeogenesis, glucose
molecules with one, two, or three 13C substitutions (M1, M2, and M3 glucose) are produced. The
appearance of mass isotopomers M1, M2, and M3 of glucose provides a measurement of the
rate of gluconeogenesis. Because the chance of two labeled triose phosphates combining to
form glucose is negligible, M6 glucose behaves as a nonrecyclable tracer, and the steady-state
enrichment of M6 glucose in plasma allows the determination of the hepatic glucose produc-
tion rate. Thus the infusion of [U-13C6]glucose has the advantage of being able to estimate
simultaneously hepatic glucose output and fractional gluconeogenesis from the MIDA of
plasma glucose and lactate and has been used to estimate gluconeogenesis by Tayek and Katz
(Tayek and Katz, 1996, 1997; Katz and Tayek, 1999). However, different equations have been
used to calculate the contribution of gluconeogenesis to glucose production.
Methodological approaches to metabolism research 465

The dilution of the labeled lactate molecules by endogenous unlabeled lactate molecules
(D) is calculated by the equation (Landau et al., 1998; Landau, 1999b; Radziuk and Lee,
1999):

D = [0.5(M1 + M2 + M3) + M6]/(m1 + m2 + m3)

where M1, M2, M3, and M6 are, respectively, the percentages of blood glucose molecules with
one, two, three, and six 13C atoms, i.e. isotopomers M1, M2, M3, and M6. Correspondingly, m1,
m2, and m3 are the percentages for blood lactate of isotopomers m1, m2, and m3, respectively.
The fraction of glucose molecules in the blood that recycled (F), i.e. via the Cori cycle, is
calculated by the equation (Landau et al., 1998; Landau, 1999b; Radziuk and Lee, 1999):

F = 0.5(M1 + M2 + M3)/[0.5(M1 + M2 + M3) + M6]

The product of the Cori cycle and the dilution of glycolysis by endogenous lactate represents
the contribution of gluconeogenesis to the Ra glucose (Katz and Tayek, 1999). Thus, the frac-
tional gluconeogenesis (% of glucose production) can be calculated by the following equation
(Landau, 1999b; Radziuk and Lee, 1999; Mao et al., 2002), assuming that there is no loss of
labeled molecules via the tricarboxylic acid cycle because when mi → mj, i ≥ j, labeled mol-
ecule is still counted (Kelleher, 1999; Radziuk and Lee, 1999):

Fractional gluconeogenesis (% of glucose production) = (M1 + M2 + M3)/[2(m1 + m2 + m3)]

Fractional gluconeogenesis can be derived directly by a binominal expansion approach


(Kelleher, 1999; Radziuk and Lee, 1999):

Gluconeogenesis (% of glucose production) = (M1 + M2 + M3)/[2 ⋅ m0 ⋅ (m1 + m2 + m3)]

where m0 is approximate to 1. Fractional gluconeogenesis calculated from these equations is


underestimated (Landau et al., 1998; Kelleher, 1999; Landau, 1999b; Radziuk and Lee, 1999;
Mao et al., 2002), which results from the lack of isotope equilibrium in both the lactate (m3)
and glucose (M3) compartments and the tracer dilution by other unlabeled gluconeogenic
substrates (Mao et al., 2002).
Finally,

Rate of gluconeogenesis = Ra glucose × D × F

Equations for D and F are applicable only when the rate of glucose infused is small relative
to glucose production, which will result in relatively low enrichments and with negligible
formation of M4 and M5 as well as M6 isotopomers.

9. NUCLEAR MAGNETIC RESONANCE SPECTROSCOPY


Nuclear magnetic resonance (NMR) spectroscopy now provides a noninvasive means to
monitor metabolic flux and intracellular metabolite concentrations continuously. The basic
principles of in vivo NMR spectroscopy have been described in detail (Roden and Shulman,
1999). In brief, some atomic nuclei (e.g. 1H, 13C, and 31P) possess magnetic properties, i.e. the
magnetic moment or “spin”. Under experimental conditions, resonant waves (resonance)
from various nuclei/compounds can be translated into a display of peak intensities vs. fre-
quencies. The frequency of a peak is the characteristic of a certain nucleus/compound and the
area under that peak corresponds to the concentration of that nucleus/compound. The ability
to distinguish between different molecules containing the same nucleus relies on the “chemical
shift”, given in parts per million (ppm). The nuclei of different molecules thereby experience
466 X. Guan and D. G. Burrin

an altered static magnetic field and in turn resonate at an altered frequency, i.e. chemical shift,
which is typical for the respective molecule.
Several measures are used to improve the signal-to-noise ratio of NMR spectroscopy.
Increasing the field strength of the static magnetic field improves the signal-to-noise ratio and
thereby the sensitivity of the technique. Studies in humans are routinely performed at 1.5–4.7
Tesla. To examine a defined small volume of tissue, surface coils are placed tightly over the
region of interest to ensure homogeneous tissue filling in that region. The pulse angle and shape
can be selected to suppress signals from other tissues such as the subcutaneous fat layer. In vivo
NMR spectroscopy can measure the concentrations and synthesis rates of individual biologi-
cal molecules such as glycogen and neurotransmitters within precisely defined areas of specific
organs such as brain, liver, and muscle (Shulman and Rothman, 2001). Most studies to
date have used 1H, 31P, and 13C to determine skeletal muscle glucose and glycogen metabolism.
The suitability of a nucleus for NMR spectroscopy depends on its relative magnetic sensitiv-
ity, the tissue concentration range of the metabolite, and the chemical shift range.

9.1. 1H NMR spectroscopy

Protons (1H) have a natural abundance close to 100% and overall offer the highest sensitivity
for NMR spectroscopy. However, the relatively low concentration of metabolites (compared
to the proton concentration in water) and the low chemical shift range (10 ppm) have limited
the use of 1H for NMR spectroscopy. Measurement of intracellular triglyceride (TG) content
in vivo at 1.5 Tesla by 1H NMR spectroscopy has been validated biochemically by liver
biopsy (Szczepaniak et al., 1999). Furthermore, utilization of intramyocellular lipid in human
muscle is measured by 1H NMR spectroscopy (Szczepaniak et al., 1999; Krssak et al., 2000).

9.2. 31P NMR spectroscopy

Phosphors (31P) occur 100% in nature and allow quantification of intramuscular concentra-
tions of adenosine triphosphate (ATP), adenosine diphosphate, inorganic phosphate,
phosphocreatine, and glucose-6-phosphate (G6P) (Krebs et al., 2001). The concentrations of
metabolites are determined by comparing the spectral areas to the area of the β-ATP resonance,
which is used as an internal concentration standard (Bloch et al., 1993). Measurement of mus-
cular G6P concentrations by 31P NMR spectroscopy has been validated by a chemical assay
of its concentration in rat muscle frozen in situ (Bloch et al., 1993). Glucose-6-phosphate is
an intermediate in the muscle glycogen synthesis pathway, and its concentration depends on
the relative activities of muscle glycogen synthase enzyme and glucose transport into muscle.
In addition, 31P NMR spectroscopy has been used to measure mitochondrial unidirectional
ATP synthesis flux in vivo in rat skeletal muscle (Jucker et al., 2000a,b) and to measure G6P
concentration in human muscle (Rothman et al., 1995).

9.3. 13C NMR spectroscopy

In contrast to 1H and 31P, 13C has a natural abundance of 1.1% and therefore a relatively low
sensitivity. Nevertheless, 13C NMR spectroscopy has been used to measure hepatic glycogen
concentrations and thus estimate rates of net hepatic glycogen synthesis and glycogenolysis
in vivo. Since the resonance of 13C in the C-1 position of glycogen is clearly resolved at 100.5 ppm
and all 13C signals from glycogen are detected by 13C NMR spectroscopy, it can be used to meas-
ure 13C incorporation into glycogen during an infusion of [1-13C]glucose, which increases the
Methodological approaches to metabolism research 467

sensitivity of the method by up to 100-fold. Measurement of tissue glycogen content by 13C


NMR spectroscopy has been validated for skeletal muscle and liver by comparison with muscle
(Gruetter et al., 1991) and liver (Gruetter et al., 1994) biopsies (Taylor et al., 1992; Krssak et al.,
2000). Furthermore, using a 13C-glucose pulse–12C-glucose chase experiment, rates of hepatic
glycogen synthesis and glycogenolysis can be assessed (Magnusson et al., 1994; Roden et al.,
1996; Petersen et al., 1998). The peak intensity of the C-1 resonance of the glycosyl units of
glycogen is monitored with 13C NMR spectroscopy during [1-13C]glucose infusion followed by
unlabeled glucose infusion. Increment in the C-1 peak intensity during the [1-13C]glucose infu-
sion represents glycogen synthesis, while decline in the C-1 peak intensity during unlabeled
glucose infusion reflects glycogenolysis (Magnusson et al., 1994). Increments in muscle glyco-
gen concentration can be calculated from the change in [1-13C]glycogen concentration and the
isotopic enrichment of plasma [1-13C]glucose (Shulman et al., 1990).
In human and rat brains 13C NMR measurements of the in vivo flux of 13C label from [1-13C]
glucose into glutamate and glutamine simultaneously determine the rate of glucose oxidation
(tricarboxylic acid cycle rate) and glutamate/glutamine neurotransmitter cycling between
astroglia and neurons (Sibson et al., 1998, 2001; Shen et al., 1999; Shulman et al., 2001). The
glutamate/glutamine neurotransmitter cycling, measured by 13C NMR spectroscopy, is the
major pathway for neuronal glutamate repletion (Lebon et al., 2002), which accounts for 80%
of glucose oxidation in the resting state (Shen et al., 1999). 1H-decoupled 13C NMR spectra
yields sufficient signal-to-noise resonance at C-4 glutamate and C-4 glutamine in the rat brain
in vivo at 7.0 Tesla (Sibson et al., 1998). It is possible to detect 13C labeling of glutamate and
glutamine in liver by 13C NMR spectroscopy. Additionally, the in vivo 13C labeling kinetics of
glutamate and glutamine in liver and glutamine in blood can be used to calculate the liver tri-
carboxylic acid cycle flux (Jucker et al., 1998). 13C NMR and 31P NMR can be combined to
quantify glycogen synthesis rate and glucose-6-phosphate concentration in rat gastrocnemius
muscle (Chase et al., 2001). The concentration of glycogen is calculated from the increment in
the 13C spectra and the isotopic enrichment of [1-13C]glucose (Bloch et al., 1994).

10. FUTURE PERSPECTVES


In this chapter, we have discussed some new approaches aimed at understanding the biolog-
ical basis of metabolomics from systemic physiology, to intermediary metabolism, and to
molecular regulation of critical gene and protein expression. Metabolomics has recently been
developed as a platform for the quantitative measurement of the dynamic multiparametric
metabolic response of living systems to genetic modification, developmental state, patho-
physiological process, or environmental stimulus, which promises to identify gene function,
evaluate drug efficacy and toxicity, and define in vivo metabolic profiling (of all the metabo-
lites in an intact tissue, organ, or biofluid) (Raamsdonk et al., 2001; Brindle et al., 2002;
Nicholson et al., 2002; Watkins et al., 2002). Metabolomics is becoming feasible directly in
crude biological extracts with advances in nuclear magnetic resonance spectroscopy, mass
spectrometry coupled with bioinformatics techniques, and multivariate statistical analyses. In
fact, the metabolic status of an integrated biological system can be defined by its spectral
metabolic profile. Because of metabolic dynamics caused by coordinated biochemical and
molecular events, metabolic profiles are spatial-specific and temporal-dependent in response
to developmental state and environmental stimuli, which may mirror tissue-specific and time-
related changes in transcriptomic and proteomic patterns, thus limiting any physiological
relevance of single-time-point measurements of gene expression and protein abundance.
Moreover, metabolomics may provide the most direct linkage between genetic function,
468 X. Guan and D. G. Burrin

metabolic pathway, and physiological process to decipher metabolic networks (e.g. control of
glycolysis).
In molecular regulation, identifying changes in gene expression using cDNA microarrays
is just the start of a long journey from tissue to cell. At this step, the principal aim is to assem-
ble microarray hits into groups for particular metabolic pathways and/or functional processes
that provide an intelligible story of a cell’s state, or its metabolic responses to stimuli. Then,
it is usual to select a subset of these genes to independently validate changes in their expres-
sion. Combination of laser capture microdissection with real-time quantitative RT-PCR is a
helpful follow-up step that allows expression of selected genes to be quantified in a pure pop-
ulation of defined individual cells. The voyage from chip to single cell can be completed
using sensitive new in situ hybridization and immunohistochemical methods based on
tyramide signal amplification to identify cells that express mRNAs and proteins of interest
(Mills et al., 2001). Finally, RNA interference can be used as a specific and efficient method
to silence gene expression in mammalian cells and to confirm gene function on a whole-
genome scale (McManus and Sharp, 2002).
In intermediary metabolism, stable isotopic tracer methodology has become the most pow-
erful tool to quantify metabolic fluxes both in the whole body and across an organ. For
example, the arterio-venous tracer balance approach and mass isotopomer distribution analy-
sis have been widely used to estimate in vivo enzyme activity (e.g. NOS activity) and nutrient
metabolism (e.g. protein synthesis and breakdown, lipogenesis and lipolysis, and gluconeoge-
nesis). In the future, it will be possible to integrate data from transcriptomics, proteomics, and
metabolomics to provide an in vivo holistic picture of gene function and metabolic control
(Nicholson et al., 2002; Fiehn and Weckwerth, 2003).

REFERENCES
Aarsland, A., Chinkes, D., Wolfe, R.R., 1996. Contributions of de novo synthesis of fatty acids to total
VLDL-triglyceride secretion during prolonged hyperglycemia/hyperinsulinemia in normal man.
J. Clin. Invest. 98, 2008–2017.
Ackermans, M.T., Pereira Arias, A.M., Bisschop, P.H., Endert, E., Sauerwein, H.P., Romijn, J.A., 2001.
The quantification of gluconeogenesis in healthy men by 2H2O and [2-13C]glycerol yields different
results: rates of gluconeogenesis in healthy men measured with 2H2O are higher than those measured
with [2-13C]glycerol. J. Clin. Endocrinol. Metab. 86, 2220–2226.
Adegoke, O.A., McBurney, M.I., Baracos, V.E., 1999a. Jejunal mucosal protein synthesis: validation of
luminal flooding dose method and effect of luminal osmolarity. Amer. J. Physiol. 276, G14–G20.
Adegoke, O.A., McBurney, M.I., Samuels, S.E., Baracos, V.E., 1999b. Luminal amino acids acutely
decrease intestinal mucosal protein synthesis and protease mRNA in piglets. J. Nutr. 129, 1871–1878.
Ahlman, B., Charlton, M., Fu, A., Berg, C., O’Brien, P., Nair, K.S., 2001. Insulin’s effect on synthesis rates
of liver proteins: a swine model comparing various precursors of protein synthesis. Diabetes 50, 947–954.
Antelo, F., Strawford, A., Neese, R.A., Christiansen, M., Hellerstein, M., 2002. Adipose triglyceride (TG)
turnover and de novo lipogenesis (DNL) in humans: measurement by long-term 2H2O labeling and
mass isotopomer distribution analysis (MIDA). FASEB J. 16, A400.
Arcellana-Panlilio, M., Robbins, S.M., 2002. Cutting-edge technology. I. Global gene expression profiling
using DNA microarrays. Amer. J. Physiol. Gastrointest. Liver Physiol. 282, G397–G402.
Ballmer, P.E., McNurlan, M.A., Milne, E., Heys, S.D., Buchan, V., Calder, A.G., Garlick, P.J., 1990.
Measurement of albumin synthesis in humans: a new approach employing stable isotopes. Amer.
J. Physiol. 259, E797–E803.
Barazzoni, R., Meek, S.E., Ekberg, K., Wahren, J., Nair, K.S., 1999. Arterial KIC as marker of liver and muscle
intracellular leucine pools in healthy and type 1 diabetic humans. Amer. J. Physiol. 277, E238–E244.
Barle, H., Essen, P., Nyberg, B., Olivecrona, H., Tally, M., McNurlan, M.A., Wernerman, J., Garlick, P.J.,
1999. Depression of liver protein synthesis during surgery is prevented by growth hormone. Amer.
J. Physiol. 276, E620–E627.
Methodological approaches to metabolism research 469

Barrett, E.J., Revkin, J.H., Young, L.H., Zaret, B.L., Jacob, R., Gelfand, R.A., 1987. An isotopic method
for measurement of muscle protein synthesis and degradation in vivo. Biochem. J. 245, 223–228.
Basile-Filho, A., Beaumier, L., El-Khoury, A.E., Yu, Y.M., Kenneway, M., Gleason, R.E., Young, V.R.,
1998. Twenty-four-hour L-[1-13C]tyrosine and L-[3,3-2H2]phenylalanine oral tracer studies at generous,
intermediate, and low phenylalanine intakes to estimate aromatic amino acid requirements in adults.
Amer. J. Clin. Nutr. 67, 640–659.
Bassilian, S., Ahmed, S., Lim, S.K., Boros, L.G., Mao, C.S., Lee, W.N., 2002. Loss of regulation of lipoge-
nesis in the Zucker diabetic rat. II. Changes in stearate and oleate synthesis. Amer. J. Physiol. Endocrinol.
Metab. 282, E507–E513.
Bednar, M., 2000. DNA microarray technology and application. Med. Sci. Monit. 6, 796–800.
Bequette, B.J., Hanigan, M.D., Calder, A.G., Reynolds, C.K., Lobley, G.E., MacRae, J.C., 2000. Amino
acid exchange by the mammary gland of lactating goats when histidine limits milk production.
J. Dairy Sci. 83, 765–775.
Bergeron, R., Previs, S.F., Cline, G.W., Perret, P., Russell, R.R. 3rd, Young, L.H., Shulman, G.I., 2001.
Effect of 5-aminoimidazole-4-carboxamide-1-beta-D-ribofuranoside infusion on in vivo glucose and
lipid metabolism in lean and obese Zucker rats. Diabetes 50, 1076–1082.
Betsuyaku, T., Griffin, G.L., Watson, M.A., Senior, R.M., 2001. Laser capture microdissection and real-
time reverse transcriptase/polymerase chain reaction of bronchiolar epithelium after bleomycin.
Amer. J. Respir. Cell Mol. Biol. 25, 278–284.
Biolo, G., Chinkes, D., Zhang, X.J., Wolfe, R.R., 1992. Harry M. Vars Research Award. A new model to
determine in vivo the relationship between amino acid transmembrane transport and protein kinetics
in muscle. J. Parenter. Enteral Nutr. 16, 305–315.
Biolo, G., Fleming, R.Y., Maggi, S.P., Wolfe, R.R., 1995a. Transmembrane transport and intracellular
kinetics of amino acids in human skeletal muscle. Amer. J. Physiol. 268, E75–E84.
Biolo, G., Williams, B.D., Fleming, R.Y., Wolfe, R.R., 1999. Insulin action on muscle protein kinetics
and amino acid transport during recovery after resistance exercise. Diabetes 48, 949–957.
Biolo, G., Zhang, X.J., Wolfe, R.R., 1995b. Role of membrane transport in interorgan amino acid flow
between muscle and small intestine. Metabolism 44, 719–724.
Bloch, G., Chase, J.R., Avison, M.J., Shulman, R.G., 1993. In vivo 31P NMR measurement of glucose-
6-phosphate in the rat muscle after exercise. Magnet. Reson. Med. 30, 347–350.
Bloch, G., Chase, J.R., Meyer, D.B., Avison, M.J., Shulman, G.I., Shulman, R.G., 1994. In vivo regula-
tion of rat muscle glycogen resynthesis after intense exercise. Amer. J. Physiol. 266, E85–E91.
Bonner, R.F., Emmert-Buck, M., Cole, K., Pohida, T., Chuaqui, R., Goldstein, S., Liotta, L.A., 1997.
Laser capture microdissection: molecular analysis of tissue. Science 278, 1481–1483.
Bouteloup-Demange, C., Boirie, Y., Dechelotte, P., Gachon, P., Beaufrere, B., 1998. Gut mucosal protein
synthesis in fed and fasted humans. Amer. J. Physiol. 274, E541–E546.
Brindle, J.T., Antti, H., Holmes, E., Tranter, G., Nicholson, J.K., Bethell, H.W., Clarke, S., Schofield, P.M.,
McKilligin, E., Mosedale, D.E., Grainger, D.J., 2002. Rapid and noninvasive diagnosis of the
presence and severity of coronary heart disease using 1H-NMR-based metabonomics. Nat. Med. 8,
1439–1444.
Bronson, S.K., Smithies, O., 1994. Altering mice by homologous recombination using embryonic stem
cells. J. Biol. Chem. 269, 27155–27158.
Bruhat, A., Averous, J., Carraro, V., Zhong, C., Reimold, A.M., Kilberg, M.S., Fafournoux, P., 2002.
Differences in the molecular mechanisms involved in the transcriptional activation of the CHOP and
asparagine synthetase genes in response to amino acid deprivation or activation of the unfolded
protein response. J. Biol. Chem. 277, 48107–48114.
Bruhat, A., Jousse, C., Carraro, V., Reimold, A.M., Ferrara, M., Fafournoux, P., 2000. Amino acids con-
trol mammalian gene transcription: activating transcription factor 2 is essential for the amino acid
responsiveness of the CHOP promoter. Mol. Cell Biol. 20, 7192–7204.
Bruhat, A., Jousse, C., Fafournoux, P., 1999. Amino acid limitation regulates gene expression. Proc. Nutr.
Soc. 58, 625–632.
Bruhat, A., Jousse, C., Wang, X.Z., Ron, D., Ferrara, M., Fafournoux, P., 1997. Amino acid limitation
induces expression of CHOP, a CCAAT/enhancer binding protein-related gene, at both transcriptional
and post-transcriptional levels. J. Biol. Chem. 272, 17588–17593.
Bruins, M.J., Lamers, W.H., Meijer, A.J., Soeters, P.B., Deutz, N.E., 2002. In vivo measurement of nitric
oxide production in porcine gut, liver and muscle during hyperdynamic endotoxaemia. Brit. J. Pharmacol.
137, 1225–1236.
470 X. Guan and D. G. Burrin

Bustin, S.A., 2000. Absolute quantification of mRNA using real-time reverse transcription polymerase
chain reaction assays. J. Mol. Endocrinol. 25, 169–193.
Bustin, S.A., 2002. Quantification of mRNA using real-time reverse transcription PCR (RT-PCR): trends
and problems. J. Mol. Endocrinol. 29, 23–39.
Chandramouli, V., Ekberg, K., Schumann, W.C., Kalhan, S.C., Wahren, J., Landau, B.R., 1997.
Quantifying gluconeogenesis during fasting. Amer. J. Physiol. 273, E1209–E1215.
Charlton, M., Ahlman, B., Nair, K.S., 2000. The effect of insulin on human small intestinal mucosal
protein synthesis. Gastroenterology 118, 299–306.
Chase, J.R., Rothman, D.L., Shulman, R.G., 2001. Flux control in the rat gastrocnemius glycogen
synthesis pathway by in vivo 13C/31P NMR spectroscopy. Amer. J. Physiol. 280, E598–E607.
Chinkes, D., Klein, S., Zhang, X.J., Wolfe, R.R., 1996a. Infusion of labeled KIC is more accurate than
labeled leucine to determine human muscle protein synthesis. Amer. J. Physiol. 270, E67–E71.
Chinkes, D.L., Aarsland, A., Rosenblatt, J., Wolfe, R.R., 1996b. Comparison of mass isotopomer dilution
methods used to compute VLDL production in vivo. Amer. J. Physiol. 271, E373–E383.
Clarke, J.T., Bier, D.M., 1982. The conversion of phenylalanine to tyrosine in man: direct measurement
by continuous intravenous tracer infusions of L-[ring-2H5]phenylalanine and L-[1-13C] tyrosine in the
postabsorptive state. Metabolism 31, 999–1005.
Clarke, S.D., 2001. Polyunsaturated fatty acid regulation of gene transcription: a molecular mechanism
to improve the metabolic syndrome. J. Nutr. 131, 1129–1132.
Craven, R.A., Banks, R.E., 2002. Use of laser capture microdissection to selectively obtain distinct
populations of cells for proteomic analysis. Methods Enzymol. 356, 33–49.
Craven, R.A., Totty, N., Harnden, P., Selby, P.J., Banks, R.E., 2002. Laser capture microdissection and
two-dimensional polyacrylamide gel electrophoresis: evaluation of tissue preparation and sample
limitations. Amer. J. Pathol. 160, 815–822.
Crenn, P., Thuillier, F., Rakatoambinina, B., Rongier, M., Darmaun, D., Messing, B., 2000. Duodenal vs.
gastric administration of labeled leucine for the study of splanchnic metabolism in humans. J. Appl.
Physiol. 89, 573–580.
Darzynkiewicz, Z., Bedner, E., Smolewski, P., Lee, B.W., Johnson, G.L., 2002. Detection of caspases
activation in situ by fluorochrome-labeled inhibitors of caspases (FLICA). Methods Mol. Biol. 203,
289–299.
Davis, T.A., Fiorotto, M.L., Nguyen, H.V., Burrin, D.G., 1999. Aminoacyl-tRNA and tissue free amino
acid pools are equilibrated after a flooding dose of phenylalanine. Amer. J. Physiol. 277, E103–E109.
Davis, T.A., Reeds, P.J., 2001. Of flux and flooding: the advantages and problems of different isotopic
methods for quantifying protein turnover in vivo II. Methods based on the incorporation of a tracer.
Curr. Opin. Clin. Nutr. Metab. Care 4, 51–56.
Di Buono, M., Jones, P.J., Beaumier, L., Wykes, L.J., 2000. Comparison of deuterium incorporation and
mass isotopomer distribution analysis for measurement of human cholesterol biosynthesis. J. Lipid
Res. 41, 1516–1523.
Diraison, F., Pachiaudi, C., Beylot, M., 1997. Measuring lipogenesis and cholesterol synthesis in humans
with deuterated water: use of simple gas chromatographic/mass spectrometric techniques. J. Mass
Spectrom. 32, 81–86.
Dudley, M.A., Burrin, D.G., Wykes, L.J., Toffolo, G., Cobelli, C., Nichols, B.L., Rosenberger, J., Jahoor, F.,
Reeds, P.J., 1998. Protein kinetics determined in vivo with a multiple-tracer, single-sample protocol:
application to lactase synthesis. Amer. J. Physiol. 274, G591–G598.
Duplus, E., Glorian, M., Forest, C., 2000. Fatty acid regulation of gene transcription. J. Biol. Chem. 275,
30749–30752.
Emmert-Buck, M.R., Bonner, R.F., Smith, P.D., Chuaqui, R.F., Zhuang, Z., Goldstein, S.R., Weiss, R.A.,
Liotta, L.A., 1996. Laser capture microdissection. Science 274, 998–1001.
Fafournoux, P., Bruhat, A., Jousse, C., 2000. Amino acid regulation of gene expression. Biochem. J. 351, 1–12.
Fend, F., Emmert-Buck, M.R., Chuaqui, R., Cole, K., Lee, J., Liotta, L.A., Raffeld, M., 1999. Immuno-
LCM: laser capture microdissection of immunostained frozen sections for mRNA analysis. Amer.
J. Pathol. 154, 61–66.
Fiehn, O., Weckwerth, W., 2003. Deciphering metabolic networks. Eur. J. Biochem. 270, 579–588.
Garlick, P.J., McNurlan, M.A., Essen, P., Wernerman, J., 1994. Measurement of tissue protein synthesis
rates in vivo: a critical analysis of contrasting methods. Amer. J. Physiol. 266, E287–E297.
Gibson, U.E., Heid, C.A., Williams, P.M., 1996. A novel method for real time quantitative RT-PCR.
Genome Res. 6, 995–1001.
Methodological approaches to metabolism research 471

Gowrie, I.J., Roudsari, A.V., Umpleby, A.M., Hovorka, R., 1999. Estimating protein turnover with a
[13N,15C]leucine tracer: a study using simulated data. J. Theor. Biol. 198, 165–172.
Gruetter, R., Magnusson, I., Rothman, D.L., Avison, M.J., Shulman, R.G., Shulman, G.I., 1994.
Validation of 13C NMR measurements of liver glycogen in vivo. Magn. Reson. Med. 31, 583–588.
Gruetter, R., Prolla, T.A., Shulman, R.G., 1991. 13C NMR visibility of rabbit muscle glycogen in vivo.
Magn. Reson. Med. 20, 327–332.
Guan, X., Bequette, B.J., Calder, G., Ku, P.K., Ames, K.N., Trottier, N.L., 2002. Amino acid availability
affects amino acid flux and protein metabolism in the porcine mammary gland. J. Nutr. 132, 1224–1234.
Guan, X., Stoll, B., Lu, X., Tappenden, K.A., Holst, J.J., Hartmann, B., Burrin, D.G., 2003. GLP-2-
mediated up-regulation of intestinal blood flow and glucose uptake is nitric oxide-dependent in
TPN-fed piglets. Gastroenterology 125, 136–147.
Guo, Z., Jensen, M.D., 1998. Intramuscular fatty acid metabolism evaluated with stable isotopic tracers.
J. Appl. Physiol. 84, 1674–1679.
Guo, Z.K., Cella, L.K., Baum, C., Ravussin, E., Schoeller, D.A., 2000. De novo lipogenesis in adipose
tissue of lean and obese women: application of deuterated water and isotope ratio mass spectrometry.
Int. J. Obes. Relat. Metab. Disord. 24, 932–937.
Haab, B.B., Dunham, M.J., Brown, P.O., 2001. Protein microarrays for highly parallel detection and
quantitation of specific proteins and antibodies in complex solutions. Genome Biol. 2, 1–13.
Halseth, A.E., Flakoll, P.J., Reed, E.K., Messina, A.B., Krishna, M.G., Lacy, D.B., Williams, P.E.,
Wasserman, D.H., 1997. Effect of physical activity and fasting on gut and liver proteolysis in the dog.
Amer. J. Physiol. 273, E1073–E1082.
Hammond, S.M., Caudy, A.A., Hannon, G.J., 2001. Post-transcriptional gene silencing by double-
stranded RNA. Nat. Rev. Genet. 2, 110–119.
Hannon, G.J., 2002. RNA interference. Nature 418, 244–251.
Heid, C.A., Stevens, J., Livak, K.J., Williams, P.M., 1996. Real time quantitative PCR. Genome Res. 6,
986–994.
Hellerstein, M.K., 1999. Measurement of T-cell kinetics: recent methodologic advances. Immunol. Today
20, 438–441.
Hellerstein, M., Hanley, M.B., Cesar, D., Siler, S., Papageorgopoulos, C., Wieder, E., Schmidt, D., Hoh, R.,
Neese, R., Macallan, D., Deeks, S., McCune, J.M., 1999. Directly measured kinetics of circulating
T lymphocytes in normal and HIV-1-infected humans. Nat. Med. 5, 83–89.
Hellerstein, M.K., Neese, R.A., 1999. Mass isotopomer distribution analysis at eight years: theoretical,
analytic, and experimental considerations. Amer. J. Physiol. 276, E1146–E1170.
Hellerstein, M.K., Neese, R.A., Kim, Y.K., Valerie, S.S., Michelle, C., 2002. Measurement of synthesis
rates of slow-turnover proteins from 2H2O incorporation into non-essential amino acids (NEAA) and
application of mass isotopomer distribution analysis (MIDA). FASEB J. 16, A256.
Hesketh, J.E., Vasconcelos, M.H., Bermano, G., 1998. Regulatory signals in messenger RNA: determi-
nants of nutrient-gene interaction and metabolic compartmentation. Brit. J. Nutr. 80, 307–321.
Horowitz, J.F., Coppack, S.W., Paramore, D., Cryer, P.E., Zhao, G., Klein, S., 1999. Effect of short-term
fasting on lipid kinetics in lean and obese women. Amer. J. Physiol. 276, E278–E284.
Huang, R.P., 2001. Detection of multiple proteins in an antibody-based protein microarray system.
J. Immunol. Methods 255, 1–13.
Izant, J.G., Weintraub, H., 1985. Constitutive and conditional suppression of exogenous and endogenous
genes by anti-sense RNA. Science 229, 345–352.
Jucker, B.M., Dufour, S., Ren, J., Cao, X., Previs, S.F., Underhill, B., Cadman, K.S., Shulman, G.I.,
2000a. Assessment of mitochondrial energy coupling in vivo by 13C/31P NMR. Proc. Natl. Acad. Sci.
USA 97, 6880–6884.
Jucker, B.M., Lee, J.Y., Shulman, R.G., 1998. In vivo 13C NMR measurements of hepatocellular tricar-
boxylic acid cycle flux. J. Biol. Chem. 273, 12187–12194.
Jucker, B.M., Ren, J., Dufour, S., Cao, X., Previs, S.F., Cadman, K.S., Shulman, G.I., 2000b. 13C/31P
NMR assessment of mitochondrial energy coupling in skeletal muscle of awake fed and fasted rats:
relationship with uncoupling protein 3 expression. J. Biol. Chem. 275, 39279–39286.
Katome, T., Obata, T., Matsushima, R., Masuyama, N., Cantley, L.C., Gotoh, Y., Kishi, K., Shiota, H.,
Ebina, Y., 2003. Use of RNA-interference-mediated gene silencing and adenoviral overexpression to
elucidate the roles of AKT/PKB-isoforms in insulin actions. J. Biol. Chem., 278, 28312–28323.
Katz, J., Tayek, J.A., 1999. Recycling of glucose and determination of the Cori cycle and gluconeogenesis.
Amer. J. Physiol. 277, E401–E407.
472 X. Guan and D. G. Burrin

Kelleher, J.K., 1999. Estimating gluconeogenesis with [U-13C]glucose: molecular condensation requires
a molecular approach. Amer. J. Physiol. 277, E395–E400.
Kodadek, T., 2001. Protein microarrays: prospects and problems. Chem. Biol. 8, 105–115.
Kohler, U., Ayre, B.G., Goodman, H.M., Haseloff, J., 1999. Trans-splicing ribozymes for targeted gene
delivery. J. Mol. Biol. 285, 1935–1950.
Krebs, M., Krssak, M., Nowotny, P., Weghuber, D., Gruber, S., Mlynarik, V., Bischof, M., Stingl, H.,
Furnsinn, C., Waldhausl, W., Roden, M., 2001. Free fatty acids inhibit the glucose-stimulated increase of
intramuscular glucose-6-phosphate concentration in humans. J. Clin. Endocrinol. Metab. 86, 2153–2160.
Krssak, M., Petersen, K.F., Bergeron, R., Price, T., Laurent, D., Rothman, D.L., Roden, M., Shulman,
G.I., 2000. Intramuscular glycogen and intramyocellular lipid utilization during prolonged exercise
and recovery in man: a 13C and 1H nuclear magnetic resonance spectroscopy study. J. Clin.
Endocrinol. Metab. 85, 748–754.
Kurland, I.J., Alcivar, A., Bassilian, S., Lee, W.N., 2000. Loss of [13C]glycerol carbon via the pentose
cycle: implications for gluconeogenesis measurement by mass isotoper distribution analysis. J. Biol.
Chem. 275, 36787–36793.
Landau, B.R., 1999a. Glycerol production and utilization measured using stable isotopes. Proc. Nutr.
Soc. 58, 973–978.
Landau, B.R., 1999b. Limitations in the use of [U-13C6]glucose to estimate gluconeogenesis. Amer.
J. Physiol. 277, E408–E413.
Landau, B.R., 1999c. Quantifying the contribution of gluconeogenesis to glucose production in fasted
human subjects using stable isotopes. Proc. Nutr. Soc. 58, 963–972.
Landau, B.R., Wahren, J., Chandramouli, V., Schumann, W.C., Ekberg, K., Kalhan, S.C., 1996.
Contributions of gluconeogenesis to glucose production in the fasted state. J. Clin. Invest. 98, 378–385.
Landau, B.R., Wahren, J., Ekberg, K., Previs, S.F., Yang, D., Brunengraber, H., 1998. Limitations in esti-
mating gluconeogenesis and Cori cycling from mass isotopomer distributions using [U-13C6]glucose.
Amer. J. Physiol. 274, E954–E961.
Lebon, V., Petersen, K.F., Cline, G.W., Shen, J., Mason, G.F., Dufour, S., Behar, K.L., Shulman, G.I.,
Rothman, D.L., 2002. Astroglial contribution to brain energy metabolism in humans revealed by 13C
nuclear magnetic resonance spectroscopy: elucidation of the dominant pathway for neurotransmitter
glutamate repletion and measurement of astrocytic oxidative metabolism. J. Neurosci. 22, 1523–1531.
Lee Kang, S.H., Vieira, K., Bungert, J., 2002. Combining chromatin immunoprecipitation and DNA foot-
printing: a novel method to analyze protein-DNA interactions in vivo. Nucleic Acids Res. 30, e44.
Lemieux, S., Patterson, B.W., Carpentier, A., Lewis, G.F., Steiner, G., 1999. A stable isotope method
using a [2H5]glycerol bolus to measure very low density lipoprotein triglyceride kinetics in humans.
J. Lipid Res. 40, 2111–2117.
Lindenthal, B., Aldaghlas, T.A., Holleran, A.L., Sudhop, T., Berthold, H.K., Von Bergmann, K., Kelleher,
J.K., 2002. Isotopomer spectral analysis of intermediates of cholesterol synthesis in human subjects
and hepatic cells. Amer. J. Physiol. 282, E1222–E1230.
Ljungqvist, O.H., Persson, M., Ford, G.C., Nair, K.S., 1997. Functional heterogeneity of leucine pools in
human skeletal muscle. Amer. J. Physiol. 273, E564–E570.
Lockhart, D.J., Winzeler, E.A., 2000. Genomics, gene expression and DNA arrays. Nature 405, 827–836.
Luzzi, V., Holtschlag, V., Watson, M.A., 2001. Expression profiling of ductal carcinoma in situ by laser
capture microdissection and high-density oligonucleotide arrays. Amer. J. Pathol. 158, 2005–2010.
Macallan, D.C., Fullerton, C.A., Neese, R.A., Haddock, K., Park, S.S., Hellerstein, M.K., 1998.
Measurement of cell proliferation by labeling of DNA with stable isotope-labeled glucose: studies in
vitro, in animals, and in humans. Proc. Natl. Acad. Sci. USA 95, 708–713.
Maghni, K., Nicolescu, O.M., Martin, J.G., 1999. Suitability of cell metabolic colorimetric assays for
assessment of CD4+ T cell proliferation: comparison to 5-bromo-2-deoxyuridine (BrdU) ELISA.
J. Immunol. Methods 223, 185–194.
Magnusson, I., Rothman, D.L., Jucker, B., Cline, G.W., Shulman, R.G., Shulman, G.I., 1994. Liver
glycogen turnover in fed and fasted humans. Amer. J. Physiol. 266, E796–E803.
Mansoor, O., Cayol, M., Gachon, P., Boirie, Y., Schoeffler, P., Obled, C., Beaufrere, B., 1997. Albumin
and fibrinogen syntheses increase while muscle protein synthesis decreases in head-injured patients.
Amer. J. Physiol. 273, E898–E902.
Mao, C.S., Bassilian, S., Lim, S.K., Lee, W.N., 2002. Underestimation of gluconeogenesis by the
[U-13C6]glucose method: effect of lack of isotope equilibrium. Amer. J. Physiol. Endocrinol. Metab.
282, E376–E385.
Methodological approaches to metabolism research 473

Marshman, E., Ottewell, P.D., Potten, C.S., Watson, A.J., 2001. Caspase activation during spontaneous
and radiation-induced apoptosis in the murine intestine. J. Pathol. 195, 285–292.
Martini, W.Z., Chinkes, D.L., Wolfe, R.R., 2002. DNA and protein turnover in human fibroblasts and
myocytes. FASEB J. 16, A613.
Mashimo, H., Goyal, R.K., 1999. Lessons from genetically engineered animal models. IV. Nitric oxide
synthase gene knockout mice. Amer. J. Physiol. 277, G745–G750.
Matthews, D.E., Marano, M.A., Campbell, R.G., 1993a. Splanchnic bed utilization of glutamine and
glutamic acid in humans. Amer. J. Physiol. 264, E848–E854.
Matthews, D.E., Marano, M.A., Campbell, R.G., 1993b. Splanchnic bed utilization of leucine and phenylala-
nine in humans. Amer. J. Physiol. 264, E109–E118.
Matthews, D.E., Schwarz, H.P., Yang, R.D., Motil, K.J., Young, V.R., Bier, D.M., 1982. Relationship of
plasma leucine and alpha-ketoisocaproate during a L-[1-13C]leucine infusion in man: a method for
measuring human intracellular leucine tracer enrichment. Metabolism 31, 1105–1112.
McCaffrey, A.P., Meuse, L., Pham, T.T., Conklin, D.S., Hannon, G.J., Kay, M.A., 2002. RNA interfer-
ence in adult mice. Nature 418, 38–39.
McDevitt, R.M., Bott, S.J., Harding, M., Coward, W.A., Bluck, L.J., Prentice, A.M., 2001. De novo lipo-
genesis during controlled overfeeding with sucrose or glucose in lean and obese women. Amer.
J. Clin. Nutr. 74, 737–746.
McManus, M.T., Sharp, P.A., 2002. Gene silencing in mammals by small interfering RNAs. Nat. Rev.
Genet. 3, 737–747.
Meek, S.E., Persson, M., Ford, G.C., Nair, K.S., 1998. Differential regulation of amino acid exchange
and protein dynamics across splanchnic and skeletal muscle beds by insulin in healthy human
subjects. Diabetes 47, 1824–1835.
Mills, J.C., Roth, K.A., Cagan, R.L., Gordon, J.I., 2001. DNA microarrays and beyond: completing the
journey from tissue to cell. Nat. Cell Biol. 3, E175–E178.
Mittendorfer, B., Horowitz, J.F., Klein, S., 2001. Gender differences in lipid and glucose kinetics during
short-term fasting. Amer. J. Physiol. 281, E1333–E1339.
Moller, N., Meek, S., Bigelow, M., Andrews, J., Nair, K.S., 2000. The kidney is an important site for in
vivo phenylalanine-to-tyrosine conversion in adult humans: a metabolic role of the kidney. Proc. Natl.
Acad. Sci. USA 97, 1242–1246.
Neese, R.A., Misell, L.M., Turner, S., Chu, A., Kim, J., Cesar, D., Hoh, R., Antelo, F., Strawford, A.,
McCune, J.M., Christiansen, M., Hellerstein, M.K., 2002. Measurement in vivo of proliferation rates
of slow turnover cells by 2H2O labeling of the deoxyribose moiety of DNA. Proc. Natl. Acad. Sci.
USA 99, 15345–15350.
Neese, R.A., Schwarz, J.M., Faix, D., Turner, S., Letscher, A., Vu, D., Hellerstein, M.K., 1995.
Gluconeogenesis and intrahepatic triose phosphate flux in response to fasting or substrate loads:
application of the mass isotopomer distribution analysis technique with testing of assumptions and
potential problems. J. Biol. Chem. 270, 14452–14466.
Neese, R.A., Siler, S.Q., Cesar, D., Antelo, F., Lee, D., Misell, L., Patel, K., Tehrani, S., Shah, P.,
Hellerstein, M.K., 2001. Advances in the stable isotope-mass spectrometric measurement of DNA
synthesis and cell proliferation. Anal. Biochem. 298, 189–195.
Nicholson, J.K., Connelly, J., Lindon, J.C., Holmes, E., 2002. Metabonomics: a platform for studying
drug toxicity and gene function. Nat. Rev. Drug Discov. 1, 153–161.
Palmer, R.M., Ashton, D.S., Moncada, S., 1988. Vascular endothelial cells synthesize nitric oxide from
L-arginine. Nature 333, 664–666.
Pandey, A., Mann, M., 2000. Proteomics to study genes and genomes. Nature 405, 837–846.
Paolini, C.L., Meschia, G., Fennessey, P.V., Pike, A.W., Teng, C., Battaglia, F.C., Wilkening, R.B., 2001.
An in vivo study of ovine placental transport of essential amino acids. Amer. J. Physiol. 280,
E31–E39.
Papageorgopoulos, C., Caldwell, K., Shackleton, C., Schweingrubber, H., Hellerstein, M.K., 1999.
Measuring protein synthesis by mass isotopomer distribution analysis (MIDA). Anal. Biochem.
267, 1–16.
Patterson, B.W., 1997. Use of stable isotopically labeled tracers for studies of metabolic kinetics: an
overview. Metabolism 46, 322–329.
Petersen, K.F., Krssak, M., Navarro, V., Chandramouli, V., Hundal, R., Schumann, W.C., Landau, B.R.,
Shulman, G.I., 1999. Contributions of net hepatic glycogenolysis and gluconeogenesis to glucose
production in cirrhosis. Amer. J. Physiol. 276, E529–E535.
474 X. Guan and D. G. Burrin

Petersen, K.F., Laurent, D., Rothman, D.L., Cline, G.W., Shulman, G.I., 1998. Mechanism by which
glucose and insulin inhibit net hepatic glycogenolysis in humans. J. Clin. Invest. 101, 1203–1209.
Previs, S., 2002. Application of 2H2O for estimating protein synthesis (PS) and protein breakdown (PB)
in vivo. FASEB J. 16, A788.
Previs, S.F., Fernandez, C.A., Yang, D., Soloviev, M.V., David, F., Brunengraber, H., 1995. Limitations
of the mass isotopomer distribution analysis of glucose to study gluconeogenesis: substrate cycling
between glycerol and triose phosphates in liver. J. Biol. Chem. 270, 19806–19815.
Previs, S.F., Hallowell, P.T., Neimanis, K.D., David, F., Brunengraber, H., 1998. Limitations of the mass
isotopomer distribution analysis of glucose to study gluconeogenesis: heterogeneity of glucose label-
ing in incubated hepatocytes. J. Biol. Chem. 273, 16853–16859.
Raamsdonk, L.M., Teusink, B., Broadhurst, D., Zhang, N., Hayes, A., Walsh, M.C., Berden, J.A.,
Brindle, K.M., Kell, D.B., Rowland, J.J., Westerhoff, H.V., van Dam, K., Oliver, S.G., 2001. A functional
genomics strategy that uses metabolome data to reveal the phenotype of silent mutations. Nat. Biotechnol.
19, 45–50.
Radziuk, J., Lee, W.P., 1999. Measurement of gluconeogenesis and mass isotopomer analysis based on
[U-13C]glucose. Amer. J. Physiol. 277, E199–E207.
Rapley, R., Walker, J.M., 1998. Molecular Biomethods Handbook. Humana Press, Totowa, NJ.
Reeds, P.J., Davis, T.A., 1999. Of flux and flooding: the advantages and problems of different isotopic
methods for quantifying protein turnover in vivo I. Methods based on the dilution of a tracer. Curr.
Opin. Clin. Nutr. Metab. Care 2, 23–28.
Reeds, P.J., Hachey, D.L., Patterson, B.W., Motil, K.J., Klein, P.D., 1992. VLDL apolipoprotein B-100,
a potential indicator of the isotopic labeling of the hepatic protein synthetic precursor pool in humans:
studies with multiple stable isotopically labeled amino acids. J. Nutr. 122, 457–466.
Renaudeau, D., Lebreton, Y., Noblet, J., Dourmad, J.Y., 2002. Measurement of blood flow through the
mammary gland in lactating sows: methodological aspects. J. Anim. Sci. 80, 196–201.
Rennie, M.J., Smith, K., Watt, P.W., 1994. Measurement of human tissue protein synthesis: an optimal
approach. Amer. J. Physiol. 266, E298–E307.
Reue, K., 1998. mRNA quantitation techniques: considerations for experimental design and application.
J. Nutr. 128, 2038–2044.
Roden, M., Perseghin, G., Petersen, K.F., Hwang, J.H., Cline, G.W., Gerow, K., Rothman, D.L.,
Shulman, G.I., 1996. The roles of insulin and glucagon in the regulation of hepatic glycogen synthe-
sis and turnover in humans. J. Clin. Invest. 97, 642–648.
Roden, M., Shulman, G.I., 1999. Applications of NMR spectroscopy to study muscle glycogen metabo-
lism in man. Annu. Rev. Med. 50, 277–290.
Roesler, W.J., 2001. The role of C/EBP in nutrient and hormonal regulation of gene expression. Annu.
Rev. Nutr. 21, 141–165.
Rothman, D.L., Magnusson, I., Cline, G., Gerard, D., Kahn, C.R., Shulman, R.G., Shulman, G.I., 1995.
Decreased muscle glucose transport/phosphorylation is an early defect in the pathogenesis of
non-insulin-dependent diabetes mellitus. Proc. Natl. Acad. Sci. USA 92, 983–987.
Rothman, D.L., Magnusson, I., Katz, L.D., Shulman, R.G., Shulman, G.I., 1991. Quantitation of hepatic
glycogenolysis and gluconeogenesis in fasting humans with 13C NMR. Science 254, 573–576.
Ryding, A.D., Sharp, M.G., Mullins, J.J., 2001. Conditional transgenic technologies. J. Endocrinol. 171,
1–14.
Sawada, S., Asakura, S., Daimon, H., Furihata, C., 1995. Comparison of autoradiography, liquid scintil-
lation counting and immunoenzymatic staining of 5-bromo-2′-deoxyuridine for measurement of
unscheduled DNA synthesis and replicative DNA synthesis in rat liver. Mutat. Res. 344, 109–116.
Shannon, M.F., Rao, S., 2002. Transcription. of chips and ChIPs. Science 296, 666–669.
Shen, J., Petersen, K.F., Behar, K.L., Brown, P., Nixon, T.W., Mason, G.F., Petroff, O.A., Shulman, G.I.,
Shulman, R.G., Rothman, D.L., 1999. Determination of the rate of the glutamate/glutamine cycle in
the human brain by in vivo 13C NMR. Proc. Natl. Acad. Sci. USA 96, 8235–8240.
Shi, Y., 2002. Mechanisms of caspase activation and inhibition during apoptosis. Mol. Cell 9, 459–470.
Short, K.R., Meek, S.E., Moller, N., Ekberg, K., Nair, K.S., 1999. Whole body protein kinetics using Phe
and Tyr tracers: an evaluation of the accuracy of approximated flux values. Amer. J. Physiol. 276,
E1194–E1200.
Shulman, G.I., Rothman, D.L., Jue, T., Stein, P., DeFronzo, R.A., Shulman, R.G., 1990. Quantitation of
muscle glycogen synthesis in normal subjects and subjects with non-insulin-dependent diabetes by
13C nuclear magnetic resonance spectroscopy. N. Engl. J. Med. 322, 223–228.
Methodological approaches to metabolism research 475

Shulman, R.G., Hyder, F., Rothman, D.L., 2001. Cerebral energetics and the glycogen shunt: neuro-
chemical basis of functional imaging. Proc. Natl. Acad. Sci. USA 98, 6417–6422.
Shulman, R.G., Rothman, D.L., 2001. 13C NMR of intermediary metabolism: implications for systemic
physiology. Annu. Rev. Physiol. 63, 15–48.
Sibson, N.R., Dhankhar, A., Mason, G.F., Rothman, D.L., Behar, K.L., Shulman, R.G., 1998.
Stoichiometric coupling of brain glucose metabolism and glutamatergic neuronal activity. Proc. Natl.
Acad. Sci. USA 95, 316–321.
Sibson, N.R., Mason, G.F., Shen, J., Cline, G.W., Herskovits, A.Z., Wall, J.E., Behar, K.L., Rothman, D.L.,
Shulman, R.G., 2001. In vivo 13C NMR measurement of neurotransmitter glutamate cycling, anaplero-
sis and TCA cycle flux in rat brain during [2-13C]glucose infusion. J. Neurochem. 76, 975–989.
Sidossis, L.S., Coggan, A.R., Gastaldelli, A., Wolfe, R.R., 1995a. A new correction factor for use in tracer
estimations of plasma fatty acid oxidation. Amer. J. Physiol. 269, E649–E656.
Sidossis, L.S., Coggan, A.R., Gastaldelli, A., Wolfe, R.R., 1995b. Pathway of free fatty acid oxidation in
human subjects: implications for tracer studies. J. Clin. Invest. 95, 278–284.
Sidossis, L.S., Mittendorfer, B., Chinkes, D., Walser, E., Wolfe, R.R., 1999. Effect of hyperglycemia-
hyperinsulinemia on whole body and regional fatty acid metabolism. Amer. J. Physiol. 276, E427–E434.
Sidossis, L.S., Mittendorfer, B., Walser, E., Chinkes, D., Wolfe, R.R., 1998. Hyperglycemia-induced inhi-
bition of splanchnic fatty acid oxidation increases hepatic triacylglycerol secretion. Amer. J. Physiol.
275, E798–E805.
Siler, S.Q., Neese, R.A., Hellerstein, M.K., 1999. De novo lipogenesis, lipid kinetics, and whole-body
lipid balances in humans after acute alcohol consumption. Amer. J. Clin. Nutr. 70, 928–936.
Siler, S.Q., Neese, R.A., Parks, E.J., Hellerstein, M.K., 1998. VLDL-triglyceride production after alcohol
ingestion, studied using [2-13C1] glycerol. J. Lipid Res. 39, 2319–2328.
Simone, N.L., Paweletz, C.P., Charboneau, L., Petricoin, E.F. 3rd, Liotta, L.A., 2000a. Laser capture
microdissection: beyond functional genomics to proteomics. Mol. Diagn. 5, 301–307.
Simone, N.L., Remaley, A.T., Charboneau, L., Petricoin, E.F. 3rd, Glickman, J.W., Emmert-Buck, M.R.,
Fleisher, T.A., Liotta, L.A., 2000b. Sensitive immunoassay of tissue cell proteins procured by laser
capture microdissection. Amer. J. Pathol. 156, 445–452.
Smolewski, P., Grabarek, J., Halicka, H.D., Darzynkiewicz, Z., 2002. Assay of caspase activation in situ
combined with probing plasma membrane integrity to detect three distinct stages of apoptosis.
J. Immunol. Methods 265, 111–121.
Sok, J., Wang, X.Z., Batchvarova, N., Kuroda, M., Harding, H., Ron, D., 1999. CHOP-dependent stress-
inducible expression of a novel form of carbonic anhydrase VI. Mol. Cell Biol. 19, 495–504.
Stahelin, B.J., Marti, U., Solioz, M., Zimmermann, H., Reichen, J., 1998. False positive staining in the
TUNEL assay to detect apoptosis in liver and intestine is caused by endogenous nucleases and inhib-
ited by diethyl pyrocarbonate. Mol. Pathol. 51, 204–208.
Stappenbeck, T.S., Hooper, L.V., Manchester, J.K., Wong, M.H., Gordon, J.I., 2002. Laser capture
microdissection of mouse intestine: characterizing mRNA and protein expression, and profiling inter-
mediary metabolism in specified cell populations. Methods Enzymol. 356, 167–196.
Stillman, B.A., Tonkinson, J.L., 2001. Expression microarray hybridization kinetics depend on length
of the immobilized DNA but are independent of immobilization substrate. Anal. Biochem. 295,
149–157.
Stoll, B., Burrin, D.G., Henry, J., Jahoor, F., Reeds, P.J., 1997. Phenylalanine utilization by the gut and
liver measured with intravenous and intragastric tracers in pigs. Amer. J. Physiol. 273, G1208–G1217.
Stoll, B., Burrin, D.G., Henry, J., Yu, H., Jahoor, F., Reeds, P.J., 1999a. Substrate oxidation by the portal
drained viscera of fed piglets. Amer. J. Physiol. 277, E168–E175.
Stoll, B., Burrin, D.G., Henry, J.F., Jahoor, F., Reeds, P.J., 1999b. Dietary and systemic phenylalanine uti-
lization for mucosal and hepatic constitutive protein synthesis in pigs. Amer. J. Physiol. 276, G49–G57.
Stoll, B., Chang, X., Fan, M.Z., Reeds, P.J., Burrin, D.G., 2000. Enteral nutrient intake level determines
intestinal protein synthesis and accretion rates in neonatal pigs. Amer. J. Physiol. Gastrointest. Liver
Physiol. 279, G288–G294.
Stoll, B., Henry, J., Reeds, P.J., Yu, H., Jahoor, F., Burrin, D.G., 1998. Catabolism dominates the first-
pass intestinal metabolism of dietary essential amino acids in milk protein-fed piglets. J. Nutr. 128,
606–614.
Szczepaniak, L.S., Babcock, E.E., Schick, F., Dobbins, R.L., Garg, A., Burns, D.K., McGarry, J.D., Stein, D.T.,
1999. Measurement of intracellular triglyceride stores by H spectroscopy: validation in vivo. Amer.
J. Physiol. 276, E977–E989.
476 X. Guan and D. G. Burrin

Tayek, J.A., Katz, J., 1996. Glucose production, recycling, and gluconeogenesis in normals and diabetics:
a mass isotopomer [U-13C]glucose study. Amer. J. Physiol. 270, E709–E717.
Tayek, J.A., Katz, J., 1997. Glucose production, recycling, Cori cycle, and gluconeogenesis in humans:
relationship to serum cortisol. Amer. J. Physiol. 272, E476–E484.
Taylor, R., Price, T.B., Rothman, D.L., Shulman, R.G., Shulman, G.I., 1992. Validation of 13C NMR
measurement of human skeletal muscle glycogen by direct biochemical assay of needle biopsy sam-
ples. Magnet. Reson. Med. 27, 13–20.
Tessari, P., Garibotto, G., Inchiostro, S., Robaudo, C., Saffioti, S., Vettore, M., Zanetti, M., Russo, R.,
Deferrari, G., 1996. Kidney, splanchnic, and leg protein turnover in humans: insight from leucine and
phenylalanine kinetics. J. Clin. Invest. 98, 1481–1492.
Thompson, G.N., Pacy, P.J., Merritt, H., Ford, G.C., Read, M.A., Cheng, K.N., Halliday, D., 1989. Rapid
measurement of whole body and forearm protein turnover using a [2H5]phenylalanine model. Amer.
J. Physiol. 256, E631–E639.
Toth, M.J., MacCoss, M.J., Poehlman, E.T., Matthews, D.E., 2001. Recovery of 13CO2 from infused [1-
13C]leucine and [1,2-13C ]leucine in healthy humans. Amer. J. Physiol. Endocrinol. Metab. 281,
2
E233–E241.
Trimmer, J.K., Schwarz, J.M., Casazza, G.A., Horning, M.A., Rodriguez, N., Brooks, G.A., 2002.
Measurement of gluconeogenesis in exercising men by mass isotopomer distribution analysis.
J. Appl. Physiol. 93, 233–241.
Trottier, N.L., Shipley, C.F., Easter, R.A., 1997. Plasma amino acid uptake by the mammary gland of the
lactating sow. J. Anim. Sci. 75, 1266–1278.
Turner, S.M., Neese, R.A., Murphy, E., Antelo, F., Hellerstein, M., 2002. Measurement of triglyceride
(TG) synthesis in vivo by 2H2O incorporation into TG-glycerol and application of mass isotopomer
distribution analysis (MIDA). FASEB J. 16, A400.
Ubeda, M., Wang, X.Z., Zinszner, H., Wu, I., Habener, J.F., Ron, D., 1996. Stress-induced binding of the
transcriptional factor CHOP to a novel DNA control element. Mol. Cell Biol. 16, 1479–1489.
van Goudoever, J.B., Stoll, B., Henry, J.F., Burrin, D.G., Reeds, P.J., 2000. Adaptive regulation of intes-
tinal lysine metabolism. Proc. Natl. Acad. Sci. USA 97, 11620–11625.
van Spronsen, F.J., Reijngoud, D.J., Smit, G.P., Nagel, G.T., Stellaard, F., Berger, R., Heymans, H.S.,
1998. Phenylketonuria: the in vivo hydroxylation rate of phenylalanine into tyrosine is decreased.
J. Clin. Invest. 101, 2875–2880.
Vasquez, K.M., Marburger, K., Intody, Z., Wilson, J.H., 2001. Manipulating the mammalian genome by
homologous recombination. Proc. Natl. Acad. Sci. USA 98, 8403–8410.
Volpi, E., Ferrando, A.A., Yeckel, C.W., Tipton, K.D., Wolfe, R.R., 1998. Exogenous amino acids
stimulate net muscle protein synthesis in the elderly. J. Clin. Invest. 101, 2000–2007.
Volpi, E., Jeschke, M.G., Herndon, D.N., Wolfe, R.R., 2000. Measurement of skin protein breakdown in
a rat model. Amer. J. Physiol. Endocrinol. Metab. 279, E900–E906.
Walker, J.A., Quirke, P., 2001. Viewing apoptosis through a “TUNEL”. J. Pathol. 195, 275–276.
Wang, T., Brown, M.J., 1999. mRNA quantification by real time TaqMan polymerase chain reaction:
validation and comparison with RNase protection. Anal. Biochem. 269, 198–201.
Wang, W., Basinger, A., Neese, R.A., Christiansen, M., Hellerstein, M.K., 2000. Effects of nicotinic acid
on fatty acid kinetics, fuel selection, and pathways of glucose production in women. Amer. J. Physiol.
279, E50–E59.
Waterlow, J.C., Stephen, J.M., 1967. The measurement of total lysine turnover in the rat by intravenous
infusion of L-[U-14C]lysine. Clin. Sci. 33, 489–506.
Watkins, S.M., Reifsnyder, P.R., Pan, H.J., German, J.B., Leiter, E.H., 2002. Lipid metabolome-wide
effects of the PPARgamma agonist rosiglitazone. J. Lipid Res. 43, 1809–1817.
Watt, P.W., Lindsay, Y., Scrimgeour, C.M., Chien, P.A., Gibson, J.N., Taylor, D.J., Rennie, M.J., 1991.
Isolation of aminoacyl-tRNA and its labeling with stable-isotope tracers: use in studies of human
tissue protein synthesis. Proc. Natl. Acad. Sci. USA 88, 5892–5896.
Weinmann, A.S., Yan, P.S., Oberley, M.J., Huang, T.H., Farnham, P.J., 2002. Isolating human transcrip-
tion factor targets by coupling chromatin immunoprecipitation and CpG island microarray analysis.
Gene. Dev. 16, 235–244.
Witzmann, F.A., Li, J., 2002. Cutting-edge technology. II. Proteomics: core technologies and applications
in physiology. Amer. J. Physiol. 282, G735–G741.
Wolfe, R.R., 1992. Radioactive and Stable Isotope Tracers in Biomedicine: Principles and Practice of
Kinetic Analysis. Wiley-Liss, New York.
Methodological approaches to metabolism research 477

Wong, M.H., Saam, J.R., Stappenbeck, T.S., Rexer, C.H., Gordon, J.I., 2000. Genetic mosaic analysis
based on Cre recombinase and navigated laser capture microdissection. Proc. Natl. Acad. Sci. USA
97, 12601–12606.
Xu, J., Nakamura, M.T., Cho, H.P., Clarke, S.D., 1999. Sterol regulatory element binding protein-1
expression is suppressed by dietary polyunsaturated fatty acids: a mechanism for the coordinate sup-
pression of lipogenic genes by polyunsaturated fats. J. Biol. Chem. 274, 23577–23583.
Xu, J., Teran-Garcia, M., Park, J.H., Nakamura, M.T., Clarke, S.D., 2001. Polyunsaturated fatty acids
suppress hepatic sterol regulatory element-binding protein-1 expression by accelerating transcript
decay. J. Biol. Chem. 276, 9800–9807.
Yang, V.W., 1998. Eukaryotic transcription factors: identification, characterization and functions. J. Nutr.
128, 2045–2051.
Zaidi, A.U., Enomoto, H., Milbrandt, J., Roth, K.A., 2000. Dual fluorescent in situ hybridization and
immunohistochemical detection with tyramide signal amplification. J. Histochem. Cytochem. 48,
1369–1375.
Zhang, X.J., Chinkes, D.L., Sakurai, Y., Wolfe, R.R., 1996. An isotopic method for measurement of
muscle protein fractional breakdown rate in vivo. Amer. J. Physiol. 270, E759–E767.
Zhang, X.J., Chinkes, D.L., Wolfe, R.R., 2002. Measurement of muscle protein fractional synthesis and
breakdown rates from a pulse tracer injection. Amer. J. Physiol. 283, E753–E764.
Index

3-methyl histidine, 72 β3-adrenergic receptors, 305


5′deiodinase, 309, 311, 318 Adipocytes, 305
α-cardiac MHC, 360 Brown unknown gene, 305
β-adrenergic agonists, 77, 79, 289–291 Fatty acid metabolism, 310–311
Fetal development, 305
Morphology, 304
A Quantity, 304
“Absorptive use”, 202 Thermogenesis, 304
Acetyl CoA carboxylase (ACC), 224, 226 Uncoupling protein-1, 304, 308–309,
Acute-phase protein, 84, 85, 90–92 311–312
Acyl CoA synthetase (ACS), 224, 226 Brown adipose tissue, 276, 353, 357
Adaptive immunity, 85 Brown unknown gene (BUG), 6, 305
Adipose tissue, 337
Amino acid absorption, 198, 201, 207
Amino acid oxidation, 197, 207 C
Amino acid requirements, 128 Calcineurin, 13, 44
Amino acids, 5, 6, 8–10, 24–27, 29, 35, 51, 53, Calpain, 51, 84, 98–99
55–57, 59, 107–108, 110–113, 116–118, Cardiac output, 353, 356, 361
120–122 Carnitine palmitoyltransferase (CPT), 224–225,
Ammonia absorption and liver urea synthesis, 208 229, 379, 383
Animal growth, 21, 118 Catecholamines, 14,18, 20–21, 368
Anorexia, 86–88, 90 Cathepsin, 84, 99
Apoptosis, 303, 313, 315 cDNA microarray, 437
Arachidonic acid Cell apoptosis, 446
Functional roles, 13–15 Cell culture, 6, 279
Infant nutrition, 34–40 Cell number, 277
Metabolism, 8–12, 32–34 Cell proliferation, 445
Milk composition, 26–29 Cell signalling, 180
Placental transfer, 23–26 Cell-mediated immunity, 85
Arginine, 161, 166, 168–169, 173, 177–180 Chromatin immunoprecipitation assay, 441
Arterio-venous tracer balance method, 453 Chylomicron, 329–330, 332, 340
ATP, 356, 359, 361–362, 367 Citrulline, 166, 169
Cold, 354–356, 358, 360, 362, 365, 367–368
Colostrum, 2, 4, 15, 17–19, 26, 57–58
B Compartment modeling, 449
Blood flow, 361 Composition, 281
Brain development Conceptus, 4, 6, 7, 11, 24
Docosahexaenoic acid, 18–23 Confocal laser scanning microscopy, 444
N-3 fatty acid deficiency, 15–23 Conjugated linoleic acid, 290
Polyunsaturated fatty acid accretion, 29–32 Constant infusion method, 449
Polyunsaturated fatty acid metabolism, 30–34 Copper, 318
Branched chain amino acids, 205 Corticotropin-releasing hormone, 87
Bromodeoxyuridine (BrdU) labeling assay, 445 Cortisol, 16, 18–20, 23, 368, 380, 383–384
Brown adipose tissue (BAT) Cost of urea synthesis, 208
480 Index

CPT-1, 365–366 Fatty acid synthesis, 282–290


Cysteine, 169–170, 179, 181 Fatty acids
Cytochrome c oxidase, 311–312 Metabolism, 310–311
Cytokine, 83–100 Polyunsaturated (PUFA), 316–317
Cytokines, 178, 182 Saturated, 316–317
Fatty acids, 5, 8–11, 23
Feed intake, 1–9, 83–88, 90, 94
D Feeding, 51–57
Development, 108, 119, 277, 282 Fetal programming, 4, 23
Differentiation, 277–280 Fetus, 5–8, 11, 14–16, 19, 21
Distribution, 276 First pass metabolism, 141
DNase I protection assay, 441 First-pass utilization, 458
Docosahexaenoic acid Flooding dose method, 451
Behaviour, 18–23 Functions, 276
Brain accretion, 29–32
Brain function, 18–23, 37–40
Dopamine, 19–20 G
Functional roles, 241–243 Gastrocnemius muscle, 90–91, 95–97
Infant nutrition, 254 Gene expression
Metabolism, 237–241, 253–254 Polyunsaturated fatty acids, 15
Milk composition, 248–249 Gene expression, 400
Neurotransmitters, 19–21 Glucagon, 15, 158, 164, 176–177, 379–380,
Placental transfer, 23–26 383–385
Serotonin, 21 Glucocorticoids, 176, 379, 383–385
Visual function, 15–18, 36–40 Gluconeogenesis, 6, 10, 12, 15, 213, 353, 367,
462–463
Glucose transport, 395
E Glucose, 5–8, 10, 12, 14, 17–18, 21, 23, 406
Electrophoretic mobility shift assay, 441 Glucose-6-phosphatase, 377–378
Endocrine functions, 282 Glutamate, 161, 166–171, 173, 179, 182
Endocrine regulation–anabolism, 289–290t Glutamine, 92, 160–161, 166, 168–169, 173,
Endocrine regulation–catabolism, 18–22t 176–182
Endogenous amino acid secretions, 198, 206 Glutathione, 24, 39, 164, 169–171, 179, 181
Energy metabolism, 353, 355, 359, 368–369 Glycogen synthesis, 466
Energy stores, 356, 363, 369 Glycogen, 16, 17, 378–381
Enteral, 161, 169–173, 176 Glycogenolysis, 466
Enterocytes, 108–118, 120–121 Glycolytic, 16, 22, 31, 38, 44–45, 48, 53
Essential fatty acids Growth hormone releasing hormone, 84, 93
Requirements, 40–43 Growth hormone, 15–17, 19–20, 22, 24, 29, 37, 41,
Deficiency, 7,12,13 45, 56–58, 77, 84, 93, 383–384
Gene expression, 15
Infant nutrition, 34–40
Metabolism, 8–12, 32–34 H
Milk composition, 26–29 High-density lipoprotein, 333, 341
Placental transfer, 23–26 Humoral immunity, 85
Essential fatty acids, 324, 334–335 Hydroxy-methylglutarylCoA synthase
Esterification, 329, 339, 342 (HMGCS), 224
Eukaryotic initiation factor, 94–96 Hyperplasia, 277–278
Eukaryotic initiation factors, 29, 50–53, 55–57 Hypertrophy, 281–282
Excess protein digestion, 208
Expression of enzyme data, 18, 287
I
IGF-1, 25, 30–31, 77, 175, 369
F Immune system, 83–86, 90, 93, 99–100
Fast–twitch muscle fibres, 22, 38, 44–45, Immunonutrients, 39, 157, 179, 183
53, 55 in situ hybridization, 439
Fast-twitch muscle, 91 Indicator amino acid oxidation, 128
Fatty acid binding protein, 328, 336 Innate immunity, 4–5
Fatty acid oxidation, 342–343, 353 Insulin receptor substrate, 16, 93
Index 481

Insulin, 12, 14–15, 18–19, 22–23, 28, 30, 46, 50, Infant nutrition, 254
53–59, 158, 164, 174, 176, 180–181, Metabolism, 237–241, 253–254
288–289, 292, 379–383, 385 Milk, 248–249
Insulin-like growth factor 1, 16, 20 Placental transfer, 246–248
Insulin-like growth factor 2, 15–16, 18, 20–22 Low-density lipoprotein, 341
Insulin-like growth factor binding protein, 19–21, Lysine, 165, 167, 173, 176
94–96
Insulin-like growth factor-I, 84, 92–96
Insulin-like growth factors, 40–41, 48, 56–58 M
Interferon, 85, 94 Mammalian target of rapamycin, 50, 52–53
Interleukin-1, 83–91, 93, 97 Mass isotopomer distribution analysis, 462
Interleukin-6, 84–91, 93, 97–99 MCFA, 364–365
Interleukin-1 receptor antagonist, 88, 90, 95 Melanocyte-stimulating hormone, 7
Intestine, lipid metabolism, 328 Membrane function
Intrauterine growth retardation, 3, 11, 19, 21, 23 By contractile activity, 75
Involution, 310–311 By feeding and diet, 12–15, 76
Isotopic labelling, 201–202, 206 By inflammation and injury, 75t
By stress, 76–77
Endocrine and autocrine controls, 15–16, 75
K Genetic makeup, 74
Ketogenesis, 222–223, 225, 227–229, 394, Metabolism, 408
396–398, 400 Metabolomics, 467
Methionine, 23, 41–43, 169–170, 176–177, 179,
181–182
L Methionyl-tRNA, 50
Lactate, 375, 377–383, 386, 410 Microflora, 43, 163, 177–178, 182
Laser capture microdissection, 447 Milk
Leptin, 19, 22, 89, 337, 341 Polyunsaturated fatty acids, 241–246
Leucine, 161, 164, 167–168, 172–173, 176, 180–181 Mitochondria, 306, 312, 353, 357–359, 362,
Linoleic acid 364–365
Placental transfer, 23–26 Molecular aspects–differentiation, 278–279
Deficiency, 7, 12–13 mRNA quantitative technique, 438
Dietary requirements, 40–43 Mucin, 164, 169–170
Functional roles, 13–15 Multi-catheterization, 202
Infant nutrition, 34–40 Muscle, 1, 5, 7–12, 14, 16–18, 25
Metabolism, 8–12, 32–34 Muscle hyperplasia, 37, 39–41, 59
Milk composition, 26–29 Muscle hypertrophy, 40, 47
Linolenic acid (alpha) Muscle, lipid metabolism, 335
Deficiency, 7,12–14 Myoblasts, 41–42
Dietary requirements, 40–43 MyoD, 39, 41
Functional roles, 13–23 Myofibres, 37, 40–42, 58
Infant nutrition, 34–40 Myofibril, 356, 359–360
Metabolism, 8–12, 32–34 Myofibrillar protein, 12, 21, 23–24, 26
Milk, 26–29 Myofibrillar proteins, 38, 42, 46, 53, 57, 59
Lipid degradation, 290f, 292 Myogenesis, 39–40
Lipid digestion, preruminants, 324, 327 Myogenic regulatory factors, 24, 27–30
Lipid digestion, ruminants, 326–327 Myogenin, 39, 41
Lipid metabolism, 337, 341 Myosin, 36, 38, 43–44
Lipid synthesis, 13f, 282 Myostatin, 40–41, 58, 60
Lipogenesis, 337, 339, 353, 366, 461, 463 Myotubes, 41–42
Lipolysis, 13, 24–26, 282f, 290–291, 340
Lipopolysaccharide, 6–8, 10–13, 16, 19, 23, 86
Lipoprotein lipase, 13, 20, 282, 288, 330 N
Lipoprotein metabolism, 329–330, 332, 341 Neuropeptide Y, 87
Liver amino acid metabolism, 211 Newborn mammals, 275–276
Liver, lipid metabolism, 325, 341 Nonesterified fatty acid uptake and oxidation,
Long chain polyunsaturated fatty acid See also 460–461
docosahexaenoic acid, arachidonic acid Norepinephrine (NE), 311, 316
Brain, 243–246, 251–252 Northern blotting analysis, 438
Dietary requirements, 253–258 Nuclear magnet resonance spectroscopy, 465
482 Index

Nucleotides, 3, 169, 179–180 R


Nutritional efficiency, 121 Rates, 69, 70
Real-time RT-PCR, 440
Recycling of nitrogen, 198, 207, 214
O Redox status, 157, 181–182
Ontogeny, 399 Regulation
Ornithine, 166, 168, 177, 180 Polyunsaturated fatty acids, 13–23
Oxidation, 157, 160, 165–168, 171, 173, 177 Regulation-fatty acid synthesis, 282, 284–286
Oxidative fuels, 161, 165–166, 168, 182 Regulation-lipolysis, 282t
Oxygen, 5–7, 18 Regulation-triacylglycerol, 288–289t, 290
Ribosomal protein S6 kinase, 26, 51–52,
54–55, 57
Ribosomes, 48–49, 57
P RNA interference, 436
Papillae development, 399 RNase protection assay, 439
Parenteral, 27, 170, 173 Rumen acidosis, 395
Peroxisomal β-oxidation, 342 Ruminant, 405
Peroxisome, 223–224, 227, 230
Phosphoenolpyruvate carboxykinase, 377–378,
382–385
Pig, 353
S
Satellite cells, 39, 42, 47–48, 58–60
Placenta
Sepsis, 91, 95–96, 98, 99
Polyunsaturated fatty acid metabolism, 23
Shivering, 356–357, 359–360, 364
Polyunsaturated fatty acid transfer, 23–26
Short-chain fatty acids, 414
Placenta, 5–11, 14, 19
Skeletal muscle, 14, 19, 21, 70, 73, 83–86,
Placental lactogen, 16, 18
90–100
Polyamines, 168, 170
Slow-twitch (SO) muscle fibres, 38, 44,
Portal absorption, 160
49, 55
Portal-drained viscera, 158–159, 161, 164, 167,
Slow-twitch muscle, 91
169–170, 173, 176
Small intestine, 107–121
Portal-drained visceral amino acid sequestration,
Somatotrophs, 93
198–199
Somatotropic axis, 83–84, 92
Post mortem, 78
Somatotropin, 289
Preadipocytes, 279–280
β3-Adrenergic receptors, 7, 305, 309
Proliferation, 44, 157, 163, 169, 173,
178–182
Proline, 10, 20, 161, 163, 166, 168–169
Protease T
ATP-ubiquitin-dependent, 6, 17, 71, 74 Thermogenesis, 2, 17, 311–312, 316, 318,
Calpain, 71, 74 356–357, 368
Gene expression, 73 Threonine, 169, 171, 177, 181
Lysosomal, 71, 74 Thyroid hormone, 14, 16, 44–45
Matrix metalloprotease, 72 Thyroid hormones, 353, 367
Proteasome, 22, 96–97 Thyroxine, 18
Protein accretion 83, 83–87, 90–92, 98–100 Total parenteral nutrition, 168, 172–173, 177
Protein breakdown, 448, 454 Tracee release method, 452
Protein degradation, 3, 23, 25, 48, 51, 56, 59, Transgenic technique, 435
70, 83, 90–92, 96–100, 163, Translation initiation, 84, 95
172–173, 180 Translation, 26, 28, 49–50, 52, 54–57, 59
Determination of, 7 Transsulfuration, 170, 177, 179, 182
3-methylhistidine, 72 Triacylglycerol synthesis, 21, 288t
Difference methods, 72 Triads, 360
In vitro approaches, 74 Triglyceride synthesis, 461
Isotopic tracer approaches, 73 Triiodothyronine, 18
Protein kinase B, 53–54, 57 Tumor necrosis factor binding protein, 90, 95
Protein microarray, 444 Tumor necrosis factor-α, 83–86, 88–91, 93, 95,
Protein synthesis, 4, 7, 10, 15, 37, 40, 46–57, 59, 97–99
69, 72, 83, 86, 90–95, 100, 157–158, 162, TUNEL method, 446
170, 172–182, 448, 463 Two-dimensional gel electrophoresis, 443
Pyruvate carboxylase, 375, 378 Tyramide signal amplification, 439
Index 483

U VFA metabolism, 392–394


Ubiquitin, 51, 84, 96–97, 99 Visual function
UCP, 356–357 Docosahexaenoic acid, 15–18, 36–40
Uncoupling protein (UCP), 309, 311, 314, 318,
312t, 312f
Urea synthesis, 207 W
White adipose tissue (WAT), 304, 311, 313
Whole-body lipolysis, 460
V Whole-body protein kinetics, 448
Vagus nerve, 89
Very low-density lipoprotein, 329, 330, 332, 334,
341–342

You might also like