You are on page 1of 448

VOLUME NINETY

ADVANCES IN
PARASITOLOGY
Fossil Parasites
SERIES EDITOR
D. ROLLINSON J. R. STOTHARD
Life Sciences Department Department of Parasitology
The Natural History Museum, Liverpool School of Tropical
London, UK Medicine Liverpool, UK
d.rollinson@nhm.ac.uk russell.stothard@lstmed.ac.uk

EDITORIAL BOARD
T. J. C. ANDERSON R. C. OLIVEIRA
Department of Genetics, Texas Centro de Pesquisas Rene Rachou/
Biomedical Research Institute, CPqRR - A FIOCRUZ em Minas
San Antonio, TX, USA Gerais, Rene Rachou Research
Center/CPqRR - The Oswaldo Cruz
M. G. BASÁÑEZ Foundation in the State of Minas
Professor of Neglected Tropical Gerais-Brazil, Brazil
Diseases, Department of Infectious
Disease Epidemiology, Faculty of R. E. SINDEN
Medicine (St Mary’s Campus), Immunology and Infection
Imperial College London, Section, Department of Biological
London, UK Sciences, Sir Alexander Fleming
Building, Imperial College of
Science, Technology and
S. BROOKER Medicine, London, UK
Wellcome Trust Research Fellow
and Professor, London School of D. L. SMITH
Hygiene and Tropical Medicine, Johns Hopkins Malaria Research
Faculty of Infectious and Tropical, Institute & Department of
Diseases, London, UK Epidemiology, Johns Hopkins
Bloomberg School of Public Health,
R. B. GASSER Baltimore, MD, USA
Department of Veterinary Science,
The University of Melbourne, R. C. A. THOMPSON
Parkville, Victoria, Australia Head, WHO Collaborating Centre
for the Molecular Epidemiology
of Parasitic Infections, Principal
N. HALL Investigator, Environmental
School of Biological Sciences, Biotechnology CRC (EBCRC), School
Biosciences Building, University of of Veterinary and Biomedical
Liverpool, Liverpool, UK Sciences, Murdoch University,
Murdoch, WA, Australia
J. KEISER
Head, Helminth Drug X.-N. ZHOU
Development Unit, Department Professor, Director, National
of Medical Parasitology and Institute of Parasitic Diseases,
Infection Biology, Swiss Tropical Chinese Center for Disease Control
and Public Health Institute, Basel, and Prevention, Shanghai, People’s
Switzerland Republic of China
VOLUME NINETY

ADVANCES IN
PARASITOLOGY
Fossil Parasites
Edited by

KENNETH DE BAETS
GeoZentrum Nordbayern,
Friedrich-Alexander-Universit€
at
Erlangen-N€ urnberg, Erlangen, Germany

D. TIMOTHY J. LITTLEWOOD
Department of Life Sciences, Natural History Museum,
London, UK

AMSTERDAM • BOSTON • HEIDELBERG • LONDON


NEW YORK • OXFORD • PARIS • SAN DIEGO
SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO
Academic Press is an imprint of Elsevier
Academic Press is an imprint of Elsevier
125 London Wall, London EC2Y 5AS, UK
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK
225 Wyman Street, Waltham, MA 02451, USA
525 B Street, Suite 1800, San Diego, CA 92101-4495, USA

First edition 2015

Copyright © 2015 Elsevier Ltd. All rights reserved.

No part of this publication may be reproduced or transmitted in any form or by any means,
electronic or mechanical, including photocopying, recording, or any information storage and
retrieval system, without permission in writing from the publisher. Details on how to seek
permission, further information about the Publisher’s permissions policies and our
arrangements with organizations such as the Copyright Clearance Center and the Copyright
Licensing Agency, can be found at our website: www.elsevier.com/permissions.

This book and the individual contributions contained in it are protected under copyright by
the Publisher (other than as may be noted herein).

Notices
Knowledge and best practice in this field are constantly changing. As new research and
experience broaden our understanding, changes in research methods, professional practices,
or medical treatment may become necessary.

Practitioners and researchers must always rely on their own experience and knowledge in
evaluating and using any information, methods, compounds, or experiments described
herein. In using such information or methods they should be mindful of their own safety and
the safety of others, including parties for whom they have a professional responsibility.

To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors,
assume any liability for any injury and/or damage to persons or property as a matter of
products liability, negligence or otherwise, or from any use or operation of any methods,
products, instructions, or ideas contained in the material herein.

ISBN: 978-0-12-804001-0
ISSN: 0065-308X

For information on all Academic Press publications


visit our website at store.elsevier.com
CONTRIBUTORS

Adauto Ara ujo


Fundaç~ao Oswaldo Cruz, Laborat
orio de Paleoparasitologia, Rio de Janeiro, RJ, Brazil
Geoff A. Boxshall
Department of Life Sciences, Natural History Museum, London, UK
Kenneth De Baets
Fachgruppe Pal€aoUmwelt, GeoZentrum Nordbayern, Friedrich-Alexander-Universit€at
Erlangen-N€
urnberg, Erlangen, Germany
Paula Dentzien-Dias
N
ucleo de Oceanografia Geologica, Instituto de Oceanografia, Universidade Federal do Rio
Grande, Rio Grande, Brazil
Philip C.J. Donoghue
School of Earth Sciences, University of Bristol, Life Science Building, Bristol, UK
Stephen K. Donovan
Department of Geology, Naturalis Biodiversity Center, Leiden, The Netherlands
Luiz Fernando Ferreira
Fundaç~ao Oswaldo Cruz, Laborat
orio de Paleoparasitologia, Rio de Janeiro, RJ, Brazil
Joachim T. Haug
Department of Biology II, Functional Morphology Group, University of Munich (LMU),
Planegg-Martinsried, Germany
John Warren Huntley
Department of Geological Sciences, University of Missouri, Columbia, MO, USA
Adiël A. Klompmaker
Florida Museum of Natural History, University of Florida, Gainesville, FL, USA
D. Timothy J. Littlewood
Department of Life Sciences, Natural History Museum, London, UK
Piers D. Mitchell
Department of Archaeology and Anthropology, University of Cambridge, Cambridge,
United Kingdom
Christina Nagler
Department of Biology II, Functional Morphology Group, University of Munich (LMU),
Planegg-Martinsried, Germany
George O. Poinar, Jr.
Department of Integrative Biology, Oregon State University, Corvallis, OR 97331, USA
Karl Reinhard
School of Natural Resources, University of Nebraska, Lincoln, NE, USA

ix j
x Contributors

Paul D. Taylor
Department of Earth Sciences, Natural History Museum, London, UK
Ieva Upeniece
Department of Geology, University of Latvia, Riga, Latvia
Olivier Verneau
CNRS, Centre de Formation et de Recherche sur les Environnements Méditerranéens,
University of Perpignan Via Domitia, Perpignan, France; Unit for Environmental Sciences
and Management, North-West University, Potchefstroom, South Africa
PREFACE

Parasites are ubiquitous, numerous and diverse, and it is widely considered


that this must have been the situation for much of the history of life.
However, understanding the evolutionary history of parasites and their
impact on hosts and ecology through time has relied almost exclusively on
inferences extrapolated from the diversity and nature of parasite species
living today. To a large extent the fossil record of parasites has been over-
looked. Consequently, the role and potential of fossil parasites in the field of
evolutionary biology, and indeed parasitology more generally, has been
difficult to assess. Almost 25 years ago Simon Conway Morris published a
review of fossil parasites in the Journal of Parasitology (82:489–509), consid-
ering their importance in an evolutionary context and providing a com-
prehensive singular resource amenable to parasitologists and palaeontologists
alike. Subsequent reviews on this topic have focused on particular groups of
marine or terrestrial parasites, whilst consideration of fossil parasites more
generally has been invoked through palaeoecological interpretations of
behaviour, coevolution and the nature of trace fossils. This new knowledge
on fossil parasites is mostly scattered in the literature, making it harder to get
an overview of their usefulness and potential for illuminating or framing
their evolutionary history. Newly discovered fossils, the development of
sensitive techniques that enable fossils to be better visualized, manipulated
digitally, and analyzed chemically, and the need to integrate palae-
ontological data into time-calibrated phylogenies, have provided a wealth of
new opportunities to consider fossil parasites again. The combination of
frustrations balanced by advances and opportunities provided the impetus for
this volume. Fortuitously there are also currently many active researchers
considering fossil parasites and their role in better understanding the nature
of parasitism in evolution, disease and ecology. The sheer diversity of
approaches and innovations further merits a renewed perspective, not least
to demonstrate how much more integrative the study of parasitology has
become.
This volume aims to demonstrate that direct and indirect evidence from
the fossil record can be crucial in a variety of ways, providing empirical,
complementary and supplementary evidence in testing and developing
hypotheses on the nature of parasite diversity, phylogeny and host associa-
tions through space and time. In the first chapter, Kenneth De Baets and
Tim Littlewood review recent discoveries of fossil parasites and vectors and

xi j
xii Preface

consider them in their phylogenetic context, highlighting emerging tech-


niques and pathways for further progress. The fossil record of soft-bodied
parasites is understandably poor, but George Poinar’s contribution and that
of De Baets and co-workers, each explore how the fossil record of parasitic
helminths might be richer, or at least ‘less poor’, than that of their free-living
relatives. Poinar concisely discusses the evolutionary history of nematode
parasites of invertebrates, vertebrates and plants based on fossil remains in
amber, rock and coprolites ranging from the Palaeozoic to the Holocene
drawing from his long personal research on this topic. De Baets and co-
workers review the fossil record of parasitic flatworms for the first time and
illustrate how fossil and geological evidence may be used to calibrate
molecular clocks. Christina Nagler and Joachim Haug comprehensively
review the fossil record of insects, which includes some of the most
important terrestrial parasites, parasitoids and vectors today.
In reviewing the fossil record of crustacean parasites and hosts, Adiël
Klompmaker and Geoff Boxshall demonstrate that characteristic pathologies
and traces in hosts with fossilizable skeletons can provide a less patchy record
of parasitism compared with the body fossil record, thus demonstrating the
need to better understand and detect such traces. Pathologies in several other
groups of hosts with hard parts can be linked with parasitism, particularly in
bivalve molluscs, echinoderms and colonial organisms. John Huntley and
Kenneth De Baets discuss the indirect record of trematode flatworms in
bivalve shells. Stephen Donovan adds a new take on evidence for ecto-
parasitism on echinoderms, a group with a remarkable fossil record and one
that has captured numerous lesions, traces, burrows and attachments since
their appearance in deep time. Paul Taylor reviews for the first time the fossil
evidence for parasitism in fossil colonial organisms including sponges,
bryozoans, corals and graptolites and reveals a number of questions and
conundrums. In spite of this wealth of damage and evidence of association,
the identity of the perpetrators and the true nature of their associations
remain open to interpretation and speculation. For these predominantly
marine taxa, it seems that modern studies of parasite ecology in marine
environments may yet reveal more about ancient associations, and the
nature of traces caused by species (likely extinct) in other organisms also
extinct. The systematic review of fossil evidence at least provides focus for a
renewed search amongst modern marine assemblages.
Last but not least, the field of palaeoparasitology is drawn considerably
from techniques and methods introduced and still being developed rapidly
in the study of human parasites at archaeological sites. Drawing from a
Preface xiii

lifetime of experience Adauto Ara ujo and co-workers illustrate the impor-
tance of ancient parasitic remains and review the methods to study human
parasites. Considering the links between human health, medical practice and
parasites, Piers Mitchell provides the final chapter with a focus specifically on
Medieval European parasite remains. Fascinating personal natural histories of
the rich, famous and simple peasantry provide insights into archaeology as
well as anthropology. The constraints of parasitic infections on lifestyle,
sanitation and medical treatment of humans in early Europe resonates with
much of modern parasitology today and brings into focus our own ongoing
struggles with infection and parasite-mediated disease worldwide.
We take this opportunity to thank all of the authors for their con-
tributions and to the editorial team for their support in making this volume
possible.

Kenneth De Baets and D. Timothy J. Littlewood


October 2015
CHAPTER ONE

The Importance of Fossils in


Understanding the Evolution of
Parasites and Their Vectors
Kenneth De Baets*, 1, D. Timothy J. Littlewoodx, 1
*Fachgruppe Pal€aoUmwelt, GeoZentrum Nordbayern, Friedrich-Alexander-Universit€at
Erlangen-N€urnberg, Erlangen, Germany
x
Department of Life Sciences, Natural History Museum, London, UK
1
Corresponding authors: E-mail: kenneth.debaets@gmail.com; t.littlewood@nhm.ac.uk

Contents
1. Introduction 2
2. Techniques for Ancient Parasite Discovery 4
2.1 Thin sections and computed tomography 5
2.2 Ancient biomolecules 6
2.2.1 Ancient DNA 6
2.2.2 Palaeoproteomics 7
3. The Parasite Fossil Record 8
3.1 Body fossils 14
3.2 Trace fossils and pathologies 18
3.3 Coprolites 22
4. Molecular Perspectives on Parasite Phylogeny and Evolution 26
4.1 Molecular clocks 29
4.2 HGT and ‘parasitic DNA’ 34
5. Future Perspectives 35
Acknowledgements 36
References 36

Abstract
Knowledge concerning the diversity of parasitism and its reach across our current
understanding of the tree of life has benefitted considerably from novel molecular
phylogenetic methods. However, the timing of events and the resolution of the nature
of the intimate relationships between parasites and their hosts in deep time remain
problematic. Despite its vagaries, the fossil record provides the only direct evidence
of parasites and parasitism in the fossil record of extant and extinct lineages. Here,
we demonstrate the potential of the fossil record and other lines of geological evi-
dence to calibrate the origin and evolution of parasitism by combining different kinds
of dating evidence with novel molecular clock methodologies. Other novel methods
promise to provide additional evidence for the presence or the life habit of pathogens
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.07.001 All rights reserved. 1
2 Kenneth De Baets and D. Timothy J. Littlewood

and their vectors, including the discovery and analysis of ancient DNA and other bio-
molecules, as well as computed tomographic methods.

1. INTRODUCTION
Parasitism is one of the most successful modes of life, as evidenced by
its convergent appearance in numerous lineages and its sheer absolute and
relative abundance among extant biodiversity (Poulin and Morand, 2000).
Antagonistic interactions, in the form of arms races between parasites and
their hosts, have been considered important drivers of evolution (Zaman
et al., 2014) and might also have contributed to the origin of sexual repro-
duction (Mostowy and Engelst€adter, 2012). Because parasitism also has an
obvious societal importance with many parasitic taxa being of significant
biomedical, veterinary or economic importance (Bush et al., 2001), it is
here that most of the research effort is focused. This focus is narrow and fails
to provide the wider evolutionary picture or an appreciation of the influence
of parasitism on, and as part of, biodiversity. Indeed, despite their importance
and ubiquity, the evolutionary history of parasites is still poorly known, a
phenomenon not helped by their inadequate, or rather inadequately
explored, fossil record (Littlewood and Donovan, 2003).
Establishing time-calibrated evolutionary frameworks to test the origins
and radiations of parasites in parallel with studies on environmental param-
eters, or the degree of coevolution between parasites and hosts, is a difficult
but as yet a largely unexplored means by which ancient associations may be
revealed. Parasitologists have often resorted to more circular lines of evi-
dence, such as extrapolating from current host associations or distributions
to put time constraints on the origins and evolution of parasites. For instance,
where extant hosteparasite associations appear to be combinations of early
divergent hosts and early divergent parasites, it is tempting and compelling to
assume a long and ancient association; for example, early divergent gyroco-
tylidean cestodes found only parasitizing early divergent ‘primitive’ holoce-
phalan fishes (Xylander, 2001). In these cases, when the timing of a host’s
divergence can be estimated from molecular or preferably fossil evidence,
a calibration point for the parasite’s association also appears tractable, at least
as a working hypothesis. Assumptions of cophylogeny are common but
bring their own suites of problems, not in the least because of the traps set
by multiple assumptions (Page, 2003). To reveal coevolutionary patterns,
phylogenies of hosts and parasites need to be untangled to better understand
The Importance of Fossils in the Evolution of Parasites 3

historical relationships, but the task is complex. Accurate estimates of histor-


ical events such as co-divergence, duplication or loss of an association require
complex mathematics and computationally demanding algorithms, and any
estimate is contingent upon adequate sampling (Charleston and Perkins,
2006). Usually such sampling relies on phylogenies determined from extant
organisms and pays little heed to loss of lineages through extinction. Whilst
these studies can be profitable, direct evidence from the fossil record remains
the most compelling evidence for past historical and deep evolutionary
associations, as well as extinctions. Palaeontological data could also have a
bearing on testing of how parasiteehost associations respond to environ-
mental changes across longer time-scales and to what extent parasites could
be prone to (co)extinction (Dunn et al., 2009).
The past decades have seen a wealth of new discoveries, ranging from
exceptionally preserved parasites and eggs assignable to modern (even family
level) lineages (Cressey and Boxshall, 1989; Da Silva et al., 2014; Hugot
et al., 2014), to characteristic traces of preserved biomolecules in host
remains (Dittmar, 2009; Greenwalt et al., 2013; Wood et al., 2013b). Of
particular note have been advances in X-ray, ion, electron and laser-beam
techniques, serial grinding/imaging techniques and magnetic resonance
tomography characterizing fine structures, textures and underlying chemis-
tries (Mietchen et al., 2008; Sutton, 2008; Schiffbauer and Xiao, 2011;
Dunlop et al., 2012; Cunningham et al., 2014a,b; Sutton et al., 2014). Addi-
tionally, advances in mass spectrometry have allowed the detection and
characterization of amino acid traces, particularly collagen within bone, to
a remarkable level of detail and resolution (Cappellini et al., 2014). Such
techniques open up the prospect of detecting traces of parasites and para-
sitism more frequently and revealing key systematic features and morpho-
logical characters indicative of a parasitic way of life.
Another facet of palaeoparasitology, although perhaps not widely
considered as such, is the study of horizontally (laterally) transferred
DNA including transposable elements, where DNA from one organism
can be detected buried within the genome of another. These so-called
‘genomic fossils’ offer clues as to the origins and nature of ancient associa-
tions (Gilbert and Feschotte, 2010; Gilbert et al., 2010; Katzourakis and
Gifford, 2010; Katzourakis, 2013; Koutsovoulos et al., 2014). Indeed,
Ford Doolittle considers horizontally transferred DNA fragments as
‘basically parasites’ (p. 8; Gitschier, 2015) that have been parts of their
host genomes for a considerable length of time. The rise in genome studies
has provided ever-increasing evidence for horizontal gene transfers, HGTs
4 Kenneth De Baets and D. Timothy J. Littlewood

(e.g. see Scholl et al., 2003), although such events detected among Metazoa
appear to be more common in some groups (e.g. bdelloid rotifers:
Gladyshev et al., 2008) than in others. Whereas palaeontology looks
towards the earth’s fossil and subfossil record for ancient biotic interactions,
it is clear that genomes may also be gleaned for evidence of relictual genetic
elements of nonhost (“parasitic”) origin.
Regardless of approach, time points gathered directly from fossils or
inferred from calibrated phylogenies remain critical in understanding
when, where and to some extent how hosteparasite interactions took place
and how they might respond in the future; for example, the exchange of
genes from parasitic to host plants of the genus Plantago has been shown
to be a result of their direct physical contact with one another (Mower
et al., 2004). Parasitic plants offer a particularly rich resource for understand-
ing HGT (Davis and Xi, 2015).
Morphologically based classifications of parasites have proved chal-
lenging due to frequent apparent simplifications, convergence or specializa-
tions in their morphology that make homology assessment difficult.
However, novel molecular methods, used with caution, may form the basis
for more robust phylogenetic assignments of extant and subfossil parasitic
remains, and thus more comprehensive understanding of the origin and
evolution of parasitism within single lineages (Near, 2002; Lockyer et al.,
2003; Littlewood, 2011; Wood et al., 2013b; Hartikainen et al., 2014; Sum-
mers and Rouse, 2014; Blaxter and Koutsovoulos, 2015; Littlewood and
Waeschenbach, 2015; Okamura and Gruhl, 2015). Here we provide an
updated perspective on the merits, further possibilities and frustrations
associated with using the fossil record in constraining the origins and evolu-
tion of parasitism. We highlight novel methods, which make it possible to
more fully exploit information buried in fossil or genomic sequences of their
hosts.

2. TECHNIQUES FOR ANCIENT PARASITE DISCOVERY


Various destructive methods (thin sections, rehydratation or resedi-
mentation techniques, and grinding tomography) and nondestructive
methods (e.g. phase contrast synchrotron or microcomputed tomography,
mCT) are particularly relevant to discovering and characterizing the
morphology of parasitic remains or host responses in older (fossil) samples.
Increasingly these can be supplemented with analyses of ancient
The Importance of Fossils in the Evolution of Parasites 5

biomolecules, including ancient DNA (aDNA) in younger (archaeological)


samples, or analyses of more resistant biomolecules in older, exceptionally
preserved fossil samples (Briggs and Summons, 2014).

2.1 Thin sections and computed tomography


Various studies have demonstrated the merit of conventional preparation
and imaging methods used for archaeological samples (Araujo et al., 2008;
Ara ujo et al., 2015) to provide evidence for helminth eggs or other parasitic
remains in fossil coprolites (Poinar and Boucot, 2006; Dentzien-Dias et al.,
2013; Da Silva et al., 2014). Such investigations typically rely on thin
sectioning and various dissolution/rehydration/resedimentation methods,
which destroy the original 3D structure and/or association of the parasites
and may lose less-resistant parasite remains (Dufour and Le Bailly, 2013;
Wood et al., 2013b; Ara ujo et al., 2015).
Computed tomography (Sutton et al., 2014), whereby 3D reconstruc-
tions of serially sectioned material are rendered and visualized by computer,
is developing rapidly as a means by which high resolution differentiation can
be achieved from the fossilized remains of small and even soft-bodied organ-
isms. However, the resolution of parasites and evidence of parasitism often
remains a serendipitous by-product of investigating fossil microstructures.
One such case has recently revealed the remarkable discovery of a putative
fossil pentastomid found in association with its ostracod host entombed for
425 Ma (Siveter et al., 2015). Not only did this discovery reveal the first fos-
sil occurrence of an adult pentastomid but also its host association, which had
been the subject of some considerable speculation in the literature
(Waloszek et al., 2005). Today, pentastomids mostly parasitize terrestrial ver-
tebrates exclusively as endoparasites, and although extinct representatives
have been identified in Cambrian marine sediments at a time before these
terrestrial hosts existed, it has been suggested that these forms were parasitic
on marine vertebrates (Sanders and Lee, 2010). The new Silurian record by
Siveter et al. (2015) not only reveals a new host association but also shows
the parasite as ectoparasitic. Unfortunately, because of the destructive nature
of the grinding technique, the fossil now exists only as an image. Although
many reports for terrestrial parasites or vectors come from amber (Poinar,
2014a), computed tomography has only been rarely used to test such
assertions (Dittmar et al., 2011; Dunlop et al., 2012), which have been
largely based on microscopic methods. Phase contrast tomography and other
tomographic methods are however ideally suited to characterize fossils in
three dimensions (even in nontransparent amber) and corroborate
6 Kenneth De Baets and D. Timothy J. Littlewood

parasite/vector interpretations as it has been used for phoretic associations


(Dunlop et al., 2012; Penney et al., 2012).

2.2 Ancient biomolecules


Novel techniques make it possible to identify biomolecules (Greenwalt
et al., 2013; Briggs and Summons, 2014; Yao et al., 2014) in ancient remains
and samples. The most familiar one is probably aDNA, which has been
applied to various parasitic (sub)fossils (Dittmar, 2009; Dittmar et al.,
2011; Wood et al., 2013b), although it can only be detected and character-
ized reliably in exceptionally well-preserved material up to 1 Ma (Hebsgaard
et al., 2005; Briggs and Summons, 2014). Large and fragile molecules such as
DNA cannot survive fossilization (Briggs and Summons, 2014), but other
complex organic structures such as iron-stabilized haem, can survive for
longer, at least under certain conditions (Briggs, 2013; Greenwalt et al.,
2013; Yao et al., 2014), opening up the prospects for protein-based detec-
tion methods (Cappellini et al., 2014). This can be used as an extension to
the analysis of gut contents of isolated parasite or vector remains to get an
indication of the trophic targets and, therefore, the possible identity of the
hosts. Preservation of feather remnants in a louse specimen’s foregut, for
example, confirmed its association to a waterbird ectoparasite (Wappler
et al., 2004).
Advances in chemical analysis at the nanoscale, and as applied to fossils,
opens up a whole new world in revealing ancient colours, pigments,
microbiomes, and the hidden remnants of soft-bodied organisms (Briggs
and Summons, 2014; Bertazzo et al., 2015; Vinther, 2015). A better
understanding of taphonomy (Briggs, 2013), as applied to parasitic groups,
may provide the tools to improve and apply these techniques in recognizing
parasites and their influence on hosts in the fossil record. Certainly there
seems room for developing and applying these techniques to subfossil
(unmineralized) remains where eggs are found intact without DNA and
without clear morphological identity (see also Linseele et al., 2013 and
Cano et al., 2014).

2.2.1 Ancient DNA


Earlier aDNA studies focused on PCR to amplify specific short gene
sequences targeted for particular parasite groups or species from single sam-
ples (reviewed in Dittmar, 2009; Dittmar et al., 2011; Dittmar, 2014). Novel
approaches (Wood et al., 2013b) focus on the amplification of total DNA of
whole samples followed by the application of next generation sequencing
The Importance of Fossils in the Evolution of Parasites 7

platforms for shotgun sequencing. Subsequent bioinformatic interrogation


of the data for various groups of parasites has been established in the field
of metagenomics and environmental samples (Dittmar, 2009; Wood et al.,
2013b). In spite of remarkable progress, aDNA techniques will always be
constrained by the rapid deterioration of DNA.

2.2.2 Palaeoproteomics
DNA is not believed to survive in sufficiently long lengths for sequencing
over longer geological timescales (>1 Ma) (Hebsgaard et al., 2005; Briggs
and Summons, 2014) and is rarely found well preserved. However, the
study of fossilized bone, which has been crudely defined as a composite
of collagen (protein) and hydroxyapatite (mineral) (see Hill et al., 2015),
has shown that, for vertebrates at least, palaeoproteomics can be a
rewarding insight into ancient proteins, providing evidence for phyloge-
netics and an understanding of bone biochemistry (Wadsworth and
Buckley, 2014). Collagen, in particular, has been isolated from vertebrate
fossils of considerable age, including an 80 million year (my)-old Campa-
nian hadrosaur, Brachylophosaurus canadensis (Schweitzer et al., 2009) and a
68-my-old Tyrannosaurus rex (Asara et al., 2007); some of these studies have
attracted some criticism (Pevzner et al., 2008). The characterization of the
constituent peptides of fossil bone proteins, requiring mass spectrometry,
suggests that collagen (the most abundant protein) can survive up to
340 ky at 20  C, and the second most abundant protein, osteocalcin, can
persist for w45 ky; see Ostrom et al. (2000), Hofreiter et al. (2012) and
Collins et al. (2000). There are many other bone proteins that can be iso-
lated and identified depending on the nature of preservation and the age of
the fossil (Cappellini et al., 2012), but recent studies focusing on collagen
have provided opportunities to push timescales back further in the charac-
terization of ancient biomolecules useful for phylogenetics (Welker et al.,
2015). Recently, even putative erythrocyte remains were reported in dino-
saur bones (Bertazzo et al., 2015).
The application of palaeoproteomics more broadly to parasites or to
other fossil remains, in the hope of finding evidence for parasitism, is in its
infancy, but analysis of more resistant biomolecules than DNA might
make it possible to test other hypotheses associated with parasites or host-
related biomolecules. Certainly, the prospect of verifying the presence of
porphyrins in fossilized haematophagous insects (Greenwalt et al., 2013;
Yao et al., 2014) seems a tractable goal.
8 Kenneth De Baets and D. Timothy J. Littlewood

3. THE PARASITE FOSSIL RECORD


The fossil record of parasites or other pathogens is usually poor
because they often reside within their hosts or vectors, may go through
some life cycle stages away from their hosts, are small, and lack hard tissues
(Conway Morris, 1981; Baumiller and Gahn, 2002; Labandeira, 2002;
Littlewood and Donovan, 2003; De Baets et al., 2011). This is particularly
true not only for viruses, bacteria and protozoa (Poinar, 2014a), but also
for various soft-bodied metazoan parasites such as helminths (Littlewood
and Donovan, 2003; De Baets et al., 2015a; Huntley and De Baets, 2015;
Poinar, 2015a) or weakly sclerotized arthropods (Cressey and Boxshall,
1989; Castellani et al., 2011; Klompmaker and Boxshall, 2015; Nagler
and Haug, 2015). Body fossils of parasites are rare and usually restricted to
sites of exceptional fossil preservation (KonservateLagerst€atten), particularly
those still associated with the remains of their hosts providing direct evidence
of parasitism. Characteristic traces or pathologies in the skeletons of their
hosts are more common and can be traced more continuously over longer
timescales, but the taxonomic affinity of the culprits is not always easy to
disentangle (Donovan, 2015; Huntley and De Baets, 2015; Klompmaker
and Boxshall, 2015; Taylor, 2015). The fossil record of parasitism has
been reviewed on various occasions (Conway Morris, 1981; Boucot,
1990; Littlewood and Donovan, 2003; Boucot and Poinar, 2010; Dittmar,
2010) with some others focusing particularly on marine parasites (Baumiller
and Gahn, 2002; Rouse, 2005a) or terrestrial pathogens (Labandeira, 2002;
Poinar, 2014a). The fossil record can also provide direct evidence for pres-
ence of parasiteehost associations e some of which might now be extinct, as
well as the impact of parasitism on their hosts in the geological past. We will
discuss the different sources of fossil evidence, their merits, limitations and
associated frustrations.
Major important fossil discoveries of metazoan parasites and pathogens
distributed by vectors underpinning our understanding of the ancient history
of hosteparasite associations are detailed in Table 1. The table reveals
considerable diversity of hosts and parasites, and hosteparasite interactions.
New finds also highlight the potential of finding evidence in coprolites for
parasitic remains in both marine and terrestrial realms, and amber as sources
for haematophagous vectors. Fossil pentastomids were long thought to be
rare and restricted to the CambrianeOrdovician sites with Orsten preserva-
tion, but new discoveries highlight they might be more common than
Table 1 Ancient history of hosteparasite associations
Higher taxon Taxon Fossil evidence Age Source Host References Environment
Bacteria
Spirochaetes Palaeoborrelia Direct: spirochaete- Miocene Amber Amblyomma sp. Poinar (2014b) T
dominicana like cells
Rickettsiales Palaeorickettsia Direct: rickettsial-like Early Cretaceous Amber Cornupalpatum Poinar (2015b) T
protera cells burmanicum

Eukaryota
Metamonada Parabasalia: Indirect: lesions Cretaceous Skeletal Tyrannosaurus rex Wolff et al. (2009) T
Trichomonadida deformation
Amoebozoa? Entamoebites antiquus Direct: cyst Cretaceous Coprolite Archosaur Poinar and Boucot T
(?dinosaur) (2006)

Phylum Apicomplexa
Gregarinasina Primigregarina Direct: in amber CretaceouseMiocene Amber Cockroaches Poinar (2012) T
burmanica
Coccidia Cryptosporidium, Direct: aDNA in Holocene Coprolite Moas Wood et al. (2013b) T
Eimeriorina coprolite
Eimeria lobatoi Indirect: oocysts Holocene Coprolite Deer Ferreira et al. (1992)
Archeococcidia antiquus; Indirect: oocysts Pleistocene Coprolite Ground sloth Schmidt et al. (1992)
A. nothrotheriopsae
Haemospororida Plasmodium dominicana Direct: erythocytes Miocene Amber Culex malariager Poinar (2005b) T
Vetufebrus ovatus Direct: erythocytes Miocene Amber Enischnomyia stegosoma Poinar (2011b) T
Paleohaemoproteus Direct: erythocytes Early Cretaceous Amber Proticulicoides sp. Poinar and Telford T
burmacis (2005)

Phylum Euglenozoa
Trypanosomatida Palaeotrypanosoma Direct: erythocytes Early Cretaceous Amber Leptoconops nosopheris Poinar (2008a) T
burmanicus
Trypanosoma Direct: erythocytes Miocene Amber Triatoma dominicana Poinar (2005b) T
antiquus
(Continued)
Table 1 Ancient history of hosteparasite associationsdcont'd
Higher taxon Taxon Fossil evidence Age Source Host References Environment
Palaeoleishmania Direct: erythocytes Early Cretaceous Amber Palaeomyia burmitis Poinar and Poinar T
proterus (2004a,b)
Paleoleishmania Direct: erythocytes Miocene Amber Lutzomyia adiketis Poinar (2008b) T
neotropicum

Phylum Arthropoda
Copepoda ? Indirect: exocysts Jurassic Skeletal Crinoids, echinoids Mercier (1936),
deformation Radwa nska and
Radwa nska (2005),
Radwanska and
Poirot (2010)
Kabatarina pattersoni Direct: Early Cretaceous Calcareous Fish Cressey and Patterson M
nodules (1973), Cressey and
Boxshall (1989)
Isopoda ?Bopyridae Indirect: swellings ?Early Jurassic; Middle Skeletal Crustacea Klompmaker et al. M
JurassiceRecent deformation (2014),
Klompmaker and
Boxshall (2015)
Pentastomida 5 genera; 10 species Direct: phosphatized Cambrian Calcareous ? early chordates; Waloszek and M€ uller M
remains isolated eOrdovician; nodules ostracod (1994), Waloszek
from host Silurian? et al. (1994),
Waloszek et al.
(2005a,b),
Castellani et al.
(2011), Siveter et al.
(2015)
Acari Cornupalpatum Direct: larva stage Early Cretaceous Amber ? Poinar and Brown T
burmanicum (2003)
(Ixodidae)
Compluriscutula Direct: larva stage Early Cretaceous Amber ? Poinar and Buckley T
vetulum (Ixodidae) (2008)
Thecostraca ?Ascothoracida Indirect: borings Cretaceous Skeletal Echinoidea Madsen and Wolff M
deformation (1965)
?Ascothoracida Indirect: borings Cretaceous Skeletal Octocorallia Voigt (1959), (1967) M
deformation

Insecta
Siphonaptera s.l. Pseudoculidae Direct: isolated from JurassiceCretaceous Lacustrine ?pterosaurs, dinosaurs Gao et al. (2012), T
host deposits and/or small Huang et al. (2012),
mammals Gao et al. (2013),
Huang et al. (2013),
Gao et al. (2014),
Huang (2014)
Siphonaptera s.s. Paleopsylla: 4 species Direct: isolated from Eocene Amber ?mammals Dampf (1911), T
host Beaucournu and
Wunderlich
(2001),
Beaucournu (2003)
Phthiraptera Megamenopon rasnitsyni Direct: isolated from Eocene Lacustrine ?water birds Wappler et al. (2004) F
host deposits
? Indirect: nits Eocene Baltic amber Mammals Voigt (1952) T
Diptera Qiya jurassica Direct: parasitic larvae Jurassic Lacustrine ?salamanders Chen et al. (2014) F
(Athericidae) deposits

Phylum Nematoda
Ascaridida Ascarites rufferi Indirect: eggs Upper Triassic Coprolite Cynodont Da Silva et al. (2014) T
(Ascarididae)
Ascarites gerus Direct: egg with Early Cretaceous Coprolite Archosaur Poinar and Boucot T
(Ascarididae) developing juvenile (2006)
Toxocara canis Direct: eggs Pleistocene Cave deposits Canid Bouchet et al. (2003)
(Ascarididae) (?Crocuta spelaea)
Oxyurida Paleoxyuris cockburni Indirect: eggs Upper Triassic Coprolite Cynodont Hugot et al. (2014) T
(Heteroxyne-
matidae)
(Continued)
Table 1 Ancient history of hosteparasite associationsdcont'd
Higher taxon Taxon Fossil evidence Age Source Host References Environment
Mermithida Cretacimermis libani Direct Early Cretaceous Amber Midge Poinar et al. (1994) T
(Mermithidae) (Chironomidae:
Diptera)
?Enoplida Palaeonema phyticum Direct Early Devonian Silicified plant Early land plant Poinar et al. (2008) T
(Palaeonematidae) material

Phylum Nematomorpha
Chordodidae Cretachordodes burmitis Direct Early Cretaceous Amber Poinar and Buckley T
(2006)

Phylum Platyhelminthes
Cestoda Permian Direct: eggs with Coprolite Sharks Dentzien-Dias et al. M
developing embryo (2013), De Baets
et al. (2015a)
Trematoda Digenites proterus Early Cretaceous Indirect: eggs Coprolite Archosaurs Poinar and Boucot T
(2006)
Dicrocoelidae Pleistocene Indirect: eggs Coprolite ?bear Jouy-Avantin et al. T
(1999)
?Gymnophallidae Early Eocene Indirect: characteristic Skeletal Ruiz and Lindberg M
pits in bivalve shells deformation (1989), Todd and
Harper (2011),
Huntley and De
Baets (2015)
?Monogenea Middle Devonian Direct: attachment Fine-grained Placoderms, Upeniece (2001,
structure sediments acanthodians 2011), De Baets
et al. (2015a)
Phylum Acanthocephala
? ? Holocene Indirect: eggs Coprolite Mammals (canids, Fry and Hall (1969), T
humans) Noronha et al.
(1994)

Phylum Annelida
Myzostomida ? Carboniferous Indirect: skeletal Crinoids Welch (1976), M
eJurassic deformations (galls) Hess (2010)
Clitellata ?Hirudinae Triassic Indirect: cocoons Freshwater ? Manum et al. (1991), F
deposits Bomfleur et al.
(2012), Parry et al.
(2014)

Phylum Mollusca
Eulimidae Eulima Upper Cretaceous Direct: isolated shells Marine deposit ? Sohl (1964) M
? Upper Cretaceous Indirect: trace on Skeletal Echinoids Neumann and M
echinoid hosts deformation Wisshak (2009)
Coriaphyllidae Coralliophila (Timothia) Eocene Direct: isolated shells Marine deposit ? Dockery (1980) M
aldrichi
Leptoconchus: 2 species; OligoceneeMiocene Direct: shells associated Skeletal Corals (Cladocera, Lozouet and Renard M
Coralliophila: with coral host deformation Thegioastraea, (1998)
1 species; Galeropsis: Pocillopora)
1 species
Unionidae Unio, Anodonta Quaternary Direct: glochidium Freshwater ?fish Brodniewicz (1968) F
larvae deposits
14 Kenneth De Baets and D. Timothy J. Littlewood

expected even during different ages. Such fossil Lagerst€atte should therefore
be systematically screened for parasites. Further work needs to be done on
establishing characteristic trace fossils and pathologies, particularly in verte-
brates, back in time.

3.1 Body fossils


Body fossils are usually defined as remains or representations of the whole or
actual parts of organisms (Goldring 1999; Foote and Miller, 2007). In rare
cases, parasites are so well and completely preserved that they can be accu-
rately assigned to modern taxa (Cressey and Boxshall, 1989; Siveter et al.,
2015). Often, only sclerotized attachment organs, eggs or cysts are known,
which can still be characterized sufficiently to assign them to particular taxa
(De Baets et al., 2015a). In other cases, body fossils of nonparasitic life stages
of extant parasites hint that their parasitic life stages might potentially also be
present (e.g. parasitic stages of arthropods or unionids), but this can only be
confirmed by future discoveries (Skawina and Dzik, 2011; Nagler and Haug,
2015). All extant members of Unionida have parasitic larvae, which suggests
that this is a synapomorphic trait for this group, which goes as far back as the
middle Triassic (Skawina and Dzik, 2011). Unfortunately, the oldest fossil
glochidia cannot be used to test this hypothesis as they are only found as
late as the Pleistocene (Brodniewicz, 1968). In exceptional cases, parasite
body fossils are still associated with host remains, while in other cases they
are isolated, making it difficult to identify the hosts or even decide if these
forms were parasitic or not. Establishing them as parasites is usually achieved
by comparing them with the morphology of modern relatives or analogues,
but in some cases it can only be corroborated by finding the parasite in situ
parasitizing its host. Labandeira (2002), for example, claims that Cambrian
tardigrades (Muller et al., 1995) have certain resemblances to extant parasitic
forms, which indicates at least a possible phylogenetic relationship and
potentially a mode of life, but this still needs to be further corroborated.
The body fossil record is restricted to sites of exceptional preservation.
One of the most important types of preservation is found in fossil ambers
which have preserved the remains of unicellular pathogens from a variety
of terrestrial parasitic arthropods (Nagler and Haug, 2015), to nematodes
exiting their dying hosts (Poinar, 2014a, 2015a). Other important Konservat
Lagerst€atten for the preservation of parasites include phosphatized remains of
arthropods (Cressey and Patterson, 1973; Cressey and Boxshall, 1989; Maas
and Waloszek, 2001; Maas et al., 2006) in carbonate nodules (the so-called
Orsten preservation), silicified nematodes (Poinar et al., 2008) in
The Importance of Fossils in the Evolution of Parasites 15

hydrothermal vent environments (the so-called Rhynie Chert preservation),


fine-grained lacustrine or marine deposits where the remains were quickly
buried and/or anoxic environments contributed to fossilization. The latter
includes oil shales deposited not only in maars or other lacustrine environ-
ments (Wappler et al., 2004; Hughes et al., 2010; Greenwalt et al., 2013),
but also in low energetic, marginal marine or deeper marine environments.
In some cases, ectoparasites are still attached or associated with their hosts,
while endoparasites are still found in situ within their hosts or escaping their
dead hosts. Cysts, eggs including those with developing embryos, larvae
(Ferreira et al., 1993) or remains of juveniles, are sometimes also recovered
from coprolites, one of the many types of trace fossils (see Section 3.2).
Protozoan parasites or other pathogens like viruses and bacteria are usu-
ally hard to verify in the fossil record (Frías et al., 2013; Poinar, 2014a), but
are known to be transmitted by various vectors, which might themselves
fossilize. Most common vectors are arthropods which feed on blood,
although forms feeding on feathers or hairs might also be involved. Direct
fossil evidence for vector feeding behaviour is even rarer than parasitism
(Table 2). Possible exceptions include the recent discovery of haemoglo-
bin-derived porphyrins in the stomach content of an Eocene mosquito
(Greenwalt et al., 2013) as well as higher Fe contents in fossilized true
bugs, Hemiptera (Yao et al., 2014). Direct evidence for vector behaviour
of haematophagous taxa has been restricted so far to amber deposits (Poinar
and Poinar, 2004b; Poinar, 2005a,b,c; Poinar and Telford, 2005; Poinar,
2011b, 2014b, 2015b). Some unicellular pathogens have also been reported
from fossil and subfossil archosaur coprolites, although microscopic taxa are
harder to identify in older occurrences than in more recent occurrences
where aDNA is available (Poinar and Boucot, 2006; Wood et al., 2013b).
Our ability to identify parasitic and haematophagous insects is mostly based
on their possible taxonomic affinities and morphological adaptions of their
mouth parts or other structures similar to extant taxa (Lukashevich and
Mostovski, 2003; Greenwalt et al., 2013; Pe~ nalver and Pérez-De la Fuente,
2014; Nagler and Haug, 2015). Such morphological adaptations have even
been used to suggest ectoparasitic behaviour in lineages where larvae are no
longer parasitic (Chen et al., 2014).
Modern-type lice can be traced back to the Eocene (Wappler et al., 2004),
while earlier reports are most likely erroneous determinations of mites
(Dalgleish et al., 2006), so that Phthiraptera, which are mainly specialized
on mammals, probably evolved later than did fleas (Nagler and Haug,
2015). Fossils of modern-type fleas can be traced back to the Eocene at least
16
Table 2 Body fossils of haematophagous vectors
Taxonomy Species Haematophagy Parasite Source References
Arachnida
Ixodidae Cornupalpatum burmanicum Indirect Rickettsial-like cells Burmese amber Poinar (2015b)
Amblyomma sp. Indirect Spirochaete-like cells Dominican amber Poinar (2014b)

Diptera
Ceratopogonidae Proticulicoides sp. Indirect Plasmodiidae Burmese amber Poinar and Telford (2005),
Boucot and Poinar (2010)
Leptoconops nosopheris Indirect Trypanosomatidae Burmese amber Poinar (2008a)
Culicidae Culiseta sp. Direct Fe ? Kishenehn Formation Greenwalt et al. (2013)
concentrations,

Kenneth De Baets and D. Timothy J. Littlewood


porphyrins
Culex malariager Indirect Plasmodiidae Dominican amber Poinar (2005a,b)
Phlebotomidae Palaeomyia burmitis Indirect Trypanosomatidae Burmese amber Poinar and Poinar (2004a,b)
Lutzomyia adiketis Indirect Trypanosomatidae Dominican amber Poinar (2008b)
Streblidae Enischnomyia stegosoma Indirect Plasmodiidae Dominican amber Poinar (2011b),
Poinar and Brown (2012)

Hemiptera
Reduviidae Triatoma dominicana Indirect Trypanosomatidae Dominicsssan amber Poinar (2005c)
Torirostratidae Torirostratus pilosus Direct Fe concentrations ? Yixian Formation Yao et al. (2014)
Torirostratidae Flexicorpus acutirostratus Direct Fe concentrations ? Yixian Formation Yao et al. (2014)
The Importance of Fossils in the Evolution of Parasites 17

(Perrichot et al., 2012), although several extinct families of putative stem-


group Siphonaptera have been reported from the Jurassic and Cretaceous pe-
riods (Gao et al., 2012, 2013, 2014; Huang et al., 2012, 2013; Huang, 2014,
2015). Host associations with dinosaurs and pterosaurs have been suggested
for stem-group fleas (Huang, 2014), but a novel molecular study suggests
that the earliest fleas appeared in the early Cretaceous era and had a strong as-
sociation with mammals, whereas the Jurassic stem-group forms are only
distantly related (Zhu et al., 2015). Finding direct evidence of Jurassic or
Cretaceous fleas associated with host remains would be the smoking gun to
resolve this issue. The relationship between hosts and other haematophagous
insects, which are also important vectors, is less strict. Bed bugs and their close
relatives can probably be traced back to the Eocene (Engel, 2008). Diptera can
be traced back to the PermianeTriassic. Blood-sucking is inferred to be
ancestral in this group and members of various lineages are important vectors
(Poinar, 2014a; Nagler and Haug, 2015), including blackflies (Simulidae),
sandflies (Phlebotominae, Psychodidae) and mosquitoes (Culicidae). Direct
evidence for both haematophagy and their vectors is only known from
Diptera (Greenwalt et al., 2013; compare Table 2). Ticks are also important
vectors for Spirochaetes and Ricketsiales, which might have been already the
case since the Cretaceous period (Poinar, 2014b, 2015b).
Even when isolated from their hosts, gut contents or coprolites of fossil
parasitic invertebrates (Wappler et al., 2004; Greenwalt et al., 2013) or verte-
brate hosts (McConnell and Zavada, 2013), as well as other remains found
alongside, have allowed parasite remains to be confidently attributed to
major host groups, at least at higher taxonomic levels (Dentzien-Dias
et al., 2013; Da Silva et al., 2014; Hugot et al., 2014). Archaeological exam-
ples frequently allow species level identifications of helminth eggs (Dittmar
et al., 2011).
The earliest evidence for Metazoa parasitizing a plant is nematodes found
within early land plants (Poinar et al., 2008). Nevertheless, fungi can also be
vicious pathogens of animals and plants (Sexton and Howlett, 2006). The
oldest evidence for the presence of fungi parasitic on animals derives from
the Cretaceous period (Sung et al., 2008). Sometimes, the characteristic
response of the host to a fungal parasite can also be preserved (see Section
4.2). Fossilized galls can also be important sources of information on
planteparasite interactions (Knor et al., 2013; Labandeira and Currano,
2013; Leckey and Smith, 2015). Note that galls are often caused by parasitic
insects, but can also be induced by viruses, bacteria, fungi, nematodes and
mites (Knor et al., 2013).
18 Kenneth De Baets and D. Timothy J. Littlewood

3.2 Trace fossils and pathologies


Various invertebrate hosts including arthropods, molluscs, echinoderms and
various colonial organisms can contain traces or pathologies in their
skeleton, which is evidence of an infestation or association with parasites
(Boucot and Poinar, 2010; Donovan, 2015; Huntley and De Baets, 2015;
Taylor, 2015).
In some cases, these traces or pathologies are believed to be so character-
istic that they are interpreted to represent the oldest fossil evidence for
particular lineages in the fossil record; including castration of fossil decapods
by rhizocephalan barnacles (Feldmann, 1998, 2003; compare Klompmaker
and Boxshall, 2015), deformations in echinoderms attributed to myzostomid
annelids (Welch, 1976; Hess, 2010; Parry et al., 2014), crustacean arthropods
(Madsen and Wolff, 1965; Radwa nska and Radwa nska, 2005; Radwanska
and Poirot, 2010) or eulimid gastropods (Neumann and Wisshak, 2009) as
well as borings in octocorals related to the presence of ascothoracid barnacles
(Voigt, 1959, 1967). The most convincing palaeontological model systems
are those where both extant and fossil parasiteehost interactions are compa-
rably well studied such as the gymnophallid-induced pits and igloo-shaped
shell concretions in bivalves (Campbell, 1985; Ruiz and Lindberg, 1989;
Ituarte et al., 2001, 2005; Huntley, 2007; Todd and Harper, 2011; De Baets
et al., 2015a; Huntley and De Baets, 2015) and (?bopyrid) isopod swellings
in decapods (Weinberg Rasmussen et al., 2008; Boyko and Williams, 2009;
Williams and Boyko, 2012; Klompmaker et al., 2014; Klompmaker and
Boxshall, 2015). Particularly in such cases, trace fossils and associated pathol-
ogies can not only provide direct information on the behaviour of the par-
asites but also the response of the host, which makes it possible to interpret
the type of relationship between them. However, even in the case of such
model systems, it cannot be ruled out that they were made by another
type of closely related parasite or organism with similar behaviour in the
geological past, which are now extinct or where pathological reactions are
not yet documented in extant hosts. Other (now extinct) culprits can be sus-
pected when the record of these structures is not so continuous and shows
major stratigraphic gaps (Boucot and Poinar, 2010). For example: shell pits
have been confidently linked with gymnophallid flatworms in extant
bivalves, and can be confidently traced into the Eocene (Ruiz and Lindberg,
1989; Todd and Harper, 2011; De Baets et al., 2015a; Huntley and De Baets,
2015), which is more or less consistent with the origin of their final hosts
(extant shorebirds). However, igloo-shaped concretions e attributed to
The Importance of Fossils in the Evolution of Parasites 19

gymnophallids in extant bivalves (Ituarte et al., 2001, 2005), have also been
reported from the Silurian (Liljedahl, 1985), which is not consistent with
extant parasiteehost associations (De Baets et al., 2015a) as shorebirds (Char-
adriiformes), their present day definitive hosts, are believed to have radiated
sometime between the Cretaceous and Eocene periods (Smith, 2015).
Pathologies, therefore, offer less confident evidence for the presence of
parasitic lineages in the fossil record than in body fossils, when no parasitic
remains are found associated with these traces. Direct evidence for the par-
asites associated with such pathologies is mostly restricted to parasitic organ-
isms with mineralized skeletons such as gastropods (Hayami and Kanie,
1980; Lozouet and Renard, 1998; Baumiller and Gahn, 2002). Such traces
are usually compared with known responses to parasites by extant hosts,
which are sometimes not that well-investigated, and by extrapolation it
has been assumed that the same culprits were responsible in the past. This
can be further complicated by the fact that extant phylogenies indicate
that pathology-inducing lineages might have evolved more than once
(e.g. gall- and cyst-forming myzostomids: Summers and Rouse, 2014).
Parasite-induced pathologies have also been reported from hosts that are
now extinct or no longer affected; although their interpretation becomes
more difficult if no modern analogues are available (Owen, 1985; Babcock,
2007; De Baets et al., 2011; De Baets et al., 2015b). Traces which are
reminiscent of nematode borings in foraminifer tests (Sliter, 1971) have
been reported from Cambrian and Ordovician trilobites (Babcock, 2007),
but no conclusive assignment to nematodes as culprits can be made without
direct fossil evidence for associated nematodes. Furthermore, it is still
debated whether these traces in trilobites were made during life or post-
mortem (Owen, 1985). Various pathological reactions in ammonoids (an
extinct group of externally shelled cephalopods) have been attributed to
parasitic flatworms based on their prevalence and similar pathologies in
extant shelled molluscs (De Baets et al., 2011), but as long as no parasite
remains are found associated with them, their attribution to parasitic flat-
worms remains highly speculative at best (De Baets et al., 2015b).
Many palaeontologists also point out the difficulty of defining an inter-
action as being parasitic, although this might be more a problem concerning
the definition of parasitism rather than its recognition (Tapanila, 2008;
Zapalski, 2011), which is not restricted to fossil associations. Some authors
like Tapanila (2008) have suggested that fossil studies should assume that a
symbiosis is neutral (commensal), unless demonstrated otherwise. Other au-
thors have argued that a neutral interaction is absence of an interaction,
20 Kenneth De Baets and D. Timothy J. Littlewood

which cannot be proven, and is therefore unfit for empirical science. The
detection of commensalism is difficult and rather subjective in recent asso-
ciations (usually it is understood as a weak positive or negative interaction)
and as such it seems impossible to detect in the fossil record. Zapalski (2011)
has, therefore, argued avoiding commensalism as a null hypothesis in palae-
oecology, because the possibility of making a type II error is very high. Pos-
itive or negative effects can be detected or inferred based on comparisons
with extant interactions. Identifying traces or pathologies of fossil parasites
can potentially also be performed by demonstrating a negative influence
or effect on growth, body size and/or morphology of their hosts, while a
certain positive effect for the parasite can be inferred. Of course, such inter-
pretations rely on identifying the traces (e.g. borings) or structures as being
made in vivo. This can be most convincingly demonstrated when a host
response (e.g. growth deformation or pathology) can be shown to be asso-
ciated with these structures, often most readily recognized in specimens with
sparse traces or pathologies (De Baets et al., 2011; Donovan, 2015).
Studies have focused particularly on invertebrate hosts, and especially on
those with external shells or exoskeletons, but such pathologies which could
potentially be tracked in the fossil record are also found in vertebrates
including characteristic limb malformations in amphibians (Johnson et al.,
2001, 2002, 2003; Johnson and Sutherland, 2003) or cavities in the mastoid
bone of humans (Oyediran et al., 1975) caused by digenetic trematodes or
trabecula-like bone lesions in cetacean whales (Littlewood and Donovan,
2003) and enlargement of the frontal sinuses accompanied by bone lesions
in mustelids (Rothschild and Martin, 2006) caused by cestodes. Character-
istic skeletal pathologies in terrestrial vertebrates (e.g. mammals) induced by
helminths with resistant eggs have the potential for comparison of preva-
lence of skeletal deformations directly with parasite load or prevalence in
coprolites effectively linking palaeoparasitology and palaeopathology
(compare Dutour, 2013). Some parasitic unicellular pathogens might leave
characteristic traces or pathologies in their hosts. Wolff et al. (2009) studied
erosive lesions in tyrannosaurs and attributed them to Trichomonas gallinae-
like protozoans, because they are reminiscent of similar pathologies in extant
birds caused by this parasite. Unicellular eukaryotes can also leave character-
istic traces in their hosts; for example, borings by foraminifera in marine
echinoderms and bivalves (Neumann and Wisshak, 2006; Beuck et al.,
2007, 2008). In some cases, the host performs activities or exhibits behaviour
induced by the parasites, which can occasionally also be found in the fossil
record. One spectacular example is the death-grip scars found on Eocene
The Importance of Fossils in the Evolution of Parasites 21

leaves, interpreted to have been made by ‘zombie’ ants infested by fungi


(Hughes et al., 2010); the fungal infection Ophiocordyceps forces ants to
hold onto tips of leaves so that after the death of the ant, emerging fungal
spores can be released into the wind from an elevated position. Trace fossils
can give unique information on the behaviour and prevalence of the
parasites and track the response of their host through geological time, as their
record can be more continuous (less patchy) than that of body fossils which
only fossilize in exceptional conditions.
Pathologies have their own problems as they are sometimes hard to
assign to a certain lineage of culprits, although there are some pathologies
which are believed to be diagnostic, or at least characteristic, for parasitism.
Irrespective of this problem, the temporal and spatial record of parasite-
induced pathologies in their hosts can be much more continuous than the
parasite body fossil record, particularly in molluscs, echinoderms, colonial
organisms and others with fossilizable skeletons (Donovan, 2015; Huntley
and De Baets, 2015; Klompmaker and Boxshall, 2015; Taylor, 2015).
This can provide valuable quantitative data on various aspects of parasitee
host interactions, which can be tracked over millions of years (Brett,
1978; Ruiz and Lindberg, 1989; Baumiller and Gahn, 2002; De Baets
et al., 2011; Klompmaker et al., 2014; De Baets et al., 2015b; Huntley
and De Baets, 2015; Klompmaker and Boxshall, 2015). Such data can
only rarely be obtained by looking at body fossil records of parasiteehost as-
sociations, with some rare exceptions where both the host and the parasite
have fossilizable skeletons (Lozouet and Renard, 1998; Baumiller and Gahn,
2002) or where multiple similar taphonomic windows (e.g. amber) exist.
This includes direct information concerning prevalence and virulence of
this relationship, and their possible relationship with host evolution
including diversity (Klompmaker et al., 2014) and body size (Ruiz, 1991;
Huntley and De Baets, 2015), as well as environmental factors such as
sea-level and climate change (Huntley et al., 2014). This can be particularly
relevant to predict the future response of parasiteehost systems to global
change, where studies have suggested that parasites might be more prone
to (co)extinction (Dunn et al., 2009).
The link between the prevalence of parasites and pathologies, within pop-
ulations and particularly individual hosts, might not be straightforward and has
only rarely been investigated in extant hosts. Furthermore, the nature of the
relationship and their effects might be context dependent (Bronstein, 1994;
Daskin and Alford, 2012). Various preservation, collection and taxonomic
biases can and should be accounted for in quantitative analyses of fossil
22 Kenneth De Baets and D. Timothy J. Littlewood

antagonistic interactions (Huntley and De Baets, 2015; Klompmaker and


Boxshall, 2015), which have so far mainly focused on less specific (pred-
atoreprey) interactions (Kelley et al., 2003; Huntley and Kowalewski,
2007) rather than parasiteehost interactions. Various environmental factors
can, for example, influence the invasion of molluscs by parasites on various
organizational levels (Cheng and Combes, 1990). To further progress in
the field of quantitative palaeopathology in deep time, more quantitative
data and analyses on existing model systems are required, as well as the
need to identify additional modern and/or fossil analogues of already identi-
fied parasite-related pathologies. Palaeopathologies with a more continuous
fossil record have the potential to be used to model the influence of parasite
prevalence and virulence on the evolution of their hosts and how they are
modulated by environmental parameters on longer time-scales.

3.3 Coprolites
Coprolites are usually defined as fossilized (permineralized) faeces, although
the term is often also used for desiccated, more recent faeces from archaeo-
logical sites (Ferreira et al., 1991; Reinhard and Bryant, 1992; Hunt et al.,
2012). Coprolites have yielded fossil and archaeological evidence for para-
sitic organisms (Table 3) ranging from coccidia or other protozoans (Ferreira
et al., 1992; Schmidt et al., 1992; Poinar and Boucot, 2006; Frías et al., 2013;
Wood et al., 2013b), to parasitic fungi and plant remains (Sharma et al.,
2005; Wood et al., 2012), to helminths (Gonçalves et al., 2003; Savinetsky
and Khrustalev, 2013), including acanthocephalans (Noronha et al., 1994),
but particularly nematodes (Ferreira et al., 1991, 1993; Poinar and Boucot,
2006; Leles et al., 2010; Da Silva et al., 2014; Hugot et al., 2014) and various
groups of parasitic flatworms (Schmidt et al., 1992; Jouy-Avantin et al.,
1999; Dentzien-Dias et al., 2013). Coprolites can therefore be an important
additional source of ancient parasitism supplementary to amber, where the
record is heavily biased towards arthropods and their terrestrial parasites.
In ideal cases, coprolites are still associated with their producer, which
makes it possible to confidently identify their origin and therefore the
host taxon of the fossil parasites. The coprolite producer may correspond
with the host of the parasite or more rarely as the one who ingested the para-
site and/or host. However, most frequently, coprolites are found in isola-
tion, where the identity of the producer can only be inferred from their
morphology and content (Poinar and Boucot, 2006; Dentzien-Dias et al.,
2013), and, in the case of more recent specimens, by aDNA analysis
(Wood and Wilmshurst, 2014). Invertebrate coprolites might also have
The Importance of Fossils in the Evolution of Parasites
Table 3 Coprolites depicting fossil and archaeological evidence for parasitic organisms
Taxonomic affinity Fossil evidence Age Host References
Protozoa
Coccidia
Eimeriorina Oocyst Holocene Ground sloth Schmidt et al. (1992)
(Nothrotheriops
shastensis)
Eimeridae Oocysts (Eimera) Holocene Deer Ferreira et al. (1992)
Cryptosporidiidae aDNA (Cryptosporidium) Holocene Moas (Dinomis robustus, Wood et al. (2013b)
Pachyornis
elephantopus)
Helminths
Platyhelminthes
Cestoda Egg þ developing Permian ?elasmobranchs Dentzien-Dias et al. (2013)
embryo
Trematoda Eggs Cretaceous Archosaur (?dinosaur) Poinar and Boucot (2006)
Dicrocoelidae Eggs Pleistocene Mammal (?bear) Jouy-Avantin et al. (1999)
Schistosomatidae Schistosome-like eggs Holocene Ground sloth Schmidt et al. (1992)
(Nothrotheriops
shastensis)
Nematoda
? Larvae Pleistocene Hyenid Ferreira et al. (1993)
Ascaridomorpha Eggs Triassic Cynodont Da Silva et al. (2014)
Oxyurida Eggs Triassic Cynodont Hugot et al. (2014)
(Continued)

23
Table 3 Coprolites depicting fossil and archaeological evidence for parasitic organismsdcont'd

24
Taxonomic affinity Fossil evidence Age Host References
Ascaridomorpha Egg þ developing Cretaceous Archosaur (?dinosaur) Poinar and Boucot (2006)
larvae
Heterakoidea aDNA Holocene Archosaur Wood et al. (2013b)
(Anomalopteryx,
Dinornis, Pachyornis,
Megalapteryx)
Trichocephalida Eggs (Trichuris) Pleistocene Hyenid Ferreira et al. (1991)
Acanthocephala
? Eggs Holocene Humans Fry and Hall (1969)
Eggs (Echinopardalis) Holocene Felidae Noronha et al. (1994)
Arthropods
Ticks Body remains Holocene Humans Johnson et al. (2008)

Kenneth De Baets and D. Timothy J. Littlewood


Lice Body remains Holocene Humans Fry (1977)
Fungi
Plant-parasitic fungi Spores Cretaceous Archosaur (?dinosaur) Sharma et al. (2005)
Plants
Root-parasite Pollen (Dactylanthus Holocene Kakapo (Strigops Wood et al. (2012)
taylorii) habroptilus)
The Importance of Fossils in the Evolution of Parasites 25

the potential to reveal past parasite infections, as demonstrated by reports of a


putative trypanosome from faecal droplets found in association with a fossil
triatomine in Dominican amber (Poinar, 2005c). Protozoa remains have also
been reported from the abdominal region of Miocene Tapir remains
(McConnell and Zavada, 2013) as well as in termites preserved in amber
(Poinar, 2009).
Egg remains are usually considered trace fossils, but they can occasionally
contain remains of developing embryos (Poinar and Boucot, 2006;
Dentzien-Dias et al., 2013), which are body fossils. Interestingly, parasite
eggs have often been found in coprolites or fossilized faeces, which are trace
fossils themselves. Some of these coprolites range back to the Palaeozoic
(Zangerl and Case, 1976; Dentzien-Dias et al., 2013) or Mesozoic periods
(Poinar and Boucot, 2006; Da Silva et al., 2014; Hugot et al., 2014),
although most are known from the Cenozoic, particularly the Quaternary
era (Ferreira et al., 1991; Jouy-Avantin et al., 1999; Gonçalves et al.,
2003). As some of these eggs can be quite resistant to decay, they can also
be found in Quaternary sediments, particularly in archaeological sites
(Bouchet et al., 2003; Gonçalves et al., 2003; Ara ujo et al., 2015). In other
cases, parasite eggs have been reported associated with Cretaceous fossil
feathers (Martill and Davis, 1998) or Eocene mammal hair in Baltic amber
(Voigt, 1952; Nagler and Haug, 2015).
Remains of fossil parasites have typically been revealed through destruc-
tive preparation (e.g. thin sectioning, dissolution/rehydration methods) and
classical imaging methods leaving little chance for further study, or have
been missed or destroyed because of techniques designed to reveal structures
of the hosts to which they are associated. Traditional methods of parasite
recovery from coprolites, or other ancient samples, can be combined with
aDNA techniques (Wood et al., 2013b; Ara ujo et al., 2015) or computed
tomography.
Sequencing of aDNA might successfully detect very small and/or fragile
parasites that may not preserve intact in coprolites, or can be destroyed dur-
ing the extraction or preparation methods. This method is however also
destructive (e.g. samples need to be rehydrated to extract DNA) and
restricted to younger (archaeological) samples due to the rapid deterioration
of DNA/RNA, making computer tomography particularly important to
identify parasites for older (fossil) samples.
Tomographic methods might help to reveal additional details of fossils
trapped in amber as well as help to discover parasites in coprolites or other
ancient remains which can be destroyed during the traditional destructive
26 Kenneth De Baets and D. Timothy J. Littlewood

preparation processes. The potential of these windows into parasite evolu-


tion has become clear in the last two decades by multiple new discoveries,
particularly in marine and terrestrial coprolites (Table 3) as well as in amber
fossils (Tables 1, 2). Parasitic remains in fossil coprolites have been reported
for a long time (Zangerl and Case, 1976), but were usually received with a
certain degree of scepticism (Boucot, 1990; Poinar, 2003). Various methods
can yield high-resolution reconstructions of microscopic remains or struc-
tures in larger fossils including mCT-scanning or phase-contrast synchrotron
tomography (Donoghue et al., 2006; Sutton et al., 2014), but their useful-
ness depends on the contrast between the fossils and the matrix. Computed
tomography can not only be relevant to identify and characterize parasite re-
mains, perhaps to place them in extant phylogenies (Faulwetter et al., 2013;
Garwood and Dunlop, 2014), but also to quantify their original position, 3D
structure and morphology or association in coprolites or other ancient re-
mains (Dunlop et al., 2012; Siveter et al., 2015). It could potentially be
used to study the presence of developing embryos (Donoghue and Dong,
2005; Donoghue et al., 2006; Duan et al., 2012), which have been suggested
to be present in some fossil eggs attributed to helminths based on traditional
imaging methods (Poinar and Boucot, 2006; Dentzien-Dias et al., 2013).
These could even provide supplementary information on the morphology
and content of the coprolites, which could make it possible to more confi-
dently identify their host, the coprolite producers, as well as give indications
about possible predatoreprey relationships and parasite load. In archaeolog-
ical studies, coprolites can be quantitatively studied to establish changes in
biogeographic distributions and habits of their hosts (Ara ujo et al., 2015;
Mitchell, 2015) as well as on parasite body size changes (Fugassa et al.,
2008). Additional data are required to quantitatively study these aspects
on longer (palaeontological) timescales as the parasitological coprolite studies
are still quite patchy in space and time.

4. MOLECULAR PERSPECTIVES ON PARASITE


PHYLOGENY AND EVOLUTION
Historically, for many taxonomic groups of parasites, morphologically
based systematic schemes and phylogenies have been difficult to resolve
even in the light of additional sampling or analysis. Parasites often demonstrate
high degrees of specialization compared with their free-living relatives, high
degrees of reduction or apparent simplification, specialization and/or conver-
gence in morphology related to their parasitic lifestyle. Complex life cycles
The Importance of Fossils in the Evolution of Parasites 27

involving multiple hosts or even a single host can consist of morphologically


distinct ontogenetic stages making homology assessment and the identifica-
tion of shared features even more difficult (Brooks and McLennan, 1993).
Molecular methods have the potential to resolve many problems where
morphology has been problematic, as long as one can properly deal with
biases and spurious signal related to long-branch attraction, sampling,
(host) contamination and other issues arising from a purely molecular
approach (Edgecombe et al., 2011). Contamination is not only relevant
for aDNA (Shapiro and Hofreiter, 2014), but might also be responsible
for wrong assignment of various groups of extant parasites. Previously erro-
neous assignment of Myxozoa, including the vermiform Buddenbrockia, with
other taxa has been attributed to host contamination (Jiménez-Guri et al.,
2007). On the other hand, molecular studies have had a considerable impact
on the assignment of some parasite groups within the broader context of
metazoan evolution (Zrzavý, 2001; Edgecombe et al., 2011). For example,
pentastomids, long considered to be a separate phylum, are now considered
to be closely related to fish lice based on molecular evidence (Sanders and
Lee, 2010; Oakley et al., 2012). Also, Myxozoa are now considered to be
cnidarians based on molecular evidence (Jiménez-Guri et al., 2007;
Hartikainen et al., 2014; Okamura and Gruhl, 2015; Okamura et al.,
2015). The phylogenetic position of myzostomids has also been long
debated, but most authors now agree that they belong within the Annelida
based on molecular analyses (Parry et al., 2014; Summers and Rouse, 2014).
The discovery of Chromera velia, the first photosynthetic apicomplexan
with a fully functional plastid, might also provide a powerful model to study
the evolution of parasitism in Apicomplexa. Molecular analyses indicate that
it is the closest relative to apicomplexan parasites, indicating that the plastid
of this coral symbiont shares its origin with the apicoplasts (Moore et al.,
2008; Okamoto and McFadden, 2008).
Acanthocephala (thorny-headed worms) were occasionally compared to
priapulids (penis worms) based on morphological evidence (Conway Morris
and Crompton, 1982), but are now aligned with Rotifera within the
Syndermata (Weber et al., 2013; Wey-Fabrizius et al., 2013). Most recent
molecular studies with greater coverage indicate a particular route to para-
sitism within Platyhelminthes and a closer relationship between cestodes
and trematodes (Lockyer et al., 2003; De Baets et al., 2015a; Egger et al.,
2015; see Littlewood and Waeschenbach, 2015 for a review), although there
are some exceptions (Laumer et al., 2015), which illustrate that the relative
importance of particular gene regions needed to disentangle phylogenetic
28 Kenneth De Baets and D. Timothy J. Littlewood

relationships still need to be better understood. A recent molecular study


(Struck et al., 2014) also indicates that the Platyzoa, which groups Acantho-
cephala and Platyhelminthes among various free-living taxa, is probably an
artefact of long-branch attraction, whereby distantly related lineages are
incorrectly inferred to be closely related because both lineages have under-
gone a considerable amount of change. This theory has since been
confirmed by Egger et al. (2015), who demonstrated that only the fast-
evolving quartile of their transcriptomic dataset supported the Platyzoa.
Importantly, in the absence of phylogenetic artifacts, molecular studies
can be used to investigate the origins and radiations of parasitic lineages.
This has been shown for acanthocephalans (Near, 2002; Wey-Fabrizius
et al., 2014), parasitic flatworms (Lockyer et al., 2003), myzostomid annelids
(Summers and Rouse, 2014) and nematodes (Dorris et al., 1999; Blaxter,
2003; Blaxter and Koutsovoulos, 2015). Sister-group relationships within
and between parasite lineages, the determination of free-living sister groups
and the inference of ancestral life history strategies are key to determine the
origins of parasitism and the evolution of complex life cycles. The evolution
of complex life cycles is a major research question, with two main mecha-
nisms proposed (Parker et al., 2015): upward incorporation by terminal
addition of hosts, or downward incorporation by addition of intermediate
hosts. Cladistic studies, or the application of parsimony principles to infer
ancestral life cycles, have yielded some controversial hypotheses as to the or-
igins and development of ontogenetic sequences (O’Grady, 1985). In para-
sitic flatworms (Neodermata), complex parasite life cycles appear to have
initially evolved by the addition of intermediate hosts, with vertebrate defin-
itive hosts argued to be the plesiomorphic condition for stem group neoder-
matans (Littlewood et al., 1999); the scenarios are inferred from molecular
phylogenies of extant taxa; however, fossil vertebrates may yet hold the
key to verifying or at least supporting this claim.
Since their initial application, molecular clock methodologies have
undergone major developments (Bromham and Penny, 2003; W€ orheide
et al., 2015), with ever more sophisticated models accommodating large
genomic datasets, rate variation and the uncertainties of multiple types of
calibration points (Parham et al., 2012; Ho, 2014). Nevertheless, molecular
clocks still need age constraints from the fossil record or other lines of
evidence to provide calibration points for at least one or more nodes in a
calibrated phylogeny (Ho and Phillips, 2009; Hipsley and M€ uller, 2014;
Warnock, 2014). The use of fossils in molecular clocks has changed in the
last decades (Parham et al., 2012; Ho, 2014; W€ orheide et al., 2015)
The Importance of Fossils in the Evolution of Parasites 29

including their implementation as probabilistic time priors (Ho and Phillips,


2009; Warnock et al., 2015) and tip calibrations integrated among their
living relatives by combing molecular and morphological evidence (Pyron,
2011; Ronquist et al., 2012; Wood et al., 2013a; Arcila et al., 2015). How-
ever, methods using other types of age evidence like geological calibrations
(e.g. vicariance events) or host calibrations lag behind in their development
and leave them heavily scrutinized (Goswami and Upchurch, 2010; Kodan-
daramaiah, 2011; Hipsley and M€ uller, 2014).
Finally, advances in molecular techniques have provided additional
advances in unravelling past events. aDNA techniques have made it possible
to obtain pathogen DNA from dated ancient parasites or host remains
(Dittmar, 2009; Bos et al., 2015; Hofreiter et al., 2015). Although many
of these studies have been targeted in their approach, there are also those
that have characterized complete biomes (Rawlence et al., 2014); for
example, detection and characterization of bacteria entombed in dental
calculus (Warinner et al., 2015), and the discovery of subfossil coprolites
(Santiago-Rodriguez et al., 2013; Wood et al., 2013b; Cano et al., 2014)
has opened up new avenues in palaeomicrobiology.

4.1 Molecular clocks


The ancient origin of parasitism is suggested by the deep-branching position
of various parasitic lineages within the tree of life as well as the extrapolation
of extant parasite host associations. In prokaryote and unicellular eukaryote,
parasitic relationships might have already existed in the Precambrian,
although so far no direct fossil evidence has been found. For metazoan par-
asites, for which the fossil constraints are better than they are for unicellular
organisms, parasitism must have evolved at least for some groups during or
slightly before the Cambrian explosion in the marine realm. This is corrob-
orated by the earliest sign of parasitism by an unknown metazoan parasite in
Cambrian brachiopods (Bassett et al., 2004) and body fossil remains of
various pentastomid taxa from the CambrianeOrdovician (Castellani
et al., 2011). Additional indications for parasitism are also found in other
Lower Palaeozoic groups like Ordovician graptoloids (hemichordates)
(Bates and Loydell, 2000) as well as in Silurian echinoderms (Franzen,
1974) and bivalves (De Baets et al., 2015a), although the exact identity of
the culprits is unknown. Based on extrapolation of extant host-associations,
the origin of parasitic flatworms and some lineages of parasitic nematodes
have also been estimated to lie in the CambrianeOrdovician (Littlewood,
2006; Poinar, 2011a, 2015a). Since the assignment of Cambrian
30 Kenneth De Baets and D. Timothy J. Littlewood

Cambroclavida (a group of enigmatic, phosphatized, hollow spine-shaped


sclerites) to Acanthocephala (Qian and Yin, 1984) is unsubstantiated (Kou-
chinsky et al., 2012), the earliest confidently assigned ancient acanthoceph-
alan remains are eggs derived from Quaternary archaeological sites (Fry and
Hall, 1969; Noronha et al., 1994). The earliest metazoan helminth remains
are circlets of hooks from Middle Devonian fishes (Upeniece, 2001),
although their systematic affinity is still unclear (Littlewood and Donovan,
2003; De Baets et al., 2015a). So far no direct unequivocal evidence for
metazoan parasites has been discovered in the Precambrian. Ediacaran fossils
like Dickinsonia have occasionally been compared with Spinther (Wade,
1972; Conway Morris, 1981), an annelid parasite of sponges (Rouse,
2005b). This comparison proved to be superficial at best and they are
now often interpreted to be more basal metazoans (Xingliang and Reitner,
2006; Sperling and Vinther, 2010), although the exact systematic position of
Dickinsonia remains highly controversial (Retallack, 2007; Brasier and Ant-
cliffe, 2008). More importantly, no evidence for a parasitic relationship of
Dickinsonia with other taxa could be evidenced. A parasitic mode of life of
Dickinsonia would be rather absurd too as it would mean the presence of
hosts which would have to be considerably larger than Dickinsonia, which
have remained unnoticed in the fossil record.
Considering that bacteria, viruses and various other unicellular to multi-
cellular eukaryotes have evolved before this time, many could have already
evolved towards parasitism in the Precambrian. The discovery that the
closest relative of parasitic apicomplexans is a coral symbiont might suggest
that modern parasites may have started out as mutualistic metazoan symbi-
onts before turning to parasitism (Moore et al., 2008). It has, therefore, been
suggested that symbiotic/parasitic relationships in dinoflagellates and Api-
complexa might have extended back in evolutionary time to the earliest or-
igins of Metazoa, which means that either as parasites or symbionts, these
protists have been interacting with the metazoan immune system since their
inception (Okamoto and McFadden, 2008). However, calibrating molecu-
lar clocks to test such hypotheses remains a challenge (Bensch et al., 2013).
Parasitism-like life history patterns probably first evolved in the sea, but it
is unclear if parasites closely tracked the terrestrialization of their hosts. In
some cases this might have been true, as indicated by nematode remains
in early diverging land plants from the Early Devonian (Poinar et al.,
2008), although both plants and nematodes might have colonized the
land considerably earlier based on molecular clock estimates (Clarke et al.,
2011; Rota-Stabelli et al., 2013). In other cases it is less clear as the earliest
The Importance of Fossils in the Evolution of Parasites 31

fossil evidence for terrestrial parasitism in these lineages considerably predates


the presumed origin of their host taxa (e.g. vertebrates or arthropods). How-
ever, various groups of terrestrial vertebrates (archosaurs, synapsids) were
already parasitized in now extinct lineages leading up to currently infested
host taxa (Poinar and Boucot, 2006; Da Silva et al., 2014). Although para-
sitism evolved in the sea, using this premise to constrain molecular clocks
might be dangerous in particular cases; as is the case for various groups of
nematodes parasitizing shallow marine to intertidal taxa, which have evolved
convergently from terrestrial ancestors (Sudhaus, 2010).
The fossil record can only provide minimum constraints on the origin of
parasitism, but molecular clocks might be an alternative to dating important
events without relying on recent evidence of parasiteehost associations or
the evolutionary history of their hosts. The calibration of molecular clocks
in pathogens with a poor fossil record like viruses, bacteria, protozoa (Bensch
et al., 2013; Frías et al., 2013) or soft-bodied helminths (De Baets et al., 2015a)
is not straightforward. The direct record is largely restricted to more recent
Quaternary sites, particularly archaeological sites (Ara ujo et al., 2015;
Mitchell, 2015) which can put important constraints on shallower nodes in
phylogenies, although deeper nodes, particularly the root, in phylogenies
are more crucial for dating (Warnock et al., 2012; Mello et al., 2014). Further-
more, most authors agree that multiple calibration points implemented faith-
fully across a phylogeny are ideal in achieving accurate and precise divergence
estimates. However, these calibration points should be screened and selected a
priori, rather than using posteriori selections methods, which evaluate
congruence through cross-validation, as the latter can lead to selection of
congruent, erroneous calibrations (Warnock et al., 2015).
Nevertheless, the summary above shows that the body fossil record offers
confident minimum constraints for various lineages of parasites, when crit-
ically evaluated. Various methods have been developed to estimate diver-
gence times based on the stratigraphic distribution of fossil data
(Wilkinson et al., 2011; Nowak et al., 2013; Heath et al., 2014), but most
studies apply phylogenetic bracketing (M€ uller and Reisz, 2005; Benton
and Donoghue, 2007) or probability functions that express some predefined
perception of the degree to which fossil minima approximate the true time
of divergence (Ho and Phillips, 2009). The latest tip-calibration or total ev-
idence dating methods (Pyron, 2011; Ronquist et al., 2012; Wood et al.,
2013a) allow fossils to be integrated into divergence time studies among
their living relatives, using combined morphological and molecular datasets
and evolutionary models. These have been shown, however, to yield
32 Kenneth De Baets and D. Timothy J. Littlewood

unexpectedly old age estimates of clades (Arcila et al., 2015) and their per-
formance needs to be more extensively tested. Furthermore, such methods
might be difficult to apply to soft-bodied taxa as crucial morphological char-
acters required to confidently place them in extant phylogenies might be ab-
sent or limited in fossil parasite specimens.
Computed tomography will be important to reveal additional details of
the morphology and structure of putative body fossils, which will make it
possible to assign them more accurately to extant lineages. In some cases,
the nearest free-living relatives have a good fossil record, which can be
used to put constraints on early nodes in molecular clocks. Interestingly,
this is not always the case. In some soft-bodied helminths, the body fossil
record of parasitic forms is even richer (at least less poor) than that in their
free-living relatives, such as amongst Platyhelminthes (Poinar, 2003; De
Baets et al., 2015a) and Nematoda (Poinar, 2011a, 2015a). Their fossil re-
cord remains are rare in time and space due to their restriction to sites of
exceptional preservation, but can potentially still be valuable to place con-
straints on the evolution of these groups as a whole. In the absence of reliable
body fossils, characteristic traces or pathologies could potentially also be used
to put constraints on certain nodes and computed tomography could also be
possibly used to characterize those (Dittmar et al., 2011). Unfortunately,
skeletal responses to parasitism are still comparatively poorly studied, partic-
ularly in extant taxa (Zibrowius, 1981; Ituarte et al., 2001, 2005; Keupp,
2012; Klompmaker et al., 2014). This makes interpretation of fossil traces
even more open to interpretation as they could also have been made by a
different group of organisms with a similar behaviour, but not necessarily
closely related. It therefore probably makes more sense to avoid using
them to constrain molecular clocks directly. Molecular clocks constrained
by other types of evidence (e.g. body fossils or geological events) could how-
ever be used to test the appearances of these skeletal responses.
Problematically, the fossil record does not yield body fossils or other re-
mains for multiple lineages of unicellular pathogens or soft-bodied metazoan
parasites (e.g. Myxozoa, Argulidae). In these cases, it is therefore necessary to
look for and select suitable alternatives, or supplementary ways, to constrain
the molecular clock (Bensch et al., 2013; De Baets et al., 2015a; Héritier
et al., 2015). Such solutions potentially lie in the host fossil record or biogeo-
graphic events, which have left a footprint of divergence among
evolutionary lineages. Unfortunately, host or biogeographic calibrations
have not received the same scrutiny and refinement as fossil calibrations
(Kodandaramaiah, 2011; De Baets and Donoghue, 2012; Parham et al.,
The Importance of Fossils in the Evolution of Parasites 33

2012; Hipsley and M€ uller, 2014). Biogeographic dating often relies on


geochronologically or otherwise geologically dated events. As they are
currently implemented, biogeographic calibrations and their age evidence
are rarely if ever justified; they often assume the biogeography of living or-
ganisms is a faithful reflection of ancestral distribution, which is not always,
or even rarely, the case as they might have been modified or even reset by
subsequent events. Some exceptions have been suggested in parasites but
these still need to be tested; for example, the cestode Nesolecithus and the
nematode Nilonema have an apparent Gondwanaland origin with their
present day distribution in Africa and South America (Gibson et al., 1987;
Santos and Gibson, 2007). Furthermore, tectonic episodes are protracted
and might have different impacts on lineages depending on their ecology.
These limitations can be partially overcome or at least controlled like fossil
calibrations, by implementing them in the most conservative way as proba-
bilistic constraints that span an interval of time, which takes into account
these factors (Warnock, 2014). Parasitologists often extrapolate extant
parasiteehost or biogeographic relationships to estimate the evolutionary
origin of parasites (Mejía-Madrid, 2013 for a review) or more rarely to cali-
brate molecular clocks, which can introduce a factor of circularity when
testing hypotheses of evolutionary changes in parasiteehost associations or
biogeographic distribution (Trewick and Gibb, 2010; Crisp et al., 2011;
Kodandaramaiah, 2011; Hipsley and M€ uller, 2014). Although some of these
hypotheses do stand the test of additional sampling of extant or fossil forms,
others do not if quite different host associations or biogeographic distribu-
tions are recovered. Caution should be always taken as the fossil record
also yields evidence of parasitic lineages and parasiteehost associations which
are now clearly extinct (Upeniece, 2001; Poinar and Boucot, 2006;
Castellani et al., 2011; Upeniece, 2011; Chen et al., 2014). Biogeographic
distributions of their hosts and potentially that of their parasites might also
have differed considerably in the past. Therefore, assumptions should be
at least consistent with the fossil evidence and robust molecular clock
estimates of their hosts. Not only does the fossil record provide direct and
indirect evidence for the presence of certain parasite lineages, but it can
also provide evidence for infection intensity, host response (developmental,
pathological), novel (potentially extinct) host associations and parasite life
cycles. Likewise, ancestral state reconstruction of host associations and life
cycles (host use and complexity) from phylogenies, combined with calibra-
tions can inform the interpretation of body fossils and inferred chronologies
(Zhu et al., 2015). All factors considered, we should prefer to have an
34 Kenneth De Baets and D. Timothy J. Littlewood

accurate timescale that might lack precision, than a precise timescale that
lacks the necessary accuracy (De Baets et al., 2015a). Even if no suitable cali-
bration points can be found, methods have been developed, which can
compare the relative molecular rates of groups to test the hypotheses of
co-divergences (Loader et al., 2007; Hibbett and Matheny, 2009; Loss-Oli-
veira et al., 2012; Silva et al., 2015).

4.2 HGT and ‘parasitic DNA’


Even for groups where no fossil biomolecules have been found, for example,
viruses or symbionts, parasites can leave footprints in their hosts’ genomes
(Gilbert and Feschotte, 2010; Thézé et al., 2011; Katzourakis, 2013;
Koutsovoulos et al., 2014). Interrogation of genomes has made it possible
to identify horizontal transfers of genetic elements (HGTs) in the deep
history of living organisms. HGTs are the transfer of DNA between two
nonvertically related individuals belonging to the same or different species
(Sj€
ostrand et al., 2014). Some of these transfers retain an apparent parasitic
role (Kidwell and Lisch, 2001) or become integrated into biochemical
pathways that are functionally important in lineages that become parasitic
(Alsmark et al., 2013).
Comparison of such genomic signatures between species provides a
means of determining their origins, diversification and change through
time. Studies have focused particularly on ancient viruses (Gilbert and
Feschotte, 2010; Thézé et al., 2011; Gifford, 2012; Herniou et al., 2013;
Lee et al., 2013), typically revealing ‘hosteparasite’ interactions over
prehistoric or geological timescales. There is increasing evidence that
HGTs have left genomic signatures of other more highly organized symbi-
onts like bacteria in their metazoan host genomes, too (Cerveau et al., 2011;
Koutsovoulos et al., 2014). Things may be even more complicated as evi-
dence of HGT may have occurred between various types of endosymbiotic
bacteria (Duron, 2013) or viruses within their hosts (Niewiadomska and
Gifford, 2013). There is no physical ‘fossil record’ of these viruses or signa-
tures (Katzourakis and Gifford, 2010; Katzourakis, 2013), and their long
unchanged history is inferred such that they cannot be referred to as
genomic ‘fossils’ as such. However, further evidence of historical hoste
parasite interactions will undoubtedly arise from future genomic studies of
both hosts and parasites, although an understanding of the role of HGT in
eukaryotes is still in its infancy (Hirt et al., 2015).
HGT events are common in prokaryotes and many microbial eukary-
otes, but are expected to become more commonly detected in multicellular
The Importance of Fossils in the Evolution of Parasites 35

eukaryotes as more is known about their genomes (Andersson, 2005). With


microbial phylogenies, HGTs are important in revealing which evolutionary
lineages were concurrent and when speciation (or broader divergence)
events took place. By explicitly modelling the evolution of genes present
in genomes, Sz€ ollTsi et al. (2012) provided a chronologically ordered
phylogeny for cyanobacteria, validated against the groups’ good microfossil
record, thus showing that their methods can reveal and use HGTs as a source
of information on timing (or at least chronology) of evolutionary events.
Focusing on microbial eukaryotes, Alsmark et al. (2013) showed the pattern
of HGTs retained after parasite diversification are likely to be functionally
important for the parasites (e.g. in kinetoplastids and apicomplexans).
Similarly, there is strong evidence that multiple HGT events have promoted
the plant parasitism ability in some nematodes (Danchin et al., 2010).
There is increasing evidence that the role of HGTs in eukaryote evolution
is important, particularly in the evolution of resistance, and it does not seem
overly speculative to predict that signatures from hosteparasite interactions
will be found in more genomes. Further interrogation will reveal lineages
of parasites with genomic signatures of their long-associated histories with
their host groups, and vice versa; for example, Richards et al. (2011);
Wijayawardena et al. (2013); Davis and Xi (2015). In turn, this evidence
will provide indications as to when particular lineages came into contact
with one another, but the success of these leads in revealing accurate records
of historical interactions depends on greater study. Some recent claims of
HGT in multicellular parasites (e.g. schistosomes) have been discredited,
revealing the need to be wary of technical artifacts and gene conservation is-
sues before claims of HGT can be verified (Wijayawardena et al., 2015).

5. FUTURE PERSPECTIVES
Various new advances in ancient biomolecule detection and charac-
terization (Briggs and Summons, 2014) including aDNA (Dittmar, 2009,
2014; Wood et al., 2013b; Dittmar, 2014; Shapiro and Hofreiter, 2014;
Hofreiter et al., 2015), palaeoproteomics (Hofreiter et al., 2012), novel
development in molecular clock methodologies (Parham et al., 2012; Ho,
2014; W€ orheide et al., 2015) and new possibilities for the critical evaluation
and nondestructive analysis of 3D fossil structures by computed tomography
(Cunningham et al., 2014a,b; Sutton et al., 2014) offer many new prospects
and perspectives in palaeoparasitology.
36 Kenneth De Baets and D. Timothy J. Littlewood

ACKNOWLEDGEMENTS
We are very grateful to Andrea Waeschenbach and Rod Bray for constructive comments on
an earlier draft of the manuscript. The initial research leading to this article was partially
funded by an SNF-grant for Prospective Researchers to KDB (141438).

REFERENCES
Alsmark, C., Foster, P.G., Sicheritz-Ponten, T., Nakjang, S., Embley, M.T., Hirt, R.P.,
2013. Patterns of prokaryotic lateral gene transfers affecting parasitic microbial
eukaryotes. Genome Biol. 14, R19.
Andersson, J.O., 2005. Lateral gene transfer in eukaryotes. Cell Mol. Life Sci. 62, 1182e1197.
Araujo, A., Reinhard, K., Fernando Ferreira, L., 2008. Parasite findings in archeological
remains: diagnosis and interpretation. Quat. Int. 180, 17e21.
Ara
ujo, A., Reinhard, K., Ferreira, L.F., 2015. Palaeoparasitology e human parasites in
ancient material. Adv. Parasitol. 90, 349e387.
Arcila, D., Alexander Pyron, R., Tyler, J.C., Ortí, G., Betancur-R, R., 2015. An evaluation
of fossil tip-dating versus node-age calibrations in tetraodontiform fishes (Teleostei:
Percomorphaceae). Mol. Phylogenet. Evol. 82, 131e145.
Asara, J.M., Schweitzer, M.H., Freimark, L.M., Phillips, M., Cantley, L.C., 2007. Protein
sequences from mastodon and Tyrannosaurus rex revealed by mass spectrometry. Science
316, 280e285.
Babcock, L.E., 2007. Role of malformations in elucidating trilobite paleobiology: a historical
synthesis. In: Mikulic, D.G., Landing, E., Kluessendorf, J. (Eds.), Fabulous Fossilse
300 Years of Worldwide Research on Trilobites. The State Education Department,
New York.
Bassett, M.G., Popov, L.E., Holmer, L.E., 2004. The oldest-known metazoan parasite?
J. Paleontol. 78, 1214e1216.
Bates, D.E.B., Loydell, D.K., 2000. Parasitism on graptoloid graptolites. Palaeontology 43,
1143e1151.
Baumiller, T.K., Gahn, F.J., 2002. Record of parasitism on marine invertebrates with special
emphasis on the platyceratid-crinoid interaction. In: Kowalewski, M., Kelley, P.H.
(Eds.), The Fossil Record of Predation, pp. 195e209.
Beaucournu, J.-C., 2003. Palaeopsylla groehni n. sp., quatrieme espece de Puce connue de
l’ambre de la Baltique (Siphonaptera, Ctenophthalmidae). Bulletin de la Société ento-
mologique de France 108, 217e220.
Beaucournu, J., Wunderlich, J., 2001. A third species of Palaeopsylla Wagner, 1903, from
Baltic amber (Siphonaptera: Ctenophthalmidae). Entomologische Zeitschrift 111, 296.
Bensch, S., Hellgren, O., Krizanauskien_e, A., Palinauskas, V., Valki unas, G., Outlaw, D.,
Ricklefs, R.E., 2013. How can we determine the molecular clock of malaria parasites?
Trends Parasitol. 29, 363e369.
Benton, M.J., Donoghue, P.C.J., 2007. Paleontological evidence to date the tree of life. Mol.
Biol. Evol. 24, 26e53.
Bertazzo, S., Maidment, S.C., Kallepitis, C., Fearn, S., Stevens, M.M., Xie, H.N., 2015.
Fibres and cellular structures preserved in 75-million-year-old dinosaur specimens.
Nat. Commun. 6, 7352.
Beuck, L., Correa, M., Freiwald, A., 2008. Biogeographical distribution of Hyrrokkin
(Rosalinidae, Foraminifera) and its host-specific morphological and textural trace
variability. In: Wisshak, M., Tapanila, L. (Eds.), Current Developments in Bioerosion.
Springer Berlin Heidelberg, pp. 329e360.
The Importance of Fossils in the Evolution of Parasites 37

Beuck, L., Vertino, A., Stepina, E., Karolczak, M., Pfannkuche, O., 2007. Skeletal response
of Lophelia pertusa (Scleractinia) to bioeroding sponge infestation visualised with micro-
computed tomography. Facies 53, 157e176.
Blaxter, M., Koutsovoulos, G., 2015. The evolution of parasitism in Nematoda. Parasitology
142 (Suppl. 1), S26eS39.
Blaxter, M.L., 2003. Nematoda: genes, genomes and the evolution of parasitism. Adv. Para-
sitol. 54, 101e195.
Bomfleur, B., Kerp, H., Taylor, T.N., Moestrup, Ø., Taylor, E.L., 2012. Triassic leech cocoon
from Antarctica contains fossil bell animal. Proc. Natl. Acad. Sci. U.S.A. 109, 20971e20974.
Bos, K.I., Jager, G., Schuenemann, V.J., Vagene, A.J., Spyrou, M.A., Herbig, A., Nieselt, K.,
Krause, J., 2015. Parallel detection of ancient pathogens via array-based DNA capture.
Philos. Trans. R. Soc. London B Biol. Sci. 370, 20130375.
Bouchet, F., Ara ujo, A., Harter, S., Chaves, S.M., Duarte, A.N., Monnier, J.L., Ferreira, L.F.,
2003. Toxocara canis (Werner, 1782) eggs in the Pleistocene site of Menez-Dregan,
France (300,000-500,000 years before present). Mem. Inst. Oswaldo Cruz 98, 137e139.
Boucot, A.J., 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier,
Amsterdam.
Boucot, A.J., Poinar, G., 2010. Host-parasite and host-parasitoid relationships and disease. In:
Boucot, A.J., Poinar, G.O. (Eds.), Fossil Behavior Compendium. CRC Press, Boca
Raton, Florida, pp. 27e71.
Boyko, C.B., Williams, J.D., 2009. Crustacean parasites as phylogenetic indicators in decapod
evolution. In: Martin, J.W., Crandall, K.A., Felder, D.L. (Eds.), Decapod Crustacean
Phylogenetics, pp. 197e220.
Brasier, M.D., Antcliffe, J.B., 2008. Dickinsonia from Ediacara: a new look at morphology and
body construction. Palaeogeogr. Palaeoclimatol. Palaeoecol. 270, 311e323.
Brett, C.E., 1978. Host-specific pit-forming epizoans on Silurian crinoids. Lethaia 11, 217e232.
Briggs, D.E., Summons, R.E., 2014. Ancient biomolecules: their origins, fossilization, and
role in revealing the history of life. Bioessays 36, 482e490.
Briggs, D.E.G., 2013. A mosquito’s last supper reminds us not to underestimate the fossil
record. Proc. Natl. Acad. Sci. U.S.A. 110, 18353e18354.
Brodniewicz, I., 1968. On glochidia of the genera Unio and Anodonta from the Quaternary
freshwater sediments of Poland. Acta Palaeontologica Polonica 13, 619e628.
Bromham, L., Penny, D., 2003. The modern molecular clock. Nat. Rev. Genet. 4, 216e224.
Bronstein, J.L., 1994. Conditional outcomes in mutualistic interactions. Trends Ecol. Evol. 9,
214e217.
Brooks, D.R., McLennan, D.A., 1993. Parascript: Parasites and the Language of Evolution.
Smithsonian Institution Press, Washington.
Bush, A.O., Fernandez, J.C., Esch, G.W., Seed, J.R., 2001. Parasitism: The Diversity and
Ecology of Animal Parasites. Cambridge University Press, Cambridge.
Campbell, D., 1985. The life cycle of Gymnophallus rebecqui (Digenea: Gymnophallidae) and
the response of the bivalve Abra tenuis to its metacercariae. J. Mar. Biol. Assoc. U.K. 65,
589e601.
Cano, R.J., Rivera-Perez, J., Toranzos, G.A., Santiago-Rodriguez, T.M., Narganes-
Storde, Y.M., Chanlatte-Baik, L., Garcia-Roldan, E., Bunkley-Williams, L.,
Massey, S.E., 2014. Paleomicrobiology: revealing fecal microbiomes of ancient indige-
nous cultures. PLoS One 9, e106833.
Cappellini, E., Collins, M.J., Gilbert, M.T.P., 2014. Unlocking ancient protein palimpsests.
Science 343, 1320e1322.
Cappellini, E., Jensen, L.J., Szklarczyk, D., Ginolhac, A., Da Fonseca, R.a.R., Stafford, T.W.,
Holen, S.R., Collins, M.J., Orlando, L., Willerslev, E., Gilbert, M.T.P., Olsen, J.V.,
2012. Proteomic analysis of a pleistocene mammoth femur reveals more than one hun-
dred ancient bone proteins. J. Proteome Res. 11, 917e926.
38 Kenneth De Baets and D. Timothy J. Littlewood

Castellani, C., Maas, A., Waloszek, D., Haug, J.T., 2011. New pentastomids from the Late
Cambrian of Sweden - deeper insight of the ontogeny of fossil tongue worms. Palaeon-
tographica Abteilung A 293, 95e145.
Cerveau, N., Leclercq, S., Leroy, E., Bouchon, D., Cordaux, R., 2011. Short- and long-term
evolutionary dynamics of bacterial insertion sequences: insights from Wolbachia
endosymbionts. Genome Biol. Evol. 3, 1175e1186.
Charleston, N.A., Perkins, S.L., 2006. Traversing the tangle: algorithms and applications for
cophylogenetic studies. J. Biomed. Inf. 39, 62e71.
Chen, J., Wang, B., Engel, M.S., Wappler, T., Jarzembowski, E.A., Zhang, H., Wang, X.,
Zheng, X., Rust, J., 2014. Extreme adaptations for aquatic ectoparasitism in a Jurassic fly
larva. eLife 3, e02844.
Cheng, T., Combes, C., 1990. Influence of environmental factors on the invasion of molluscs
by parasites: with special reference to Europe. In: Di Castri, F., Hansen, A.J.,
Debussche, M. (Eds.), Biological Invasions in Europe and the Mediterranean Basin.
Springer, Netherlands, pp. 307e332.
Clarke, J.T., Warnock, R.C.M., Donoghue, P.C.J., 2011. Establishing a time-scale for plant
evolution. New Phytol. 192, 266e301.
Collins, M.J., Gernaey, A.M., Nielsen-Marsh, C.M., Vermeer, C., Westbroek, P., 2000.
Slow rates of degradation of osteocalcin: green light for fossil bone protein? Geology
28, 1139e1142.
Conway Morris, S., 1981. Parasites and the fossil record. Parasitology 83, 489e509.
Conway Morris, S., Crompton, D.W.T., 1982. The origins and evolution of the
Acanthocephala. Biol. Rev. 57, 85e115.
Cressey, R., Boxshall, G., 1989. Kabatarina pattersoni, a fossil parasitic copepod (Dichelesthii-
dae) from a Lower Cretaceous fish. Micropaleontology 35, 150e167.
Cressey, R., Patterson, C., 1973. Fossil parasitic copepods from a Lower Cretaceous fish.
Science 180, 1283e1285.
Crisp, M.D., Trewick, S.A., Cook, L.G., 2011. Hypothesis testing in biogeography. Trends
Ecol. Evol. 26, 66e72.
Cunningham, J.A., Donoghue, P.C.J., Bengtson, S., 2014a. Distinguishing biology from ge-
ology in soft-tissue preservation. In: Laflamme, M., Schiffbauer, J.D., Darroch, S.a.F.
(Eds.), Reading and Writing of the Fossil Record: Preservational Pathways to Excep-
tional FossilIzation, Paleontol Soc Papers, vol. 20, pp. 275e287.
Cunningham, J.A., Rahman, I.A., Lautenschlager, S., Rayfield, E.J., Donoghue, P.C.,
2014b. A virtual world of paleontology. Trends Ecol. Evol. 29, 347e357.
Da Silva, P.A., Borba, V.H., Dutra, J.M.F., Leles, D., Da-Rosa, A.a.S., Ferreira, L.F.,
Araujo, A., 2014. A new ascarid species in cynodont coprolite dated of 240 million
years. Anais da Academia Brasileira de Ciências 86, 265e270.
Dalgleish, R.C., Palma, R.L., Price, R.D., Smith, V.S., 2006. Fossil lice (Insecta: Phthirap-
tera) reconsidered. Syst. Entomol. 31, 648e651.
Dampf, A., 1911. Palaeopsylla klebsiana n. sp: ein fossiler Floh aus dem baltischen Bernstein.

Schriften der Physikalisch-Okonomischen Gesellschaft zu K€ onigsberg 51, 248e249.
Danchin, E.G.J., Rosso, M.-N., Vieira, P., De Almeida-Engler, J., Coutinho, P.M.,
Henrissat, B., Abad, P., 2010. Multiple lateral gene transfers and duplications have promoted
plant parasitism ability in nematodes. Proc. Natl. Acad. Sci. U.S.A. 107, 17651e17656.
Daskin, J.H., Alford, R.A., 2012. Context-dependent symbioses and their potential roles in
wildlife diseases. Proc. R. Soc. B 279, 1457e1465.
Davis, C.C., Xi, Z., 2015. Horizontal gene transfer in parasitic plants. Curr. Opin. Plant Biol.
26, 14e19.
De Baets, K., Dentzien-Dias, P.C., Upeniece, I., Verneau, O., Donoghue, P.C.J., 2015a.
Constraining the deep origin of parasitic flatworms and host-interactions with fossil
evidence. Adv. Parasitol. 90, 93e135.
The Importance of Fossils in the Evolution of Parasites 39

De Baets, K., Keupp, H., Klug, C., 2015b. Parasites of ammonoids. In: Klug, C., Korn, D.,
De Baets, K., Kruta, I., Mapes, R.H. (Eds.), Ammonoid Paleobiology: From Anatomy to
Paleoecology. Springer, The Netherlands.
De Baets, K., Donoghue, P.C.J., 2012. Molecular clocks and tectonic blocks. Geol. Soc. Am.
Abstr. 44, 331.
De Baets, K., Klug, C., Korn, D., 2011. Devonian pearls and ammonoid-endoparasite co-
evolution. Acta Palaeontologica Polonica 56, 159e180.
Dentzien-Dias, P.C., Poinar, G., De Figueiredo, A.E.Q., Pacheco, A.C.L., Horn, B.L.D.,
Schultz, C.L., 2013. Tapeworm eggs in a 270 million-year-old shark coprolite. PLoS
One 8, e55007.
Dittmar, K., 2009. Old parasites for a new world: the future of paleoparasitological research.
A review. J. Parasitol. 95, 365e371.
Dittmar, K., 2010. Palaeogeography of parasites. In: Morand, S., Krasnov, B.R. (Eds.), The
Biogeography of Host-parasite Interactions. Oxford University Press, Oxford.
Dittmar, K., 2014. Paleoparasitology and ancient DNA. In: Ferreira, L.F., Reinhard, K.J.,
Araujo, A. (Eds.), Foundations of Paleoparasitology. Editora Fiocruz/International
Federation of Tropical Medicine, Rio de Janeiro.
Dittmar, K., Araujo, A., Reinhard, K.J., 2011. The study of parasites through time: archae-
oparasitology and paleoparasitology. In: Grauer, A.L. (Ed.), A Companion to
Paleopathology. Wiley-Blackwell, Oxford, UK.
Dockery, D.T., 1980. The Invertebrate Macropaleontology of the Clarke County, Missis-
sippi, Area. Mississippi Department of Natural Resources. Bureau of Geology Bulletin,
122, pp. 1e387.
Donoghue, P.C.J., Bengtson, S., Dong, X.-P., Gostling, N.J., Huldtgren, T.,
Cunningham, J.A., Yin, C., Yue, Z., Peng, F., Stampanoni, M., 2006. Synchrotron
X-ray tomographic microscopy of fossil embryos. Nature 442, 680e683.
Donoghue, P.C.J., Dong, X.-P., 2005. Embryos and ancestors. In: Briggs, D.E.G. (Ed.),
Evolving Form and Function: Fossils and Development. Yale Peabody Museum of
Natural History, Yale University, New Haven, pp. 81e99.
Donovan, S.K., 2015. A prejudiced review of ancient parasites and their host echinoderms:
CSI fossil record or just an excuse for speculation? Adv. Parasitol. 90, 291e328.
Dorris, M., De Ley, P., Blaxter, M.L., 1999. Molecular analysis of nematode diversity and the
evolution of parasitism. Parasitol. Today 15, 188e193.
Duan, B., Dong, X.-P., Donoghue, P.C.J., 2012. New palaeoscolecid worms from the
Furongian (upper Cambrian) of Hunan, South China: is Markuelia an embryonic
palaeoscolecid? Palaeontology 55, 613e622.
Dufour, B., Le Bailly, M., 2013. Testing new parasite egg extraction methods in paleopara-
sitology and an attempt at quantification. Int. J. Paleopathol. 3, 199e203.
Dunlop, J.A., Wirth, S., Penney, D., Mcneil, A., Bradley, R.S., Withers, P.J., Preziosi, R.F.,
2012. A minute fossil phoretic mite recovered by phase-contrast X-ray computed
tomography. Biol. Lett. http://dx.doi.org/10.1098/rsbl.2011.0923.
Dunn, R.R., Harris, N.C., Colwell, R.K., Koh, L.P., Sodhi, N.S., 2009. The sixth mass
coextinction: are most endangered species parasites and mutualists? Proc. R. Soc. B
276, 3037e3045.
Duron, O., 2013. Lateral transfers of insertion sequences between Wolbachia, Cardinium and
Rickettsia bacterial endosymbionts. Heredity 111, 330e337.
Dutour, O., 2013. Paleoparasitology and paleopathology. Synergies for reconstructing
the past of human infectious diseases and their pathocenosis. Int. J. Paleopathol. 3,
145e149.
Edgecombe, G., Giribet, G., Dunn, C., Hejnol, A., Kristensen, R., Neves, R., Rouse, G.,
Worsaae, K., Sørensen, M., 2011. Higher-level metazoan relationships: recent progress
and remaining questions. Org. Div. Evol. 11, 151e172.
40 Kenneth De Baets and D. Timothy J. Littlewood

Egger, B., Lapraz, F., Tomiczek, B., M€ 


uller, S., Dessimoz, C., Girstmair, J., Skunca, N.,
Rawlinson, K.A., Cameron, C.B., Beli, E., 2015. A transcriptomic-phylogenomic
analysis of the evolutionary relationships of flatworms. Curr. Biol. 25, 1347e1353.
Engel, M.S., 2008. A stem-group cimicid in mid-Cretaceous amber from Myanmar
(Hemiptera: Cimicoidea). Alavesia 2, 233e237.
Faulwetter, S., Vasileiadou, A., Kouratoras, M., Dailianis, T., Arvanitidis, C., 2013. Micro-
computed tomography: Introducing new dimensions to taxonomy. ZooKeys 1e45.
Feldmann, R.M., 1998. Parasitic Castration of the crab, tumidocarcinus giganteus glaessner,
from the miocene of New Zealand: coevolution within the Crustacea. J. Paleontol.
72, 493e498.
Feldmann, R.M., 2003. Interpreting ecology and physiology of fossil decapod crustaceans.
Contr. Zool. 72, 111e117.
Ferreira, L., Araujo, A., Duarte, A., 1993. Nematode larvae in fossilized animal coprolites
from lower and middle Pleistocene sites, Central Italy. J. Parasitol. 440e442.
Ferreira, L.F., Araujo, A., Confalonieri, U., Chame, M., Gomes, D.C., 1991. Trichuris eggs
in animal coprolites dated from 30,000 years ago. J. Parasitol. 77, 491e493.
Ferreira, L.F., Araujo, A., Confalonieri, U., Chame, M., Ribeiro, B., 1992. Eimeria oocysts
in deer coprolites dated from 9,000 years BP. Mem. Inst. Oswaldo Cruz 87, 105e106.
Foote, M., Miller, A.I., 2007. Principles of Paleontology, Third edition. W. H. Freeman,
New York.
Franzen, C., 1974. Epizoans on Silurian-Devonian crinoids. Lethaia 7, 287e301.
Frías, L., Leles, D., Araujo, A., 2013. Studies on protozoa in ancient remains - a review.
Mem. Inst. Oswaldo Cruz 108, 1e12.
Fry, G.F., 1977. Analysis of prehistoric coprolites from Utah. Anthropol. Pap. 97, 1e45.
Fry, G.F., Hall, H., 1969. Parasitological examination of prehistoric human coprolites from
Utah. Proc. Utah Acad. Sci. Art. Lett. 102e105.
Fugassa, M.H., Sardella, N.H., Taglioretti, V., Reinhard, K.J., Ara
ujo, A., 2008. Eimeriid oo-
cysts from archaeological samples in Patagonia, Argentina. J. Parasitol. 94, 1418e1420.
Gao, T.-P., Shih, C.-K., Xu, X., Wang, S., Ren, D., 2012. Mid-Mesozoic flea-like ectopar-
asites of feathered or haired vertebrates. Curr. Biol. 22, 732e735.
Gao, T., Shih, C., Rasnitsyn, A.p., Xu, X., Wang, S., Ren, D., 2013. New transitional fleas
from China highlighting diversity of Early Cretaceous ectoparasitic insects. Curr. Biol.
23, 1261e1266.
Gao, T., Shih, C., Rasnitsyn, A.P., Xu, X., Wang, S., Ren, D., 2014. The first flea with fully
distended abdomen from the Early Cretaceous of China. BMC Evol. Biol. 14, 168.
Garwood, R.J., Dunlop, J., 2014. Three-dimensional reconstruction and the phylogeny of
extinct chelicerate orders. PeerJ 2, e641.
Gibson, D.I., Bray, R.A., Powell, C.B., 1987. Aspects of the life history and origins of Neso-
lecithus africanus (Cestoda: Amphilinidea). J. Nat. Hist. 21, 785e794.
Gifford, R.J., 2012. Viral evolution in deep time: lentiviruses and mammals. Trends Genet.
28, 89e100.
Gilbert, C., Feschotte, C., 2010. Genomic fossils calibrate the long-term evolution of
hepadnaviruses. PLoS Biol. 8, e1000495.
Gilbert, C., Schaack, S., Feschotte, C., 2010. Quand les éléments génétiques mobiles bon-
dissent entre especes animales [Mobile elements jump between parasites and vertebrate
hosts]. Med. Sci. Paris 26, 1025e1027.
Gitschier, J., 2015. The philosophical approach: an interview with Ford Doolittle. PLoS
Genet. 11, e1005173.
Gladyshev, E.A., Meselson, M., Arkhipova, I.R., 2008. Massive horizontal gene transfer in
bdelloid rotifers. Science 320, 1210e1213.
Goldring, R., 1999. Field Palaeontology, Second Edition. Longman, Harlow.
The Importance of Fossils in the Evolution of Parasites 41

Gonçalves, M.L.C., Ara ujo, A., Ferreira, L.F., 2003. Human intestinal parasites in the past:
new findings and a review. Mem. Inst. Oswaldo Cruz 98, 103e118.
Goswami, A., Upchurch, P., 2010. The dating game: a reply to Heads (2010). Zool. Scr. 39,
406e409.
Greenwalt, D.E., Goreva, Y.S., Siljestr€ om, S.M., Rose, T., Harbach, R.E., 2013. Hemoglo-
bin-derived porphyrins preserved in a Middle Eocene blood-engorged mosquito. Proc.
Natl. Acad. Sci. U.S.A. 110, 18496e18500.
Hartikainen, H., Gruhl, A., Okamura, B., 2014. Diversification and repeated morphological
transitions in endoparasitic cnidarians (Myxozoa: Malacosporea). Mol. Phylogenet. Evol.
76, 261e269.
Hayami, I., Kanie, Y., 1980. Mode of life of a giant capulid gastropod from the Upper Creta-
ceous of Saghalien and Japan. Palaeontology 23, 689e698.
Heath, T.A., Huelsenbeck, J.P., Stadler, T., 2014. The fossilized birthedeath process for
coherent calibration of divergence-time estimates. Proc. Natl. Acad. Sci. U.S.A. 111,
E2957eE2966.
Hebsgaard, M.B., Phillips, M.J., Willerslev, E., 2005. Geologically ancient DNA: fact or
artefact? Trends Microbiol. 13, 212e220.
Héritier, L., Badets, M., Du Preez, L.H., Aisien, M.S.O., Lixian, F., Combes, C.,
Verneau, O., 2015. Evolutionary processes involved in the diversification of chelonian
and mammal polystomatid parasites (Platyhelminthes, Monogenea, Polystomatidae)
revealed by palaeoecology of their hosts. Mol. Phylogenet. Evol. 92, 1e10.
Herniou, E.A., Huguet, E., Thézé, J., Bézier, A., Periquet, G., Drezen, J.-M., 2013. When
parasitic wasps hijacked viruses: genomic and functional evolution of polydnaviruses.
Philos. Trans. R. Soc. B 368, 20130051.
Hess, H., 2010. Myzostome deformation on arms of the Early Jurassic crinoid Balanocrinus
gracilis (Charlesworth). J. Paleontol. 84, 1031e1034.
Hibbett, D., Matheny, P.B., 2009. The relative ages of ectomycorrhizal mushrooms and their
plant hosts estimated using Bayesian relaxed molecular clock analyses. BMC Biol. 7, 13.
Hill, R.C., Wither, M.J., Nemkov, T., Barrett, A., D’Alessandro, A., Dzieciatkowska, M.,
Hansen, K.C., 2015. Preserved proteins from extinct Bison latifrons identified by tandem
mass spectrometry; hydroxylysine glycosides are a common feature of ancient collagen.
Mol. Cell. Proteomics 14, 1946e1958.
Hipsley, C.A., M€ uller, J., 2014. Beyond fossil calibrations: realities of molecular clock prac-
tices in evolutionary biology. Front. Genet. 5, 138.
Hirt, R.P., Alsmark, C., Embley, T.M., 2015. Lateral gene transfers and the origins of the
eukaryote proteome: a view from microbial parasites. Curr. Opin. Microbiol. 23,
155e162.
Ho, S.Y., 2014. The changing face of the molecular evolutionary clock. Trends Ecol. Evol.
29, 496e503.
Ho, S.Y.W., Phillips, M.J., 2009. Accounting for calibration uncertainty in phylogenetic esti-
mation of evolutionary divergence times. Syst. Biol. 58, 367e380.
Hofreiter, M., Collins, M., Stewart, J.R., 2012. Ancient biomolecules in Quaternary
palaeoecology. Quat. Sci. Rev. 33, 1e13.
Hofreiter, M., Paijmans, J.L., Goodchild, H., Speller, C.F., Barlow, A., Fortes, G.G.,
Thomas, J.A., Ludwig, A., Collins, M.J., 2015. The future of ancient DNA: technical
advances and conceptual shifts. Bioessays 37, 284e293.
Huang, D., 2014. The diversity and host associations of Mesozoic giant fleas. Natl. Sci. Rev.
http://dx.doi.org/10.1093/nsr/nwu022.
Huang, D., 2015. Tarwinia australis (Siphonaptera: Tarwiniidae) from the Lower Cretaceous
Koonwarra fossil bed: morphological revision and analysis of its evolutionary
relationship. Cretaceous Res. B 52, 507e515.
42 Kenneth De Baets and D. Timothy J. Littlewood

Huang, D., Engel, M., Cai, C., Nel, A., 2013. Mesozoic giant fleas from northeastern China
(Siphonaptera): taxonomy and implications for palaeodiversity. Chin. Sci. Bull. 58,
1682e1690.
Huang, D., Engel, M.S., Cai, C., Wu, H., Nel, A., 2012. Diverse transitional giant fleas from
the Mesozoic era of China. Nature 483, 201e204.
Hughes, D.P., Wappler, T., Labandeira, C.C., 2010. Ancient death-grip leaf scars reveal ante
fungal parasitism. Biol. Lett. http://dx.doi.org/10.1098/rsbl.2010.0521.
Hugot, J.-P., Gardner, S., Borba, V., Araujo, P., Leles, D., Stock Da-Rosa, A., Dutra, J.,
Ferreira, L., Araujo, A., 2014. Discovery of a 240 million year old nematode parasite
egg in a cynodont coprolite sheds light on the early origin of nematode parasites in
vertebrates. Parasites Vectors 7, 486.
Hunt, A.P., Lucas, S.G., Milan, J., Spielmann, J.A., 2012. Vertebrate coprolite studies: status
and prospectus. N. M. Mus. Nat. Hist. Sci. Bull. 57, 5e24.
Huntley, J.W., 2007. Towards establishing a modern baseline for paleopathology: trace-
producing parasites in a bivalve host. J. Shellfish Res. 26, 253e259.
Huntley, J.W., De Baets, K., 2015. Trace fossil evidence of trematodeebivalve parasiteehost
interactions in deep time. Adv. Parasitol. 90, 201e231.
Huntley, J.W., Kowalewski, M., 2007. Strong coupling of predation intensity and diversity
in the Phanerozoic fossil record. Proceedings of the National Academy of Sciences 104,
15006e15010.
Huntley, J.W., Fursich, F.T., Alberti, M., Hethke, M., Liu, C., 2014. A complete Holocene
record of trematode-bivalve infection and implications for the response of parasitism to
climate change. Proc. Natl. Acad. Sci. U.S.A. 111, 18150e18155.
Ituarte, C., Cremonte, F., Zelaya, D.G., 2005. Parasite-mediated shell alterations in recent
and Holocene sub-Antarctic bivalves: the parasite as modeler of host reaction. Invert.
Biol. 124, 220e229.
Ituarte, C., Cremonte, F., Deferrari, G., 2001. Mantle-shell complex reactions elicited by
digenean metacercariae in Gaimardia trapesina (Bivalvia : Gaimardiidae) from the South-
western Atlantic Ocean and Magellan Strait. Dis. Aquat. Org. 48, 47e56.
Jiménez-Guri, E., Philippe, H., Okamura, B., Holland, P.W.H., 2007. Buddenbrockia is a
cnidarian worm. Science 317, 116e118.
Johnson, K.L., Reinhard, K.J., Sianto, L., Araujo, A., Gardner, S.L., Janovy, J., 2008. A tick
from a prehistoric arizona coprolite. J. Parasitol. 94, 296e298.
Johnson, P.T., Lunde, K.B., Haight, R.W., Bowerman, J., Blaustein, A.R., 2001. Ribeiroia
ondatrae (Trematoda: Digenea) infection induces severe limb malformations in western
toads (Bufo boreas). Can. J. Zool. 79, 370e379.
Johnson, P.T.J., Lunde, K.B., Thurman, E.M., Ritchie, E.G., Wray, S.N., Sutherland, D.R.,
Kapfer, J.M., Frest, T.J., Bowerman, J., Blaustein, A.R., 2002. Parasite (Ribeiroia ondatrae)
infection linked to amphibian malformations in the western United States. Ecol. Monog.
72, 151e168.
Johnson, P.T.J., Lunde, K.B., Zelmer, D.A., Werner, J.K., 2003. Limb deformities as an
emerging parasitic disease in amphibians: evidence from museum specimens and resurvey
data. Conserv. Biol. 17, 1724e1737.
Johnson, P.T.J., Sutherland, D.R., 2003. Amphibian deformities and Ribeiroia infection: an
emerging helminthiasis. Trends Parasitol. 19, 332e335.
Jouy-Avantin, F., Combes, C., Lumley, H., Miskovsky, J.C., Moné, H., 1999. Helminth
eggs in animal coprolites from a Middle Pleistocene site in Europe. J. Parasitol. 85,
376e379.
Katzourakis, A., 2013. Paleovirology: inferring viral evolution from host genome sequence
data. Philos. Trans. R. Soc. B 368, 20120493.
Katzourakis, A., Gifford, R.J., 2010. Endogenous viral elements in animal genomes. PLoS
Genet. 6, e1001191.
The Importance of Fossils in the Evolution of Parasites 43

Kelley, P.H., Kowalewski, M., Hansen, T.A. (Eds.), 2003. Predator-prey Interactions in the
Fossil Record. Topics in Geobiology, vol. 20. Springer Science & Business Media.
Keupp, H., 2012. Atlas zur Pal€aopathologie der Cephalopoden. Berliner Pal€aobiologische
Abhandlungen 12, 1e392.
Kidwell, M.G., Lisch, D.R., 2001. Perspective: transposable elements, parasitic DNA, and
genome evolution. Evolution 55, 1e24.
Klompmaker, A.A., Artal, P., Van Bakel, B.W.M., Fraaije, R.H.B., Jagt, J.W.M., 2014. Par-
asites in the Fossil Record: a Cretaceous fauna with isopod-infested decapod crustaceans,
infestation patterns through time, and a new ichnotaxon. PLoS One 9, e92551.
Klompmaker, A.A., Boxshall, G., 2015. Fossil crustaceans as parasites and hosts. Adv.
Parasitol. 90, 233e289.
Knor, S., Skuhrava, M., Wappler, T., Prokop, J., 2013. Galls and gall makers on plant leaves
from the lower Miocene (Burdigalian) of the Czech Republic: systematic and palaeoe-
cological implications. Rev. Palaeobot. Palynol. 188, 38e51.
Kodandaramaiah, U., 2011. Tectonic calibrations in molecular dating. Curr. Zool. 57,
116e124.
Kouchinsky, A., Bengtson, S., Runnegar, B., Skovsted, C., Steiner, M., Vendrasco, M.,
2012. Chronology of early Cambrian biomineralization. Geol. Mag. 149, 221e251.
Koutsovoulos, G., Makepeace, B., Tanya, V.N., Blaxter, M., 2014. Palaeosymbiosis revealed
by genomic fossils of Wolbachia in a strongyloidean nematode. PLoS Genet. 10,
e1004397.
Labandeira, C.C., 2002. Paleobiology of predators, parasitoids, and parasites: death and
accomodation in the fossil record of continental invertebrates. Paleontol. Soc. Pap. 8,
211e250.
Labandeira, C.C., Currano, E.D., 2013. The fossil record of plant-insect dynamics. Ann.
Rev. Earth Planet Sci. 41, 287e311.
Laumer, C.E., Hejnol, A., Giribet, G., 2015. Nuclear genomic signals of the ‘microturbellar-
ian’ roots of platyhelminth evolutionary innovation. eLife 4, e05503.
Leckey, E.H., Smith, D.M., 2015. Host fidelity over geologic time: restricted use of oaks by
oak gallwasps. J. Paleontol. 89, 236e244.
Lee, A., Nolan, A., Watson, J., Tristem, M., 2013. Identification of an ancient endogenous
retrovirus, predating the divergence of the placental mammals. Philos. Trans. R. Soc. B
368, 20120503.
Leles, D., Reinhard, K.J., Fugassa, M., Ferreira, L.F., I~ niguez, A.M., Ara ujo, A., 2010. A
parasitological paradox: why is ascarid infection so rare in the prehistoric Americas?
J. Archaeol. Sci. 37, 1510e1520.
Liljedahl, L., 1985. Ecological aspects of a silicified bivalve fauna from the Silurian of Gotland.
Lethaia 18, 53e66.
Linseele, V., Riemer, H., Baeten, J., De Vos, D., Marinova, E., Ottoni, C., 2013. Species
identification of archaeological dung remains: a critical review of potential methods. En-
viron. Archaeol. 18, 5e17.
Littlewood, D.T., Waeschenbach, A., 2015. Evolution: a turn up for the worms. Curr. Biol.
25, R457eR460.
Littlewood, D.T.J., 2006. The evolution of parasitism in flatworms. In: Maule, A.G.,
Marks, N.J. (Eds.), Parasitic Flatworms: Molecular Biology, Biochemistry, Immunology
and Physiology. CABI, Wallingford, pp. 96e123.
Littlewood, D.T.J., 2011. Systematics as a cornerstone of parasitology: overview and preface.
Parasitology 138, 1633e1637.
Littlewood, D.T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
Littlewood, D.T.J., Rohde, K., Bray, R.A., Herniou, E.A., 1999. Phylogeny of the Platyhel-
minthes and the evolution of parasitism. Biol. J. Linn. Soc. 68, 257e287.
44 Kenneth De Baets and D. Timothy J. Littlewood

Loader, S.P., Pisani, D., Cotton, J.A., Gower, D.J., Day, J.J., Wilkinson, M., 2007. Relative
time scales reveal multiple origins of parallel disjunct distributions of African caecilian
amphibians. Biol. Lett. 3, 505e508.
Lockyer, A.E., Olson, P.D., Littlewood, D.T.J., 2003. Utility of complete large and small
subunit rRNA genes in resolving the phylogeny of the Neodermata (Platyhel-
minthes): implications and a review of the cercomer theory. Biol. J. Linn. Soc. 78,
155e171.
Loss-Oliveira, L., Aguiar, B.O., Schrago, C.G., 2012. Testing synchrony in historical bioge-
ography: the case of new world primates and Hystricognathi rodents. Evol. Bioinform. 8,
127e137.
Lozouet, P., Renard, P., 1998. Les Coralliophilidae, Gastropoda de l’Oligocene et du
Miocene inférieur d’Aquitaine (Sud-Ouest de la France): Systématique et coraux
h^otes. Geobios 31, 171e185.
Lukashevich, E.D., Mostovski, M.B., 2003. Hematophagous insects in the Fossil record.
Paleontol. J. 37, 153e161.
Maas, A., Braun, A., Dong, X.-P., Donoghue, P.C.J., M€ uller, K.J., Olempska, E.,
Repetski, J.E., Siveter, D.J., Stein, M., Waloszek, D., 2006. The ‘Orsten’ e More
than a cambrian Konservat-Lagerst€atte yielding exceptional preservation. Palaeoworld
15, 266e282.
Maas, A., Waloszek, D., 2001. Cambrian derivatives of the early arthropod stem lineage, pen-
tastomids, tardigrades and lobopodians an ‘Orsten’Perspective. Zoologischer Anzeiger
240, 451e459.
Madsen, F., Wolff, T., 1965. Evidence of the occurrence of Ascothoracica (parasitic cirripeds)
in Upper Cretaceous. Meddelelser fra Dansk Geologisk Forening 15, 556e558.
Manum, S.B., Bose, M.N., Sawyer, R.T., 1991. Clitellate cocoons in freshwater deposits
since the Triassic. Zool. Scripta 20, 347e366.
Martill, D.M., Davis, P.G., 1998. Did dinosaurs come up to scratch? Nature 396, 528e529.
McConnell, S.M., Zavada, M.S., 2013. The occurrence of an abdominal fauna in an articu-
lated tapir (Tapirus polkensis) from the Late Miocene Gray Fossil Site, northeastern
Tennessee. Integr. Zool. 8, 74e83.
Mejía-Madrid, H.H., 2013. Parascript, parasites and historical biogeography. In: Silva-
Opps, M. (Ed.), Advances in Biology. InTech, Rijeka, pp. 125e148.
Mello, B., Schrago, C.G., Mello, B., Schrago, C.G., 2014. Assignment of calibration infor-
mation to deeper phylogenetic nodes is more effective in obtaining precise and accurate
divergence time estimates. Evol. Bioinform. 10, 79e85.
Mercier, J., 1936. Zoothylacies d’E chinide fossile provoquées par un Crustacé: Castexia dou-
villei nov. gen., nov. sp. Bulletin de la Société Géologique de France, sér 5, 149e154.
Mietchen, D., Aberhan, M., Manz, B., Hampe, O., Mohr, B., Neumann, C., Volke, F.,
2008. Three-dimensional magnetic resonance imaging of fossils across taxa. Bio-
geosciences 5, 25e41.
Mitchell, P.D., 2015. Human parasites in medieval Europe: lifestyle, sanitation and medical
treatment. Adv. Parasitol. 90, 389e420.
Moore, R.B., Obornik, M., Janouskovec, J., Chrudimsky, T., Vancova, M., Green, D.H.,
Wright, S.W., Davies, N.W., Bolch, C.J.S., Heimann, K., Slapeta, J., Hoegh-
Guldberg, O., Logsdon, J.M., Carter, D.A., 2008. A photosynthetic alveolate closely
related to apicomplexan parasites. Nature 451, 959e963.
Mostowy, R., Engelst€adter, J., 2012. Hosteparasite coevolution induces selection for condi-
tion-dependent sex. J. Evol. Biol. 25, 2033e2046.
Mower, J.P., Stefanovic, S., Young, G.J., Palmer, J.D., 2004. Plant genetics: gene transfer
from parasitic to host plants. Nature 432, 165e166.
M€uller, J., Reisz, R.R., 2005. Four well-constrained calibration points from the vertebrate
fossil record for molecular clock estimates. BioEssays 27, 1069e1075.
The Importance of Fossils in the Evolution of Parasites 45

Muller, K., Walossek, D., Zakharov, A., 1995. ‘Orsten’type phosphatized soft-integument
preservation and a new record from the Middle Cambrian Kuonamka Formation in
Siberia. Neues Jahrbuch fur Geologie und Palaontologie-Abhandlungen 197, 101e118.
Nagler, C., Haug, J.T., 2015. From fossil parasitoids to vectors: insects as parasites and hosts.
Adv. Parasitol. 90, 137e200.
Near, T.J., 2002. Acanthocephalan phylogeny and the evolution of parasitism. Integr. Comp.
Biol. 42, 668e677.
Neumann, C., Wisshak, M., 2006. A foraminiferal parasite on the sea urchin echinocorys:
ichnological evidence from the Late Cretaceous (Lower Maastrichtian, Northern
Germany). Ichnos 13, 185e190.
Neumann, C., Wisshak, M., 2009. Gastropod parasitism on Late Cretaceous to Early Paleo-
cene holasteroid echinoids e Evidence from Oichnus halo isp. n. Palaeogeogr. Palaeocli-
matol. Palaeoecol. 284, 115e119.
Niewiadomska, A.M., Gifford, R.J., 2013. The extraordinary evolutionary history of the
reticuloendotheliosis viruses. PLoS Biol. 11, e1001642.
Noronha, D., Ferreira, L.F., Rangel, A., Araujo, A., Gomes, D.C., 1994. Echinopardalis sp.
(Acanthocephala, Oligacanthorhynchidae) eggs in felid coprolites dated from 9,000 years
before present, found in the Brazilian northeast. Mem. Inst. Oswaldo Cruz 89, 119e120.
Nowak, M.D., Smith, A.B., Simpson, C., Zwickl, D.J., 2013. A simple method for
estimating informative node age priors for the fossil calibration of molecular divergence
time analyses. PLoS One 8, e66245.
O’Grady, R.T., 1985. Ontogenetic sequences and the phylogenetics of parasitic flatworm life
cycles. Cladistics 1, 159e170.
Oakley, T.H., Wolfe, J.M., Lindgren, A.R., Zaharoff, A.K., 2012. Phylotranscriptomics to
bring the understudied into the fold: monophyletic Ostracoda, fossil placement and
pancrustacean phylogeny. Mol. Biol. Evol. 30, 215e233.
Okamoto, N., McFadden, G.I., 2008. The mother of all parasites. Future Microbiol. 3,
391e395.
Okamura, B., Gruhl, A., 2015. Myxozoan affinities and route to endoparasitism. In:
Okamura, B., Gruhl, A., Bartholomew, J.L. (Eds.), Myxozoan Evolution, Ecology
and Development. Springer International Publishing, pp. 23e44.
Okamura, B., Gruhl, A., Reft, A., 2015. Cnidarian origins of the myxozoa. In: Okamura, B.,
Gruhl, A., Bartholomew, J.L. (Eds.), Myxozoan Evolution, Ecology and Development.
Springer International Publishing, pp. 45e68.
Ostrom, P.H., Schall, M., Gandhi, H., Shen, T.-L., Hauschka, P.V., Strahler, J.R.,
Gage, D.A., 2000. New strategies for characterizing ancient proteins using matrix-assis-
ted laser desorption ionization mass spectrometry. Geochimica et Cosmochimica Acta
64, 1043e1050.
Owen, A.W., 1985. Trilobite abnormalities. Trans. R. Soc. Edin. Earth Sci. 76, 255e272.
Oyediran, A., Fajemisin, A., Abioye, A., Lagundoye, S., Olugbile, A., 1975. Infection of the
mastoid bone with a Paragonimus-like trematode. Am. J. Trop. Med. Hyg. 24, 268e273.
Page, R.D.M., 2003. Tangled Trees: Phylogeny, Cospeciation, and Coevolution. University
of Chicago Press, Chicago.
Parham, J.F., Donoghue, P.C.J., Bell, C.J., Calway, T.D., Head, J.J., Holroyd, P.A.,
Inoue, J.G., Irmis, R.B., Joyce, W.G., Ksepka, D.T., Patané, J.S.L., Smith, N.D.,
Tarver, J.E., Van Tuinen, M., Yang, Z., Angielczyk, K.D., Greenwood, J.M.,
Hipsley, C.A., Jacobs, L., Makovicky, P.J., M€ uller, J., Smith, K.T., Theodor, J.M.,
Warnock, R.C.M., Benton, M.J., 2012. Best practices for justifying fossil calibrations.
Syst. Biol. 61, 346e359.
Parker, G.A., Ball, M.A., Chubb, J.C., 2015. Evolution of complex life cycles in trophi-
cally transmitted helminths. I. Host incorporation and trophic ascent. J. Evol. Biol. 28,
267e291.
46 Kenneth De Baets and D. Timothy J. Littlewood

Parry, L., Tanner, A., Vinther, J., 2014. The origin of annelids. Palaeontology 57, 1091e1103.
Pe~
nalver, E., Pérez-De La Fuente, R., 2014. Unearthing the secrets of ancient immature
insects. eLife 3, e03443.
Penney, D., Mcneil, A., Green, D.I., Bradley, R.S., Jepson, J.E., Withers, P.J., Preziosi, R.F.,
2012. Ancient EphemeropteraeCollembola symbiosis fossilized in amber predicts
contemporary phoretic associations. PLoS One 7, e47651.
Perrichot, V., Beaucournu, J.-C., Velten, J., 2012. First extinct genus of a flea (Siphonaptera:
Pulicidae) in Miocene amber from the Dominican Republic. Zootaxa 54e61.
Pevzner, P.A., Kim, S., Ng, J., 2008. Comment on ‘Protein sequences from mastodon and
Tyrannosaurus rex revealed by mass spectrometry’. Science 321, 1040.
Poinar, G., 2003. A rhabdocoel turbellarian (Platyhelminthes, Typhloplanoida) in Baltic
amber with a review of fossil and sub-fossil platyhelminths. Invert. Biol. 122, 308e312.
Poinar, G., 2005a. Culex malariager, n. sp. (Diptera: Culicidae) from Dominican amber : the
first fossil mosquito vector of plasmodium. Proc. Entomol. Soc. Wash. 107, 548e553.
Poinar, G., 2005b. Plasmodium dominicana n. sp. (Plasmodiidae: Haemospororida) from
Tertiary Dominican amber. Syst. Parasitol. 61, 47e52.
Poinar, G., 2005c. Triatoma dominicana sp. n. (Hemiptera: Reduviidae: Triatominae), and
Trypanosoma antiquus sp. n. (Stercoraria: Trypanosomatidae), the first fossil evidence
of a triatomine-trypanosomatid vector association. Vector-Borne Zoonotic Dis. 5,
72e81.
Poinar, G., 2008a. Leptoconops nosopheris sp. n. (Diptera: Ceratopogonidae) and Paleotrypano-
soma burmanicus gen. n., sp. n. (Kinetoplastida: Trypanosomatidae), a biting midge -
trypanosome vector association from the Early Cretaceous. Mem. Inst. Oswaldo Cruz
103, 468e471.
Poinar, G., 2008b. Lutzomyia adiketis sp. n. (Diptera: Phlebotomidae), a vector of Paleoleish-
mania neotropicum sp. n. (Kinetoplastida: Trypanosomatidae) in Dominican amber. Para-
sites Vectors 1, 1e8.
Poinar, G., 2009. Description of an early Cretaceous termite (Isoptera: Kalotermitidae) and
its associated intestinal protozoa, with comments on their co-evolution. Parasites
Vectors 2, 12.
Poinar, G., 2011a. The Evolutionary History of Nematodes: As Revealed in Stone, Amber
and Mummies. Brill, Netherlands.
Poinar, G., 2011b. Vetufebrus ovatus n. gen., n. sp. (Haemospororida: Plasmodiidae) vectored
by a streblid bat fly (Diptera: Streblidae) in Dominican amber. Parasites Vectors 4, 229.
Poinar, G., 2012. Fossil gregarines in Dominican and Burmese amber: examples of acceler-
ated development? Palaeodiversity 5, 1e6.
Poinar, G., 2014a. Evolutionary history of terrestrial pathogens and endoparasites as revealed
in fossils and subfossils. Adv. Biol. 2014, 29.
Poinar, G., 2014b. Spirochaete-like cells in a Dominican amber Ambylomma tick (Arachnida:
Ixodidae). Hist. Biol. 27, 565e570.
Poinar, G., 2015a. The geological record of parasitic nematode evolution. Adv. Parasitol. 90,
53e92.
Poinar, G., 2015b. Rickettsial-like cells in the Cretaceous tick, Cornupalpatum burmanicum
(Ixodida: Ixodidae). Cretaceous Res. 52 (Part B), 623e627.
Poinar, G., Acra, A., Acra, F., 1994. Earliest fossil nematode (Mermithidae) in Cretaceous
Lebanese amber. Fundam. Appl. Nematol. 17, 475e477.
Poinar, G., Boucot, A.J., 2006. Evidence of intestinal parasites of dinosaurs. Parasitology 133,
245e249.
Poinar, G., Brown, A., 2003. A new genus of hard ticks in Cretaceous Burmese amber (Acari:
Ixodida: Ixodidae). Syst. Parasitol. 54, 199e205.
Poinar, G., Brown, A., 2012. The first fossil streblid bat fly, Enischnomyia stegosoma n. g., n. sp.
(Diptera: Hippoboscoidea: Streblidae). Syst. Parasitol. 81, 79e86.
The Importance of Fossils in the Evolution of Parasites 47

Poinar, G., Buckley, R., 2006. Nematode (Nematoda: Mermithidae) and hairworm
(Nematomorpha: Chordodidae) parasites in early cretaceous amber. J. Invert. Pathol.
93, 36e41.
Poinar, G., Buckley, R., 2008. Compluriscutula vetulum (Acari: Ixodida: Ixodidae), a new
genus and species of hard tick from Lower Cretaceous Burmese amber. Proc. Entomol.
Soc. Wash. 110, 445e450.
Poinar, G., Kerp, H., Hass, H., 2008. Palaeonema phyticum gen. n., sp. n. (Nematoda: Palae-
onematidae fam. n.), a Devonian nematode associated with early land plants. Nema-
tology 10, 9e14.
Poinar, G., Poinar, R., 2004a. Evidence of vector-borne disease of early Cretaceous reptiles.
Vector-Borne Zoonotic Dis. 4, 281e284.
Poinar, G., Poinar, R., 2004b. Paleoleishmania proterus n. gen., n. sp., (Trypanosomatidae:
Kinetoplastida) from Cretaceous Burmese Amber. Protist 155, 305e310.
Poinar, G., Telford, S.R., 2005. Paleohaemoproteus burmacis gen. n., sp. n. (Haemospororida:
Plasmodiidae) from an Early Cretaceous biting midge (Diptera: Ceratopogonidae).
Parasitology 131, 79e84.
Poulin, R., Morand, S., 2000. The diversity of parasites. Quart. Rev. Biol. 75, 277e293.
Pyron, R.A., 2011. Divergence time estimation using fossils as terminal taxa and the origins of
Lissamphibia. Syst. Biol. 60, 466e481.
Qian, Y., Yin, G., 1984. Zhijinitidae and its stratigraphical significance. Acta Palaeontol. Sin.
23, 215e223.
Radwanska, U., Poirot, E., 2010. Copepod-infested Bathonian (Middle jurassic) echinoids
from northern France. Acta Geol. Pol. 60, 549e555.
Radwa nska, U., Radwa nska, A., 2005. Myzostomid and copepod infestation of Jurassic
echinoderms: a general approach, some new occurrences, and/or re-interpretation of
previous reports. Acta Geol. Pol. 55, 109e130.
Rawlence, N.J., Lowe, D.J., Wood, J.R., Young, J.M., Churchman, G.J., Huang, Y.-T.,
Cooper, A., 2014. Using palaeoenvironmental DNA to reconstruct past environments:
progress and prospects. J. Quat. Sci. 29, 610e626.
Reinhard, K.J., Bryant, V.M., 1992. Coprolite Analysis: A Biological Perspective on Archae-
ology. Archaeological Method and Theory. Springer.
Retallack, G.J., 2007. Growth, decay and burial compaction of Dickinsonia, an iconic
Ediacaran fossil. Alcheringa 31, 215e240.
Richards, T.A., Soanes, D.M., Jones, M.D., Vasieva, O., Leonard, G., Paszkiewicz, K.,
Foster, P.G., Hall, N., Talbot, N.J., 2011. Horizontal gene transfer facilitated the evolu-
tion of plant parasitic mechanisms in the oomycetes. Proc. Natl. Acad. Sci. U.S.A. 108,
15258e15263.
Ronquist, F., Klopfstein, S., Vilhelmsen, L., Schulmeister, S., Murray, D.L., Rasnitsyn, A.P.,
2012. A total-evidence approach to dating with fossils, applied to the early radiation of
the Hymenoptera. Syst. Biol. 61, 973e999.
Rota-Stabelli, O., Daley, A.c, Pisani, D., 2013. Molecular timetrees reveal a Cambrian colo-
nization of land and a new scenario for ecdysozoan evolution. Curr. Biol. 23, 392e398.
Rothschild, B.M., Martin, L.D., 2006. Skeletal Impact of Disease. Museum of Natural
History & Science, New Mexico.
Rouse, G.W., 2005a. Fossil parasites. In: Rohde, K. (Ed.), Marine Parasitology. CABI
Publishing, Oxon, pp. 172e174.
Rouse, G.W., 2005b. Polychaeta (Bristle Worms). In: Rohde, K. (Ed.), Marine Parasitology.
CABI Publishing, Oxon, pp. 193e196.
Ruiz, G.M., 1991. Consequences of parasitism to marine invertebrates: host evolution? Am.
Zool. 31, 831e839.
Ruiz, G.M., Lindberg, D.R., 1989. A fossil record for trematodes: extent and potential uses.
Lethaia 22, 431e438.
48 Kenneth De Baets and D. Timothy J. Littlewood

Sanders, K.L., Lee, M.S.Y., 2010. Arthropod molecular divergence times and the Cambrian
origin of pentastomids. Syst. Biodiv. 8, 63e74.
Santiago-Rodriguez, T.M., Narganes-Storde, Y.M., Chanlatte, L., Crespo-Torres, E.,
Toranzos, G.A., Jimenez-Flores, R., Hamrick, A., Cano, R.J., 2013. Microbial commu-
nities in pre-Columbian coprolites. PLoS One 8, e65191.
Santos, C., Gibson, D., 2007. Nilonema gymnarchi Khalil, 1960 and N. senticosum (Baylis, 1922)
(Nematoda: Dracunculoidea): Gondwana relicts? Syst. Parasitol. 67, 225e234.
Savinetsky, A.B., Khrustalev, A.V., 2013. Paleoparasitological investigations in Mongolia,
middle Asia and Russia. Int. J. Paleopathol. 3, 176e181.
Schiffbauer, J.D., Xiao, S., 2011. Paleobiological Applications of Focused Ion Beam Electron
Microscopy (FIB-em): An Ultrastructural Approach to the (Micro) Fossil Record. Quan-
tifying the Evolution of Early Life. Springer.
Schmidt, G.D., Duszynski, D.W., Martin, P.S., 1992. Parasites of the extinct Shasta Ground
Sloth, Nothrotheriops shastensis, in Rampart Cave, Arizona. J. Parasitol. 78, 811e816.
Scholl, E.H., Thorne, J.L., Mccarter, J.P., Bird, D.M., 2003. Horizontally transferred genes in
plant-parasitic nematodes: a high-throughput genomic approach. Genome Biol. 4, R39.
Schweitzer, M.H., Zheng, W., Organ, C.L., Avci, R., Suo, Z., Freimark, L.M., Lebleu, V.S.,
Duncan, M.B., Vander Heiden, M.G., Neveu, J.M., Lane, W.S., Cottrell, J.S.,
Horner, J.R., Cantley, L.C., Kalluri, R., Asara, J.M., 2009. Biomolecular characterization
and protein sequences of the Campanian Hadrosaur B. canadensis. Science 324, 626e631.
Sexton, A.C., Howlett, B.J., 2006. Parallels in fungal pathogenesis on plant and animal hosts.
Eukaryot. Cell 5, 1941e1949.
Shapiro, B., Hofreiter, M., 2014. A paleogenomic perspective on evolution and gene func-
tion: new insights from ancient DNA. Science 343, 1236573.
Sharma, N., Kar, R.K., Agarwal, A., Kar, R., 2005. Fungi in dinosaurian (Isisaurus) coprolites
from the Lameta Formation (Maastrichtian) and its reflection on food habit and
environment. Micropaleontology 51, 73e82.
Silva, J.C., Egan, A., Arze, C., Spouge, J.L., Harris, D.G., 2015. A new method for estimating
species age supports the co-existence of malaria parasites and their mammalian hosts.
Mol. Biol. Evol. 32, 1354e1364.
Siveter, D.J., Briggs, D.E., Siveter, D.J., Sutton, M.D., 2015. A 425-million-year-old Silurian
pentastomid parasitic on ostracods. Curr. Biol. 25, 1632e1637.
Sj€
ostrand, J., Tofigh, A., Daubin, V., Arvestad, L., Sennblad, B., Lagergren, J., 2014. A
Bayesian method for analyzing lateral gene transfer. Syst. Biol. 63, 409e420.
Skawina, A., Dzik, J., 2011. Umbonal musculature and relationships of the Late Triassic
filibranch unionoid bivalves. Zool. J. Linn. Soc. 163, 863e883.
Sliter, W.V., 1971. Predation on benthic foraminifers. J. Foraminiferal Res. 1, 20e28.
Smith, N.A., 2015. Sixteen vetted fossil calibrations for divergence dating of Charadriiformes
(Aves, Neognathae). Palaeontologia Electronica 18.1.4FC, 1e18.
Sohl, N.F., 1964. Gastropods from the Coffee Sand (Upper Cretaceous) of Mississippi.
Sperling, E.A., Vinther, J., 2010. A placozoan affinity for Dickinsonia and the evolution of late
Proterozoic metazoan feeding modes. Evol. Dev. 12, 201e209.
Struck, T.H., Wey-Fabrizius, A.R., Golombek, A., Hering, L., Weigert, A., Bleidorn, C.,
Klebow, S., Iakovenko, N., Hausdorf, B., Petersen, M., K€ uck, P., Herlyn, H.,
Hankeln, T., 2014. Platyzoan paraphyly based on phylogenomic data supports a noncoe-
lomate ancestry of Spiralia. Mol. Biol. Evol. 31, 1833e1849.
Sudhaus, W., 2010. Preadaptive plateau in Rhabditida (Nematoda) allow the repeated
evolution of zooparasites, with an outlook on evolution of life cycles with
Spiroascarida. Palaeodiversity 3, 117e130.
Summers, M.M., Rouse, G.W., 2014. Phylogeny of Myzostomida (Annelida) and their
relationships with echinoderm hosts. BMC Evol. Biol. 14, 170.
The Importance of Fossils in the Evolution of Parasites 49

Sung, G.-H., Poinar, G., Spatafora, J.W., 2008. The oldest fossil evidence of animal
parasitism by fungi supports a Cretaceous diversification of fungalearthropod
symbioses. Mol. Phylogenet. Evol. 49, 495e502.
Sutton, M.A., Rahman, I.A., Garwood, R.J., 2014. Techniques for Virtual Palaeontology.
Wiley-Blackwell, UK.
Sutton, M.D., 2008. Tomographic techniques for the study of exceptionally preserved fossils.
Proc. R. Soc. B 275, 1587e1593.
Sz€
ollTsi, G.J., Boussau, B., Abby, S.S., Tannier, E., Daubin, V., 2012. Phylogenetic model-
ling of lateral gene transfer reconstructs the pattern and relative timing of speciations.
Proc. Natl. Acad. Sci. U.S.A. 109, 17513e17518.
Tapanila, L., October 4, 2008. Direct evidence of ancient symbiosis using trace fossils. In:
Kelley, P.H., Bambach, R.K. (Eds.), From Evolution to Geobiology: Research Ques-
tions Driving Paleontology at the Start of a New Century, Paleontological Society Short
Course, vol. 14. Paleontol Soc Papers, pp. 271e287.
Taylor, P.D., 2015. Differentiating parasitism and other interactions in fossilized colonial
organisms. Adv. Parasitol. 90, 329e347.
Thézé, J., Bézier, A., Periquet, G., Drezen, J.-M., Herniou, E.A., 2011. Paleozoic origin of
insect large dsDNA viruses. Proc. Natl. Acad. Sci. U.S.A. 108, 15931e15935.
Todd, J.A., Harper, E.M., 2011. Stereotypic boring behaviour inferred from the earliest
known octopod feeding traces: early Eocene, southern England. Lethaia 44, 214e222.
Trewick, S.A., Gibb, G.C., 2010. Vicars, tramps and assembly of the New Zealand avifauna:
a review of molecular phylogenetic evidence. Ibis 152, 226e253.
Upeniece, I., 2001. The unique fossil assemblage from the Lode Quarry (Upper devonian,
Latvia). Fossil Rec. 4, 101e119.
Upeniece, I., 2011. Palaeoecology and Juvenile Individuals of the Devonian Placoderm and
Acanthodian Fishes from Lode Site, Latvia. Disertations Geologicae Universitas Latvien-
sis 21, 1e221.
Vinther, J., 2015. A guide to the field of palaeo colour: Melanin and other pigments can
fossilise: reconstructing colour patterns from ancient organisms can give new insights
to ecology and behaviour. BioEssays 37, 643e656.
Voigt, E., 1952. Ein Haareinschluß mit Phthirapteren-Eiern im Bernstein. Mitteilungen aus
dem Geologischen Staatsinstitut in Hamburg 21, 59e74.
Voigt, E., 1959. Endosacculus moltkiae n. g. n. sp., ein vermutlicher fossiler Ascothoracide
(Entomostr.) als Cystenbildner bei der Oktokoralle Moltkia minuta. Pal€aontologische
Zeitschrift 33, 211e223.
Voigt, E., 1967. Ein vermutlicher Ascothoracide (Endosacculus (?) najdini n. sp.) als Bewohner
einer kretazischen Isis aus der UdSSR. Pal€aontologische Zeitschrift 41, 86e90.
Wade, M., 1972. Dickinsonia: polychaete worms from the late Precambrian Ediacara fauna,
South Australia. Memoirs Qld. Mus. 16, 171e190.
Wadsworth, C., Buckley, M., 2014. Proteome degradation in fossils: investigating the
longevity of protein survival in ancient bone. Rapid Commun. Mass Spectrom. 28,
605e615.
Waloszek, D., M€ uller, K.J., 1994. Pentastomid parasites from the Lower Palaeozoic of
Sweden. Trans. R. Soc. Edinburgh Earth Sci. 85, 1e37.
Waloszek, D., Repetski, J.E., Maas, A., 2005. A new Late Cambrian pentastomid and a
review of the relationships of this parasitic group. Trans. R. Soc. Edinburgh Earth Sci.
96, 163e176.
Waloszek, D., Repetski, J.E., M€ uller, K.J., 1994. An exceptionally preserved parasitic
arthropod, Heymonsicambria taylori n. sp. (Arthropoda incertae sedis: Pentastomida),
from Cambrian-Ordovician boundary beds of Newfoundland, Canada. Can. J. Earth
Sci. 31, 1664e1671.
50 Kenneth De Baets and D. Timothy J. Littlewood

Wappler, T., Smith, V.S., Dalgleish, R.C., 2004. Scratching an ancient itch: an Eocene bird
louse fossil. Proc. R. Soc. B 271, S255eS258.
Warinner, C., Speller, C., Collins, M.J., 2015. A new era in palaeomicrobiology: prospects
for ancient dental calculus as a long-term record of the human oral microbiome. Philos.
Trans. R. Soc. B 370, 20130376.
Warnock, R., 2014. Molecular clock calibration. In: Rink, W.J., Thompson, J. (Eds.), Ency-
clopedia of Scientific Dating Methods. Springer, Netherlands.
Warnock, R.C., Yang, Z., Donoghue, P.C., 2012. Exploring uncertainty in the calibration
of the molecular clock. Biol. Lett. 8, 156e159.
Warnock, R.C.M., Parham, J.F., Joyce, W.G., Lyson, T.R., Donoghue, P.C.J., 2015.
Calibration uncertainty in molecular dating analyses: there is no substitute for the prior
evaluation of time priors. Proc. R. Soc. B 282, 20141013.
Weber, M., Wey-Fabrizius, A.R., Podsiadlowski, L., Witek, A., Schill, R.O., Sugar, L.,
Herlyn, H., Hankeln, T., 2013. Phylogenetic analyses of endoparasitic Acanthocephala
based on mitochondrial genomes suggest secondary loss of sensory organs. Mol. Phylo-
genet. Evol. 66, 182e189.
Weinberg Rasmussen, H., Jakobsen, S., Collins, J., 2008. Raninidae infested by parasitic Iso-
poda (Epicaridea). Bull. Mizunami Fossil Mus. 34, 31e49.
Welch, J.R., 1976. Phosphannulus on Paleozoic crinoid stems. J. Paleontol. 50, 218e225.
Welker, F., Collins, M.J., Thomas, J.A., Wadsley, M., Brace, S., Cappellini, E., Turvey, S.T.,
Reguero, M., Gelfo, J.N., Kramarz, A., Burger, J., Thomas-Oates, J., Ashford, D.A.,
Ashton, P.D., Rowsell, K., Porter, D.M., Kessler, B., Fischer, R., Baessmann, C.,
Kaspar, S., Olsen, J.V., Kiley, P., Elliott, J.A., Kelstrup, C.D., Mullin, V.,
Hofreiter, M., Willerslev, E., Hublin, J.J., Orlando, L., Barnes, I., Macphee, R.D.,
2015. Ancient proteins resolve the evolutionary history of Darwin’s South American
ungulates. Nature 522, 81e84.
Wey-Fabrizius, A.R., Herlyn, H., Rieger, B., Rosenkranz, D., Witek, A., Welch, D.B.M.,
Ebersberger, I., Hankeln, T., 2014. Transcriptome data reveal syndermatan relationships
and suggest the evolution of endoparasitism in Acanthocephala via an epizoic stage. PLoS
One 9, e88618.
Wey-Fabrizius, A.R., Podsiadlowski, L., Herlyn, H., Hankeln, T., 2013. Platyzoan mito-
chondrial genomes. Mol. Phylogenet. Evol. 69, 365e375.
Wijayawardena, B.K., Minchella, D.J., Dewoody, J.A., 2013. Hosts, parasites, and horizontal
gene transfer. Trends Parasitol. 29, 329e338.
Wijayawardena, B.K., Minchella, D.J., Dewoody, J.A., 2015. Horizontal gene transfer in
schistosomes: a critical assessment. Mol. Biochem. Parasitol 201, 57e65.
Wilkinson, R.D., Steiper, M.E., Soligo, C., Martin, R.D., Yang, Z., Tavaré, S., 2011. Dating
primate divergences through an integrated analysis of palaeontological and molecular
data. Syst. Biol. 60, 16e31.
Williams, J.D., Boyko, C.B., 2012. The global diversity of parasitic isopods associated with
crustacean hosts (Isopoda: Bopyroidea and Cryptoniscoidea). PLoS One 7, e35350.
Wolff, E.D.S., Salisbury, S.W., Horner, J.R., Varricchio, D.J., 2009. Common avian infec-
tion plagued the tyrant dinosaurs. PLoS One 4, e7288.
Wood, H.M., Matzke, N.J., Gillespie, R.G., Griswold, C.E., 2013a. Treating fossils as ter-
minal taxa in divergence time estimation reveals ancient vicariance patterns in the palpi-
manoid spiders. Syst. Biol. 62, 264e284.
Wood, J.R., Wilmshurst, J.M., Rawlence, N.J., Bonner, K.I., Worthy, T.H., Kinsella, J.M.,
Cooper, A., 2013b. A megafauna’s microfauna: gastrointestinal parasites of New Zea-
land’s extinct moa (Aves: Dinornithiformes). PLoS One 8, e57315.
Wood, J.R., Wilmshurst, J.M., 2014. Late Quaternary terrestrial vertebrate coprolites from
New Zealand. Quat. Sci. Rev. 98, 33e44.
The Importance of Fossils in the Evolution of Parasites 51

Wood, J.R., Wilmshurst, J.M., Worthy, T.H., Holzapfel, A.S., Cooper, A., 2012. A lost link
between a flightless parrot and a parasitic plant and the potential role of coprolites in con-
servation paleobiology. Conserv. Biol. 26, 1091e1099.
W€orheide, G., Dohrmann, M., Yang, Q., 2015. Molecular paleobiology e progress and
perspectives. Palaeoworld, in press. http://dx.doi.org/10.1016/j.palwor.2015.01.005.
Xingliang, Z., Reitner, J., 2006. A fresh look at Dickinsonia: Removing it from Vendobionta.
Acta Geol. Sin. Engl. Ed. 80, 636e642.
Xylander, W.E.R., 2001. The Gyrocotylidea, Amphilinidea and the early evolution of
Cestoda. In: Littlewood, D.T.J., Bray, R.A. (Eds.), Interrelationships of the
Platyhelminthes. Taylor & Francis, London, pp. 103e111.
Yao, Y., Cai, W., Xu, X., Shih, C., Engel, M.S., Zheng, X., Zhao, Y., Ren, D., 2014.
Blood-feeding true bugs in the Early Cretaceous. Curr. Biol. 24, 1786e1792.
Zaman, L., Meyer, J.R., Devangam, S., Bryson, D.M., Lenski, R.E., Ofria, C., 2014. Coevo-
lution drives the emergence of complex traits and promotes evolvability. PLoS Biol. 12,
e1002023.
Zangerl, R., Case, G.R., 1976. Cobelodus aculeatus (Cope), an anacanthous shark from Penn-
sylvanian black shales of North America. Palaeontographica, Abteilung A 154, 107e157.
Zapalski, M.K., 2011. Is absence of proof a proof of absence? Comments on commensalism.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 302, 484e488.
Zhu, Q., Hastriter, M.W., Whiting, M.F., Dittmar, K., 2015. Fleas (Siphonaptera) are Creta-
ceous, and evolved with Theria. Mol. Phylogenet. Evol. 90, 129e139.
Zibrowius, H., 1981. Associations of Hydrocorallia Stylasterina with gall-inhabiting Cope-
poda Siphonostomatoidea from the south-west Pacific. Part I. On the stylasterine hosts,
including two new species, Stylaster papuensis and Crypthelia cryptotrema. Bijdragen tot de
Dierkunde 51, 268e286.
Zrzavý, J., 2001. The interrelationships of metazoan parasites: a review of phylum-and
higher-level hypotheses from recent morphological and molecular phylogenetic
analyses. Folia Parasitologica 48, 81e103.
CHAPTER TWO

The Geological Record of


Parasitic Nematode Evolution
George O. Poinar, Jr.
Department of Integrative Biology, Oregon State University, Corvallis, OR 97331, USA
E-mail: poinarg@science.oregonstate.edu

Contents
1. Introduction 54
2. Media for the Study of Fossil Nematodes 54
2.1 Amber 54
2.2 Rock fossils 54
2.3 Coprolites 55
3. Palaeozoic Parasitic Nematodes 55
4. Parasitic Nematode Body Fossils from the Mesozoic 55
5. Nematode Parasites from the Early Cenozoic 60
5.1 Baltic amber 60
6. Nematode Parasites from the OligoceneeMiocene 65
6.1 Dominican amber nematodes 65
6.2 Mexican amber nematodes 75
7. Nematode Parasites from the Pliocene 77
8. Nematode Parasites from the Pleistocene and Holocene 77
8.1 Nematode parasites of humans from the Pleistocene and Holocene 78
9. Stages in the Evolution of Nematode Parasites of Invertebrates 79
10. Origin of Nematode Parasites of Vertebrates 81
11. Origin of Nematode Parasites of Plants 83
12. Summary 83
Acknowledgements 86
References 86

Abstract
This chapter discusses the evolutionary history of nematode parasites of invertebrates,
vertebrates and plants based on fossil remains in amber, stone and coprolites dating
from the Palaeozoic to the Holocene. The earliest parasitic nematode is a primitive plant
parasite from the Devonian. Fossil invertebrate-parasitic nematodes first appeared in
the Early Cretaceous, while the earliest fossil vertebrate-parasitic nematodes are from
Upper Triassic coprolites. Specific examples of fossil nematode parasites over time
are presented, along with views on the origin and evolution of nematodes and their
hosts.
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.03.002 All rights reserved. 53
54 George O. Poinar, Jr.

1. INTRODUCTION
The body fossil record of parasitic nematodes is limited by their small
size and soft bodies (Littlewood and Donovan, 2003). While parasitic forms
can reach up to several metres in length (Yeates and Boag, 2006), decompo-
sition after death is quite rapid, and sclerotized structures that could be fossil-
ized are microscopic, which makes them extremely difficult to locate. While
some nematodes have been preserved in fine-grained cherts dating back to
the Devonian, fossilized resin (amber) and coprolites have been the most use-
ful media for studying the evolution and early hosts of parasitic nematodes.

2. MEDIA FOR THE STUDY OF FOSSIL NEMATODES


2.1 Amber
Amber is fossilized tree resin that preserves a wide range of organisms,
from microbes to vertebrates (Poinar, 1992). Preservation appears to be the
result of inert dehydration and fixation by natural plant products in the orig-
inal resin. The occurrence of parasitic nematodes in fossilized resin is
frequently associated with the habits of their hosts, the great majority of
which are forest living arthropods (cf. Poinar, 2011). The urge of parasitic
nematodes to escape from hosts that become entrapped in tree resin, even
if they have not completely finished their development, provides most of
our fossil records of parasitic nematodes.
Amber-containing parasitic nematodes extend from the early Cretaceous
(130e135 mya) to the Miocene (15e20 mya). Dating of amber deposits is
usually based on fossils in the surrounding bedrock. Using this method,
Dominican amber has been dated at 20e45 mya (Cepek in Schlee, 1990;
Iturralde-Vincent and MacPhee, 1996; Iturralde-Vinent, 2001), Mexican
amber at 22e26 mya (Poinar, 1992), Baltic amber, 40e50 mya (Larsson,
1978), Burmese amber, 97e110 mya (Cruickshank and Ko, 2003) and Leb-
anese amber, 130e135 mya (Poinar and Milki, 2001). Geological ages
mentioned here are based on Gradstein et al. (2012). Taxonomic groupings
of nematode categories follow treatments of De Ley and Blaxter (2004),
Eyualim-Abebe et al. (2006) and Poinar (2011).

2.2 Rock fossils


Nematode remains in sedimentary deposits mostly occur as compression fos-
sils with only the body outline and external ornamentation evident.
The Geological Record of Parasitic Nematode Evolution 55

However, in some fine-grained sedimentary deposits known as Lagerst€atten,


sclerotized parts such as onchia and spicules, and occasionally even portions
of the alimentary and reproductive systems (Poinar, 2011) can be found.
In some rare instances where silica and other dissolved minerals impreg-
nate rocks, nematodes inside can be preserved in amazing detail. This type of
permineralization has provided us with the earliest record of parasitic nem-
atodes from the Early Devonian Rhynie Chert (Poinar et al., 2008).

2.3 Coprolites
Analysis of lithified coprolites dating back millions of years (e.g. Poinar and
Boucot, 2006; Da Silva et al., 2014; Hugot et al., 2014) and desiccated dung
of more recent ages (Poinar, 2014) can reveal the presence of vertebrate-
parasitic nematodes. Ancient dung samples also are one of the best resources
for establishing early records of human nematode parasites (Gonçalves et al.,
2003). However, it can be difficult to distinguish between actual nematode
parasites and microbotrophic nematodes that entered the dung after it was
deposited (Poinar, 1983).

3. PALAEOZOIC PARASITIC NEMATODES


The only Palaeozoic body fossil of a parasitic nematode is the Early
Devonian Rhynie Chert plant-parasitic nematode, Palaeonema phyticum
Poinar et al. (2008) (Figure 1). Eggs, juveniles and adults were present in
the stomatal chambers of the Early Devonian (396 mya) land plant, Aglao-
phyton major Kidson & Lang. Their presence provides the earliest evidence
of a symbiotic association between terrestrial plants and animals and repre-
sents an early stage in the evolution of plant parasitism by nematodes and
their presence on land.

4. PARASITIC NEMATODE BODY FOSSILS FROM THE


MESOZOIC
The oldest animal-parasitic nematodes have recently been found in
Triassic cynodont coprolites in Brazil. These include a 240 mya ascarid,
Ascarites rufferi Da Silva et al. (2014) and a 240 mya old oxyurid, Paleoxyuris
cockburni Hugot et al. (2014). The next oldest member of this group is the
130 mya mermithid, Cretacimermis libani Poinar et al. (1994) from Early
Cretaceous Lebanese amber. The single specimen was coiled up in the
body cavity of an adult midge (Chironomidae: Diptera) (Figure 2).
56 George O. Poinar, Jr.

Figure 1 Palaeonema phyticum Poinar et al., 2008 surrounded by cortical cells of Aglao-
phyton major in Devonian Rhynie Chert, Aberdeen, Scotland (Scale bar ¼ 59 mm).

Figure 2 The mermithid, Cretacimermis protus Poinar and Buckley, 2006 emerging from
a biting midge in Early Cretaceous Burmese amber (Ron Buckley amber collection)
(Scale bar ¼ 1 mm).
The Geological Record of Parasitic Nematode Evolution 57

Infection was probably initiated in the larval stage of the midge and the
parasite was carried through the pupal and into the adult stage. Such hoste
parasite associations still occur today (Poinar and Poinar, 2003).
Mid-Cretaceous Burmese amber contains both plant and animal nema-
tode parasites. A small population of the fungal-feeding aphelenchoidid,
Cretaciaphelenchoides burmensis Poinar (2011), together with mycelium, pro-
vides the earliest record of mycetophagous nematodes. Adult males and
females, as well as resistant juvenile stages were present. The nematodes
could have been carried to the site by wood-boring insects, as occurs in
extant members of the family (Hunt, 1993).
A snail in Burmese amber with juveniles of Palaeocosmocerca burmanicum
Poinar (2011) adjacent to its mantle cavity provide early evidence of the
family Cosmocercidae. Features of the fossil nematodes resemble those of
the extant snail parasite, Cosmocercoides dukae (Anderson, 1960). The adults
of C. dukae live in the intestinal tract of the snail, and second stage juveniles
and third stage infectives occur in the mantle cavity. A sibling species of C.
dukae occurs in the rectum of frogs and toads (Harwood, 1930; Vanderburgh
and Anderson, 1987), but it is likely that molluscs were the original host of
Cosmocercoides spp.
Oxyurids may be the first nematodes to form parasitic associations with
terrestrial animals, beginning with invertebrates and then expanding their
host range to vertebrates. While the origin of oxyurids could extend back to
the Silurian based on the earliest fossils of their millipede hosts (Wilson and
Anderson, 2004), when they first invaded insects is unknown. The thelasto-
matid, Paleothelastoma tipulae Poinar (2011) adjacent to a cranefly in Burmese
amber (Figure 3) establishes the earliest fossil record of this group in inverte-
brates. Extant craneflies are common hosts of thelastomatids (Poinar, 1975).

Figure 3 The thelastomatid, Paleothelostoma tipulae Poinar (2011) (arrow) adjacent to


its tipulid host in Early Cretaceous Burmese amber (Scale bar ¼ 251 mm).
58 George O. Poinar, Jr.

Figure 4 Proheterorhabditis burmanicus Poinar (2011) (arrow) adjacent to its beetle host
in Early Cretaceous Burmese amber (Scale bar ¼ 152 mm).

Entomopathogenic nematodes, for example those whose infective stages


carry symbiotic bacteria in their gut and release them in a potential host, date
back to the Early Cretaceous with Proheterorhabditis burmanica Poinar (2011)
and its rove beetle (Staphylinidae) host in Burmese amber (Figure 4). Rod-
shaped bacteria similar to those associated with extant species of Heterorhab-
ditis occurred in the fossil rove beetle.
Mermithid nematodes emerging from insects also occur in Burmese
amber. Two specimens of Cretacimermis chironomae Poinar (2011) were asso-
ciated with an adult chironomid midge. These specimens, as well as Creta-
cimermis lebani from Lebanese amber, show that mermithid parasitism of
the Chironomidae occurred in Laurasia and Gondwanaland in the Early
Cretaceous.
The Burmese amber mermithid, Cretacimermis protus Poinar and
Buckley (2006) parasitized biting midges (Ceratopogonidae) of the genera
Atriculoides Remm and Leptoconops Skuse (Poinar and Monteys, 2008)
(Figure 5).
Extant species of Leptoconops feed on the blood of mammals, birds and
reptiles. While the genus Articulicoides is now extinct, it is likely that the
females also fed on vertebrate blood (Szadziewski and Poinar, 2005).
An analysis of an dinosaur coprolite from the Early Cretaceous Iguan-
odon shaft in Belgium revealed ascarid eggs. The egg of Ascarites priscus
Poinar and Boucot (2006) (Figure 6) still had its outer mammillated surface
and contained a developing juvenile nematode.
The Geological Record of Parasitic Nematode Evolution 59

Figure 5 The mermithid, Cretacimermis protus Poinar and Buckley, 2006 emerging from
a biting midge in Early Cretaceous Burmese amber (Scale bar ¼ 1 mm).

Figure 6 Egg of the early Cretaceous Ascarites priscus Poinar and Boucot, 2006 in a
coprolite of the predatory dinosaur, Megalosaurus from the Bernissant Wealden Iguan-
odon locality in Belgium (Scale bar ¼ 11 mm).
60 George O. Poinar, Jr.

Nematodes that occur in extant foraminifera tests provide a challenge to


both biologists and paleontologists. Sliter (1971) found nematodes living
within borings in the tests of the foraminifera Rosalina globularis d’Orbigny
and Bolivlina doniezi Cushman and Widkenden off Southern California
and felt that the borings were made by the nematodes rather than by
some other invertebrate. Later, other nematodes recovered from forami-
nifera tests were identified as Syringonomous typicus Hope and Murphy
(1969), Smithsoninema inaequale Hope and Tchesunov (1999) and an unde-
scribed species (Myers, 1943). Sliter (1971) used identical borings in bathyal
and neritic Cretaceous foraminifera to determine the distribution and abun-
dance of marine nematodes. Additional studies are needed to verify the abil-
ity of specialized groups of nematodes to bore through the shells of
foraminifera before fossil tests with bore holes can be used to establish the
presence of marine nematodes.

5. NEMATODE PARASITES FROM THE EARLY


CENOZOIC
5.1 Baltic amber
A wide range of parasitic nematodes occur in Baltic amber. Myce-
tophagous parasites include a reproducing population of the aphelenchoidid,
Palaeoaphelenchoides balticus Poinar (2011) that were probably feeding on
fungi on the resin-producing tree. There are many extant nematodes that
live in the tunnels of bark beetles and develop on fungi in the insect galleries.
A number of mermithids occur in Baltic amber, especially as parasites of
the family Chironomidae. The first one described in 1866, Heydenius matu-
tinus (Menge, 1866) is now considered a collective species for all mermithids
parasitizing chironomid midges in Baltic amber (Figure 7). While there may
be several genera and species involved, the absence of diagnostic characters
prohibits further identification.
It is unusual to find nematodes in coal deposits, but Heydenius antiqua
(von Heyden) was found protruding from the body of a long-horned beetle
(Coleoptera: Cerambycidae) in German brown coal strata (von Heyden,
1860), showing that mermithid lineages were parasitizing Coleoptera by
the Eocene.
Extant ants are known to host a number of mermithids and this parasitic
association dates back at least to the Eocene, as shown by Heydenius formicinus
Poinar (2002) emerging from a winged male of Prenolepis henschei Mayr in
Baltic amber (Figure 8).
The Geological Record of Parasitic Nematode Evolution 61

Figure 7 Two specimens of Heydenius matutinus (Menge, 1866): one completely


emerged and the second still in the body cavity of their chironomid midge host in Baltic
amber (Scale bar ¼ 53 mm).

Also, in Baltic amber the first fossil record of a mermithid parasite of a


hemipteran, Heydenius brownii Poinar (2001), was found protruding from
the body of a planthopper (Fulgoroidea) (Figure 9). Hemipteran hosts of
mermithids are rare today, with the only records coming from Ireland
(Helden, 2008). With a fossil record extending back to the Early Permian

Figure 8 The mermithid, Heydenius formicinus Poinar, 2002 emerging from a male
winged ant (Prenolepis henschei Mayr) in Baltic amber (Scale bar ¼ 640 mm).
62 George O. Poinar, Jr.

Figure 9 The mermithid, Heydenius brownii, Poinar, 2001, emerging from an achiliid
planthopper in Baltic amber (Scale bar ¼ 2.6 mm).

(Rasnitsyn and Quicke, 2002), hemipterans could have been ancient terres-
trial hosts for mermithids.
Several different lineages of parasitic nematodes have been found to
attack sciarid fungus gnats (Diptera: Sciaridae) in Baltic amber. One was a
mermithid, Heydenius sciarophilus Poinar (2011) (Figure 10) and another a
tetradonematid, Palaeotetradonema sciarae Poinar (2011). A third parasite line-
age was represented by Tripius balticus Poinar (2011) of the family Sphaeru-
lariidae. Several nematodes were still inside the body cavity of the adult host
while others had emerged (Figure 11). The life cycle of the fossil was prob-
ably similar to that of the extant species, Tripius sciarae (Bovien).
Other dipterous hosts of mermithids in Baltic amber were gall gnats
(Cecidomyiidae). One specimen of Heydenius cecidomyae Poinar (2011)
had completely emerged from the gall gnat while the second was only partly

Figure 10 Heydenius sciarophilus Poinar (2011) emerging from a sciarid fungus gnat in
Baltic amber (Scale bar ¼ 207 mm).
The Geological Record of Parasitic Nematode Evolution 63

Figure 11 Parasitic juveniles of Tripius balticus Poinar (2011) (arrow) that emerged from
a sciarid fungus gnat in Baltic amber (Scale bar ¼ 322 mm).

emerged (Figure 12). There are no reports of extant gall gnats parasitized by
mermithids.
Another parasite lineage of Diptera was represented by the allantonematid,
Howardula helenoschini Poinar (2003) that parasitized scuttle flies (Phoridae).

Figure 12 Two specimens of Heydenius cecidiomyae Poinar (2011), a parasite of a gall


gnat in Baltic amber (Scale bar ¼ 575 mm).
64 George O. Poinar, Jr.

Figure 13 Specimens of the allantonematid, Howardula helenoschini Poinar (2003)


(arrow) emerging from a phorid fly, Triphleba sp., in Baltic amber (Scale bar ¼ 347 mm).

Five specimens of H. helenoschini were clustered at the posterior tip of the fly’s
abdomen, obviously having just exited from the host’s body cavity (Figure 13).
The lemon-shaped parasitic female of H. helenoschini could be detected
through the host’s abdomen. Extant phorid flies are parasitized by Howardula
nematodes globally (Richardson et al., 1977; Poinar, 1975; Disney, 1994).
The nematodes normally leave the fly host via the intestine or reproductive
system (Richardson et al., 1977) and that is probably the route used by the
five fossil nematodes.
Another Baltic amber nematode in the allantonematid lineage was Palae-
oallantonema baltica Poinar (2011) that was parasitizing a rove beetle (Staph-
ylinidae). A number of last stage juveniles that had exited the host consisted
of two morphotypes: females of the first generation and pre-adults of the
second generation.
Hematophagous flies in Baltic amber also were associated with vertebrate
nematode parasitic lineages. Blackflies (Simuliidae) and biting midges (Cera-
topogonidae) are the most common biting insects in Baltic amber. A micro-
filaria of Cascofilaria baltica Poinar (2011, 2012) was adjacent to a blackfly with
a swollen abdomen, suggesting that the fly had taken a blood meal shortly
before falling in the resin. The microfilaria falls within the size range of those
of extant Onchocerca spp., a worldwide genus infecting mammals and vectored
by blackflies. This represents the oldest fossil record of a filarial nematode.
Still additional insect hosts of mermithid nematodes in Baltic amber
include a moth (Lepidoptera) parasitized by Heydenius podenasae Poinar
The Geological Record of Parasitic Nematode Evolution 65

Figure 14 Adult caddis fly (Triaenodes balticus Wichard and Barnard, 2005) parasitized
by Heydenius trichorosus Poinar (2012) in Baltic amber (Scale bar ¼ 1.4 mm).

(2012), an adult caddis fly, Triaenodes balticus Wichard and Barnard, 2005
(Trichoptera) parasitized by Heydenius trichorosus Poinar (2012) (Figure 14)
and a phasmatid, Balticophasma sp. (Phasmatodea: Phasmatidae) parasitized
by Heydenius phasmatophilus Poinar (2012) (Figure 15).
Spiders were also parasitized by mermithid nematodes in the Baltic
amber forest, and Heydenius araneus Poinar (2000) from a crab spider (Tho-
misidae) (Figure 16) is the earliest record of nematode parasitism of spiders.

6. NEMATODE PARASITES FROM THE


OLIGOCENEeMIOCENE
6.1 Dominican amber nematodes
One of the major host groups of nematode parasites in amber from the
Dominican Republic is ants and there is evidence that three lineages of para-
sitic nematodes were involved.
A mermithid lineage is represented by Heydenius myrmecophilia Poinar
et al. (2006) that had recently emerged from a species of the ant genus Line-
pithema Mayr.
A second lineage was a member of the family Tetradonematidae, Myr-
meconema antiqua Poinar (2011). This species was represented by a mass of
66 George O. Poinar, Jr.

Figure 15 A phasmatid, Balticophasma sp. (Phasmatodea: Phasmatidae) parasitized by


Heydenius phasmatophilus Poinar (2012) in Baltic amber (Scale bar ¼ 2.1 mm).

eggs (Figure 17) that had escaped through a hole in the gaster (abdomen) of
a species of Cephalotes Latreille, 1803. The fossil probably had a life cycle
similar to the extant Myrmeconema neotropicum Poinar and Yanoviak
(2008), a parasite of Cephalotes atratus (L.). The life cycle of this group is
especially fascinating and unique. When females of M. neotropicum have
matured and the body cavity of the ant is filled with eggs, the color of

Figure 16 The spider mermithid, Heydenius araneus Poinar (2000) adjacent to its clu-
bionid spider host in Baltic amber (Scale bar ¼ 2 mm).
The Geological Record of Parasitic Nematode Evolution 67

Figure 17 Eggs of Myrmeconema antiqua Poinar (2011) from an ant of the genus Ceph-
alotes Latreille in Dominican amber (Scale bar ¼ 360 mm).

the host’s gaster changes from black to a shiny red. The bright red abdomen
resembles ripe berries that are relished by birds. The infected ants stand on
stems and raise their abdomens high in the air. When a bird ingests an
infected ant, the eggs of M. neotropicum pass thought the bird’s digestive tract
and are deposited in the droppings. Worker ants collect and feed the
infested bird droppings to their brood. The ingested nematode eggs hatch
in the larval guts and the juveniles penetrate into the insect’s body cavity. It
appears that this complicated life cycle was established already some 20e30
million years ago.
Cephalotes ants are also hosts to a third parasite lineage, the Allantonema-
tidae. Mature juveniles of Palaeoallantonema cephalotae Poinar (2011) were
emerging from a worker of Cephalotes serratus in a piece of amber (Figure 18).
A similar hosteparasite association occurs with an undescribed allantonema-
tid and workers of Cephalotes christopherseni in Peru.
A separate lineage of allantonematids, represented by Palaeoparasitylenchus
dominicana Poinar (2011) parasitized drosophilid fruit flies in the Dominican
amber forest. A female fly was heavily infected and large numbers of juve-
niles and second-generation female nematodes were adjacent to the host
(Figure 19). The life cycle of P. dominicanus was probably similar to that
68 George O. Poinar, Jr.

Figure 18 Detail of a specimen of Palaeoallantonema cephalotae Poinar (2011)


from the ant, Cephalotes serratus Vierbergen & Scheve, in Dominican amber (Scale
bar ¼ 24 mm).

of the extant species, Parasitylenchus nearcticus Poinar et al. (1997) from


Eastern North America.
Moth flies (Diptera: Psychodidae) were also attacked by mermithid
nematodes in the Dominican amber forest. Two juvenile specimens of
Heydenius psychodae Poinar (2011) had completely emerged from an adult
moth fly (Figure 20). There appears to be no record of mermithids parasit-
izing extant moth flies. Another mermithid lineage attacked scavenger flies

Figure 19 Parasitic juveniles of Palaeoparasitylenchus dominicanus Poinar (2011)


emerging from a drosophilid fly in Dominican amber (Scale bar ¼ 432 mm).
The Geological Record of Parasitic Nematode Evolution 69

Figure 20 Two specimens of the mermithid, Heydenius psychodae Poinar (2011), with
their moth fly host in Dominican amber (Scale bar ¼ 800 mm).

(Diptera: Scatopsidae). Five specimens of Heydenius scatophilus Poinar


(2011) were adjacent to their scatopsid host in Dominican amber. There
appears to be no report of mermithids infecting extant scavenger flies.
Another fly group in the Dominican amber that was parasitized by mer-
mithid nematodes was milichids (Diptera: Milichidae). The mermithid, Hey-
denius dipterophilus Poinar (2011) had completely emerged from its milichid
host in Dominican amber. There are no extant records of mermithids from
this family of flies.
Fungus gnats of the family Mycetophilidae are common in Dominican
amber, and one representative was parasitized by a member of the Iotonchi-
dae lineage. Iotonchid nematodes have evolved a complex life cycle alter-
nating between insects and fungi. The Dominican amber Paleoiotonchium
dominicanum Poinar (2011) probably had a similar dual host life cycle. The
swollen female nematode and a number of juveniles remain in the body
cavity of the fossil fly, while additional juveniles are in the amber surround-
ing the host (Figure 21). The life cycle of P. dominicanum may have been
similar to that of the extant Iotonchium californicum Poinar (1991). The latter
has a mycetophagous cycle that occurs in mushrooms also invaded by fungus
gnats, which become a secondary host. The fossil shows that life cycles
70 George O. Poinar, Jr.

Figure 21 Parasitic juveniles of Paleoiotonchium dominicanum Poinar (2011) inside the


body cavity of a mycetophilid fungus gnat in Dominican amber (Scale bar ¼ 75 mm).

alternating between insects and fungi are ancient and provides an evolu-
tionary scenario that insect parasitism by tylenchoid nematodes may have
evolved from mycetophagous lineages.
Strictly mycetophagous nematodes of aphelenchoidid lineages have also
been described from Dominican amber. The stylet-bearing Bursaphelenchus
similus Poinar (2011) probably originated from a gallery belonging to an
adjacent platypodid beetle. Some 60 specimens of the aphelenchoidid Crypt-
aphelenchus dominicus Poinar (2011) were associated with an adult platypodid
beetle. Many of them were dauer (resistant) juveniles, which is a phoretic
stage. A third lineage was represented by numerous individuals of Oligaphe-
lenchoides dominicanus Poinar (2011) feeding on a large mat of adjacent fungal
hyphae.
Also in Dominican amber are nematode parasites of herbivorous insects.
The allantonematid, Palaeoallantonema apionae Poinar (2011) was parasitizing
an apionid weevil, and numerous individuals of the diplogastrid, Synconema
dominicana Poinar (2011) were adjacent to a fig wasp (Hymenoptera: Agao-
nidae) (Figure 22), thus providing indirect evidence of fig trees in the ancient
forest. Based on their attenuated head, lack of a stylet, rounded lip region
and elongate tail, S. dominicana was placed in the Parasitodiplogaster clade,
an extant genus that parasitizes figs wasps in Africa and Mesoamerica (Poinar,
1979; Poinar and Herre, 1991).
Fossil-parasitic nematodes of higher plants are extremely rare; however,
numerous specimens of Oligaphelenchoides maximus Poinar (2011) in all
developmental stages were found in a section of a monocot rootlet in
Dominican amber (Figure 23). The long, slender, stylet-bearing adults
The Geological Record of Parasitic Nematode Evolution 71

Figure 22 Parasitic stages of Synconema dominicana Poinar (2011) emerging from a fig


wasp in Dominican amber (Scale bar ¼ 545 mm).

and males with a C-shaped terminus align the species with members of the
Aphelenchoididae.
Another species of stylet-bearing plant-parasitic nematode in Dominican
amber was the anguinid, Palaeoanguina dominicana Poinar (2011) (Figure 24).
It apparently had been feeding within and then attempted to escape from a
seed that fell in the resin. Extant anguinids can withstand desiccation and
survive for years in dried seeds.
One interesting lineage of aphelenchoidids in Dominican amber is the
Acugutturidae Hunt (1993). Extant representatives are ectoparasites on
the external surface of insects, especially moths. Several specimens of Seto-
nema protera Poinar (2011) belong to this family. One large specimen is adja-
cent to its moth host (Figure 25), while another is some distance away.

Figure 23 Population of the plant nematode, Oligaphelenchoides maximus Poinar


(2011) in Dominican amber (Scale bar ¼ 857 mm).
72 George O. Poinar, Jr.

Figure 24 Juvenile female of Palaeoanguina dominicana Poinar (2011) in Dominican


amber (Scale bar ¼ 128 mm).

Several additional specimens are attached to the moth’s abdomen. Extant


species of this family occur on six families of moths, especially noctuids,
and apparently transfer from moth to moth during mating.
While the high rate of mermithid parasitism of chironomid midges in
Eocene Baltic amber was previously noted, mermithid parasitism of this
host group also occurs in Dominican amber as exemplified by Heydenius

Figure 25 The ectoparasitic nematode Setonema protera Poinar (2011) (arrow) adja-
cent to its moth host in Dominican amber (Scale bar ¼ 245 mm).
The Geological Record of Parasitic Nematode Evolution 73

Figure 26 A specimen of Heydenius dominicus Poinar, 1984 inside the body cavity of a
Culex mosquito in Dominican amber (Scale bar ¼ 348 mm).

neotropicus Poinar (2011). Other mermithid lineages, such as Heydenius dom-


inicus Poinar (1984) (Figure 26) parasitized mosquitoes.
A few fossil blood-sucking flies were found vectoring filarial parasites in
Dominican amber. One female mosquito was adjacent to an infective and
pre-infective stage of the filarial nematode, Cascofilaria dominicana Poinar
(2011). The nematodes resemble the infective stages of the extant frog para-
site, Foleyella duboisi (Witenberg and Gerichter, 1944). Members of Foleyella
infect tree frogs (Leptodactylus spp.) in South America and Eleutherodactylus
tree frogs have been found in Dominican amber (Poinar and Poinar, 1999).
A second filarial nematode adjacent to another adult mosquito was
described as Cascofilaria parvus Poinar (2011). It is thought to represent an
infective stage juvenile but could not be further identified.
Nematodes also parasitized sand flies, another group of bloodsuckers in
the Dominican amber forest. A female sand fly of the genus Lutzomyia was
parasitized by Palaeoallantonema phlebotomae Poinar (2011). Aside from those
adjacent to the sand fly, nematode parasites also occurred inside the host.
Saprophagous flies were also attacked by mermithids in Dominican
amber. Heydenius saprophilus Poinar (2011) had been parasitizing a wood
gnat (Anisopodidae) (Figure 27).
The host range of mermithid nematodes in Dominican amber is
amazingly broad and also includes predatory beetles. Heydenius lamprophilus
Poinar (2011) was parasitizing an adult firefly (Coleoptera: Lampyridae) in
Dominican amber (Figure 28), and a rove beetle (Staphylinidae) was sur-
rounded by 44 juveniles of the allantonematid parasite, Palaeoallantonema
dominicana Poinar (2011) (Figure 29).
74 George O. Poinar, Jr.

Figure 27 Parasitic juvenile of Heydenius saprophilus Poinar (2011) adjacent to its wood
gnat host (Diptera: Anisopodidae) in Dominican amber (Scale bar ¼ 971 mm).

One interesting record of parasitism unknown today involved several


nematodes that had emerged from a parasitic wasp (Hymenoptera: Pteroma-
lidae). Two individuals were projecting from the body of the wasp and one
was still inside the wasp. The nematodes, described as Chalcidonema paradoxa
Poinar (2011) were tentatively placed in the Aphelenchoididae (Figure 30).

Figure 28 Heydenius lamprophilus Poinar (2011) from a lampyrid beetle in Dominican


amber (Scale bar ¼ 914 mm).
The Geological Record of Parasitic Nematode Evolution 75

Figure 29 A group of parasitic juveniles of Palaeoallantonema dominicana Poinar


(2011) adjacent to their rove beetle host in Dominican amber (Scale bar ¼ 233 mm).

Since the wasp larvae develop as internal insect parasites, it is possible that the
nematodes were also parasitizing the same host and entered the wasp larva.

6.2 Mexican amber nematodes


The Mexican amber forest was in large part similar to that of the Dominican
amber forest with similar insect genera but different species, showing that the
forests were separated long enough for speciation to occur.
One of the most spectacular nematode parasites in Mexican amber was a
population of the aphelenchoidid, Oligaphelenchoides atrebora Poinar (1977).
Males, females, eggs and juveniles were present, along with fungal hyphae

Figure 30 Stages of Chalcidonema paradoxa Poinar (2011) (arrows) associated with a


chalcidoid wasp in Dominican amber (Scale bar ¼ 471 mm).
76 George O. Poinar, Jr.

Figure 31 A reproducing population of Oligaphelenchoides atrebora Poinar, 1977 in


Mexican amber (Scale bar ¼ 176 mm).

that served as their food source (Figure 31). Also present were nematopha-
gous fungi that had parasitized several individuals of O. atrebora (Jansson and
Poinar, 1986) (Figure 32).
Few parasitic nematodes have been recovered from Mexican amber;
however, one interesting species was the tetratonematid, Palaeotetradonema

Figure 32 Thick hyphae of a nematophagous fungus inside a specimen of Oligaphelen-


choides atrebora Poinar, 1977 in Mexican amber (Scale bar ¼ 9 mm).
The Geological Record of Parasitic Nematode Evolution 77

phlebotomae Poinar (2011) that attacked a phlebotomine sand fly. Placement


in the Tetradonematidae was based on the structure of developing ova in
one specimen that was similar in size to ova in the extant sand fly tetrado-
nematid, Didilea ooglypta Tang et al. (1993).

7. NEMATODE PARASITES FROM THE PLIOCENE


An interesting Pliocene nematode parasite is the mermithid, Heydenius
tabanae Poinar (2011) emerging from an adult horsefly (Diptera: Tabanidae)
in sedimentary deposits in Germany dated at about 2.5 million years
(Grabenhorst, 1985). The horsefly was identified as the extant Tabanus sude-
ticus Zeller. It is interesting that H. tabanae was emerging from an adult
horsefly, while all extant cases of horsefly parasitism by mermithids are in
larvae (Poinar, 1985; Poinar and Lane, 1978).

8. NEMATODE PARASITES FROM THE PLEISTOCENE


AND HOLOCENE
Several vertebrate-parasitic nematodes have been reported from
deposits in the Pleistocene that extends from 1.81 mya to 11,500 BP and
the Holocene from 11,500 BP to 3300 BC. The oldest records of various
vertebrate parasite lineages are presented below. In most cases, the hosts
are still extant.
Palisade worms, Strongylus edentatus (Looss, 1900), were recovered from
the intestine of a mummified Late Pleistocene horse (Dubinina, 1972).
These nematodes are widespread and reproduce in the caecum and colon
of horses today (Anderson, 2000). Horse pinworm eggs (Oxyuris equi
Schrank) can be used to place horses at ancient locations (Jansen and
Over, 1962, 1966).
Pinworms of the genus Syphacia Seurat were recovered from fossilized
remains of a 10,000e12,000-year-old ground squirrel (Citellus sp.) (Dubinin,
1948). The fossil worms closely resemble the extant species Syphacia obvetata
(Rudolphi), which occurs in the intestine of rodents worldwide.
Whipworm (Trichuris Roederer) remains date from 30,000-year-old
eggs in coprolites of the caviid rodent, Kerodon rupestris, in Brazil. The
authors suggested that climate change was responsible for the disappearance
of Trichuris sp. in extant populations of K. rupestris at the same location
(Ferreira et al., 1991).
78 George O. Poinar, Jr.

Nine-thousand-year-old coprolites from a Brazilian iguaniid, Tropidarus


sp., contained eggs of an oxyurid considered to be Parapharyngodon sceleratus
(Travassos, 1923) that infects the extant lizard, Tropidarus torquatus (Ara
ujo
et al., 1982).
Some rare fossil nematodes have been obtained from remains of extinct
animals such as coprolites from a 1.5-million-year-old extinct Hyaena Bris-
son, 1762. The nematode juveniles could not be identified, but they could
have belonged to either Ancyclostoma or Toxascaris since both genera para-
sitize extant hyaenids (Ferreira et al., 1993).
An analysis of dried 10,000-year-old boluses of the extinct ground sloth,
Mylodon listai Ameghino, 1889, in Argentina produced elliptical bodies that
are thought to be nematode eggs (Ringuelet, 1957). Shasta Ground Sloth
(Nothrotheriops shastensis) boluses dated at 10,000  180 years from Rampart
Cave, Arizona contained first stage juveniles of the oxyurid, Agamofilaria oxy-
ura Schmidt et al. (1992) and a strongyloid, Strongyloides shastensis Schmidt
et al. (1992).
Rarely it is possible to obtain molecular evidence of parasites from sub-
fossilized remains. Wood et al. (2013) reported extracting ancient nematode
DNA from four moa species in New Zealand. Using these markers and
microscopic analysis, nematode representatives of the Heterakoidea, Tri-
chostrongylidae and Trichinellidae were identified in the moa coprolites.

8.1 Nematode parasites of humans from the Pleistocene and


Holocene
The search for human nematode parasites by anthropologists and archaeol-
ogists falls under the discipline of paleoparasitology or archaeoparasitology
(Taylor, 1955; Klicks, 1990; Reinhard, 1992; Reinhard and Ara ujo,
2008). Eggs and larvae of most gastrointestinal parasites have been obtained
from coprolites, cesspits, latrines and mummies (Gonçalves et al., 2003;
Sandison and Tapp, 1998). While the earliest evidence of human nematode
parasites are 30,000e24,000-year-old ascarid eggs from the Pleistocene
(Bouchet et al., 1996), these records are not old enough to shed light on
the origin of nematode lineages in humans. However, they do provide
important information on the types and geographical distribution of human
nematode parasites.
We know very little about how humans acquired their 138 species of
nematode parasites (Crompton, 1999). Egyptian mummies have been an
excellent source of human nematodes, especially those ‘natural mummies’
that were preserved with all their organs intact. Pages from the Papyrus Ebers
The Geological Record of Parasitic Nematode Evolution 79

dating from 3553 to 3550 BC, include treatments for ascarids (Ascaris lumbri-
coides L.), hookworms (Ancyclostoma duodenale Dubini) and Guinea worms
(Dracunculus medinensis L.) (Bryan, 1931; Ebbell, 1937).
Additional nematodes found in Egyptian mummies, but not mentioned
in the Papyrus Ebers, were whipworms (Trichuris trichiura L.), filarial nema-
todes (probably Wuchereria Silvo Araujo or Brugia spp.), and Trichinella spiralis
Owen and Stronglyoides stercoralis Bavay. Mummies show that Egyptians
were parasitized by ascarids some 2200 years ago (Cockburn et al., 1975),
infections of D. medinensis 3991e3786 years ago (Tapp and Wildsmith,
1993) and scrotal filarial nematodes (Tapp and Wildsmith, 1993).
It is curious that pinworms (Enterobius vermicularis L.) have only been
rarely recovered from ancient Egyptian mummies (Horne, 2002) and are
not mentioned in the Papyrus Ebers. Since the earliest record of pinworms
is from the New World, perhaps this species originated in the Americas,
although this might be a sampling bias.
The oldest records of human hookworms are 7230-year-old eggs recov-
ered from coprolites in Brazil (Montenegro et al., 2006; Ferreira et al.,
1987), showing that the parasites were present in South America well before
the Spanish invasion.
The whipworm, T. trichiura, infects slightly over 1000 million humans
worldwide today (Crompton, 1999). The oldest record of human parasitism
by this species dates back between 7000 and 8000 BP based on eggs from
human remains in a pre-Colombian bog in Chile, suggesting that this species
is endemic to the New World although Klicks (1990) concluded that both
T. trichiura and E. vermicularis were introduced to the New World by people
migrating across the Bering strait from Siberia.
When parasitic nematodes were originally acquired by humans is
unknown. Recent studies on nematode parasites of nonhuman primates
suggest that at least all human intestinal nematodes could have been acquired
from other primates (Dupain et al., 2009; Kaur and Singh, 2009). In fact, it is
likely that most, if not all, nematode parasites that plague humans today were
obtained through their distant ancestors and domesticated animals.

9. STAGES IN THE EVOLUTION OF NEMATODE


PARASITES OF INVERTEBRATES
Nematode parasitism of animals probably first originated in invertebrate
marine hosts with the ‘aphasmidians’ and later on land with the ‘phasmidians’.
It is likely that nematodes arose in the sea as free-living microbotrophs in the
80 George O. Poinar, Jr.

Precambrian, even earlier than has been proposed using relaxed molecular
clock methodologies (Rota-Stabelli et al., 2013), and were parasitizing marine
invertebrates by the Cambrian and terrestrial invertebrates by the Ordovician
based on extrapolations from extant parasiteehost relationships and the fossil
record of the hosts of nematode parasites (Poinar, 2011).
Early hosts of nematode parasites of invertebrates were probably repre-
sentatives of Tetradonematidae, Marimermithidae, Echinomermellidae,
Benthimermithidae, Monhysteridae and Leptolaimidae (see Poinar (2011)
for the systematic placement of these groups) that today parasitize marine
ostracods, copepods, shrimp, amphipods and isopods (Petter, 1980; Poinar
et al., 2002, 2009), starfish (Asteroidea) (Rubtsov and Platonova, 1974;
Rubtsov, 1977), brittle stars (Ophiuroidea) (Ward, 1933), sea urchins
(Echinoidea) (Gemmill and Von Linstow, 1902; Jones and Hagen, 1987),
priapulids (Rubtsov, 1980), polychaetes (Petter, 1983), tubicifid oligo-
chaetes (Hallett et al., 2001), foraminifera (Hope and Tchesunov, 1999)
and marine nematodes (Chesunov, 1988; Tchesunov and Spiridonov,
1993). Some of these host groups, such as polychaetes, ostracods and fora-
minifera, have fossil records extending back to the Cambrian (Lehmann and
Hillmer, 1983).
Today, members of the Mermithidae parasitize invertebrates in the
marine and terrestrial habitat. When mermithids shifted towards land,
marine amphipods and intertidal chironomid midges could have served as
hosts (Poinar, 1975; Schlinger, 1975; Poinar et al., 2002). Chironomid
midges have a fossil record extending back to the Late Triassic (Rasnitsyn
and Quick, 2002) and Cretacimermis and Heydenius show that mermithid
parasitism of these flies was present in the Early Cretaceous.
Later, mermithids selected freshwater aquatic or semiaquatic Diptera as
hosts, such as biting midges (Ceratopogonidae), scavenger flies (Scatopsi-
dae), gall gnats (Cecidiomyiidae), moth flies (Psychodidae), wood gnats
(Anisopodidae), scuttle flies (Phoridae), fruit flies (Drosophilidae) and
fungus gnats (Sciaridae and Mycetophilidae). These insect families all
have fossil records dating back to the Jurassic (Rasnitsyn and Quicke,
2002).
Three terrestrial groups of invertebrate-parasitic nematodes, the oxy-
urids, drilonematids and cosmocercoids, are particularly primitive and
were the possible first terrestrial parasites of invertebrates (Osche, 1963;
Inglis, 1965). Oxyurids probably parasitized invertebrates before verte-
brates and while P. tipulae) establishes the oxyurids as parasites of Diptera
in the Early Cretaceous, pinworms could have potentially already occurred
The Geological Record of Parasitic Nematode Evolution 81

in earliest known (Silurian) millipedes. Other early hosts for oxyurids could
have been isopods (Schwenk, 1927) and annelids (Poinar, 1978b; Yeates
et al., 1998). The drilonematids represent a primitive, but highly special-
ized lineage now confined to the coelom of earthworms. The discovery
of Mesidionema praecomasculatis Poinar (1978a), an earthworm parasite that
has morphological and biological features of both oxyurids and drilonema-
tids, suggests that drilonematids could have evolved as a specialized clade
from oxyurids.
The Cosmocercoids may have evolved in gastropods Anderson, 2000;
McClelland, 2005), and P. burmanicum shows that a snail-infecting lineage
was already in existence by the Early Cretaceous (Poinar, 2011). Infected
snails eaten by amphibians could have initiated the vertebrate-parasitic
taxa, although it has been suggested that amphibians may have been the
original hosts (Vanderburgh and Anderson, 1987).
The first fossil record of parasitic rhabditids is the Early Cretaceous Pro-
heterorhabdites burmanicus parasitizing a rove beetle. Heterorhabditids are
thought to have evolved from free-living, intertidal rhabditids (Poinar,
1993) and their early hosts could have been beach dwelling crustaceans (iso-
pods, amphipods, etc.) and intertidial insects.
Insect-parasitic tylenchs and aphelenchs probably evolved from myce-
tophagous lineages like Cryptaphylenchus dominicus that were associated
with platypodid beetles. This could lead to dual fungal and insect parasitism
as seen in the Dominican amber Paleoiotonchium dominicanum that probably
had a life cycle alternating between mushrooms and fungus gnats. While
there is no evidence of fungal fruiting bodies in the amber piece with Pale-
oiotonchium, mushrooms and other fungi occur in Dominican amber (Poinar
and Poinar, 1999; Boucot and Poinar, 2010). The fossil representatives of
Howardula, Palaeoallantonema, Palaeoparasitylenchus and Tripius show that the
typical allantonematidesphaerularid life cycle involving only insect hosts
existed by the beginning of the Cenozoic.

10. ORIGIN OF NEMATODE PARASITES OF


VERTEBRATES
The earliest fossil vertebrates were in shallow, near shore Cambrian
and Ordovician deposits. There is no evidence of their presence in fresh-
water deposits prior to the Early Devonian (Carroll, 1988). In the Silurian
seas, nematodes probably parasitized the primitive-jawed acanthodians
82 George O. Poinar, Jr.

as well as elasmobranchs and the ancestors of coelacanths based on present-


day host records (Poinar, 2011).
It is likely that oxyurids were the earliest nematode parasites of terrestrial
vertebrates, and their antiquity has recently been demonstrated with the dis-
covery of P. cockburni Hugot et al. (2014) in a 240-million-year-old cyno-
dont coprolite (Hugot et al., 2014). There is quite a time lapse to the
next oldest record of vertebrate pinworms in Holocene and Pleistocene de-
posits (Jansen and Over, 1962, 1966; Poinar, 2014).
However, the ascarid lineage may be just as ancient since the Triassic fos-
sil, A. rufferi Da Silva et al. (2014) dates from the same time period and same
host as P. cockburni (Da Silva et al., 2014). The next oldest fossil ascarid par-
asites are the Early Cretaceous A. priscus and A. gerus recovered from an
dinosaur coprolite (Poinar and Boucot, 2006).
The three records of filarial nematodes in amber establish this nematode
lineage in the Cenozoic and show what types of insect vectors were trans-
mitting these parasites. The next oldest records of vertebrate parasitism are
30,000-year-old eggs of A. lumbricoides from human coprolites (Patrucco
et al., 1983) and 30,000-year-old whipworm eggs from coprolites of a caviid
rodent (Ferreira et al., 1991).
The question regarding the original host of heteroxenous parasites (those
using an invertebrate as an intermediate host and a vertebrate as a final,
developmental host) has been debated for years. These nematodes, exempli-
fied by the Spirurida, actually have two infective stages, one for the inver-
tebrate and the other for the vertebrate. Did these heteroxenous
nematodes evolve first in vertebrates with invertebrates becoming second-
arily infected by ingesting eggs in faecal material? Or were invertebrates
parasitized first and the vertebrate cycle was established when the latter hosts
were eaten? Evidence for the invertebrate first theory is that (1) some of
these parasites can complete their entire development in the invertebrate
host and (2) invertebrate hosts appeared in the fossil record before those
of vertebrates.
Unfortunately, the fossil record of nematodes is too sparse to answer this
question; however, estimated dates of the origin of vertebrate- and inverte-
brate-parasitic nematode families can be inferred from the earliest fossil re-
cord of their host group or molecular clock estimates. Regarding the
family Spiruridae Railliet and Henry, 1915, the earliest vertebrate host
group appears in the Cretaceous while the earliest fossil record of an inver-
tebrate host group is in the Carboniferous (Poinar, 2011).
The Geological Record of Parasitic Nematode Evolution 83

11. ORIGIN OF NEMATODE PARASITES OF PLANTS


Plant parasitism evolved independently in at least four terrestrial line-
ages, the Enoplida, Triplonchia, Dorylaimia and Rhabditida (Poinar, 2011).
Plant-parasitic nematodes may have existed in the Silurian or even Ordovi-
cian since the earliest known fossils are from the Devonian (Poinar et al.,
2008). There is no evidence yet of any marine lineage of plant-parasitic
nematodes. Those plant-parasitic nematodes found in the sea are terrestrial
lineages that invaded littoral and sublittoral habitats. These include the
tylench genera Halenchus Cobb (Fortuner and Maggenti, 1987), Hirschma-
niella Luc & Goodey (Luc, 1987) and the aphelench species Aphelenchoides
marinus Timm & Franklin and A. gynotylurus Timm & Franklin (Timm
and Franklin, 1969).
The most primitive higher plant parasites are members of Aphelenchi-
dae, Tylenchidae, Paraphelenchidae, Neotylenchidae and Paurodontidae,
whose lineages could well extend back to the Silurian with early fossil
records of fungi (Taylor et al., 2015). Siddiqi (1983) supposed that tylenchid
plant parasites originated from algal-feeding lineages while Paramonov
(1962) and Maggenti (1971) felt they developed from mycetophagous
lineages.
The oldest aphelench fossil is the Early Cretaceous C. burmensis followed
by the Eocene P. balticus (Poinar, 2011). The above species represent free-
living mycetophages that are probably the primitive aphelench trophic stage.
Just when aphelenchs shifted their diet to higher plants is not known, but the
Dominican amber Oligaphelenchoides maxima is considered to represent an
aerial parasite with a biology similar to some extant Aphelenchoides spp.
The fossil Palaeoanguina dominicana that was developing in a seed in Domin-
ican amber is further evidence of obligate parasitism of aerial portions of
plants. This species probably had a life cycle similar to some extant members
of the genus Anguina Scopoli (Fortuner and Maggenti, 1987).

12. SUMMARY
Since nematodes most likely evolved in the sea (Poinar, 2011), it is
highly likely that the earliest animal parasites already lived in trilobites,
eurypterids and other marine invertebrates, although so far no direct fossil
evidence has become known. The first vertebrate parasites might have used
84 George O. Poinar, Jr.

elasmobranchs and primitive marine fish as hosts based on extant parasitee


host associations (Poinar, 2011). Since their fossil record is so sparse, deter-
mining the evolutionary clocks of the many lineages of parasitic
nematodes, each of which evolved independently, is difficult. However,
the minimum longevity of some nematode orders and infraorders can be
determined by body fossils (Figure 33), and time constraints on other lin-
eages can be inferred with molecular clock methodologies. However,
morphology and molecular methods do not always match. For instance,
some molecular clock studies place the origin of ascarids in the Jurassic
(Blaxter, 2009), but this is clearly incorrect since ascarid body fossils occur
in the Triassic (Da Silva et al., 2014). Since most nematode families have
multiple hosts, the earliest fossil of the most primitive host can be used
to estimate the possible date of origin for various nematode groups (see
Poinar (2011)). Using fossil host records, the earliest animal-parasitic nem-
atodes would have been members of the marine families Benthimermithi-
dae, Tetradonematidae and Leptolaimidae, whose invertebrate host groups
extend back to the Cambrian. The earliest terrestrial animal parasites would
be Thelastomatoidea, Rhigonematoidea and Ransomnematoidea that
parasitized millipedes, which have a body fossil record dating back to the
Silurian (Wilson and Anderson, 2004). The first vertebrate parasites would
have used elasmobranchs and primitive marine fish as hosts (Poinar et al.,
2014). The earliest higher plant parasites, which could have evolved
from mycetophagous lineages as far back as the Silurian could have been
lineages of Aphelenchidae, Paraphelenchidae, Neotylenchidae, Paurodon-
tidae and Tylenchidae (Poinar, 2011). With so many different lineages of
nematode parasites, it is only natural to assume that parasitism has evolved
convergently and independently in multiple lineages of nematodes. Ac-
cording to recent molecular studies (e.g. Blaxter and Koutsovoulos,
2015), plant parasitism has evolved at least three times, animal parasitism
has arisen at least ten times across a wide range of invertebrates and five
times in vertebrates hosts within the three nematodes subclasses. However,
many additional origins of parasitism could be proposed based on firm
morphological data. Sudhaus (2008, 2010) used morphological characters
to show that there are at least 20 independent origins of insect parasitism
by nematodes.
We look forward to the discovery of additional nematode fossils associ-
ated with identifiable host remains (including coprolites) that will help us to
better understand the origin and evolution of nematode parasitism of verte-
brates, invertebrates and plants.
The Geological Record of Parasitic Nematode Evolution
Figure 33 Minimum longevity of nematode orders and infraorders based on body fossils. Numbers refer to mya (million years ago).

85
86 George O. Poinar, Jr.

Molecular clock methodologies are extremely important, and together


with morphology, the two disciplines complement each other very well.
Together they can show relationships between extant lineages, perform
ancestral host reconstruction with phylogenetic dating analyses and deter-
mine when various lineages might have appeared in geological time (Sung
et al., 2008; Poinar et al., 2011).

ACKNOWLEDGEMENTS
The author would like to acknowledge E. J. Brill and ‘Nematology’ for previously having
published many of the figures used in the present work.

REFERENCES
Anderson, R.C., 1960. On the development and transmission of Cosmocercoides dukae of
terrestrial mollusks in Ontario. Canadian J. Zool. 38, 801e825.
Anderson, R.C., 2000. Nematode Parasites of Vertebrates, second ed. CAB International,
Wallingford.
Ara
ujo, A.J.G., Confalonieri, U.E.C., Ferreira, L.F., 1982. Oxyurid (Nematoda) egg from
coprolites from Brazil. J. Parasitol. 68, 511e512.
Bouchet, F., Baffier, D., Girard, M., Morel, P., Paicheler, J.C., David, F., 1996. Paléopara-
sitologie en contexte Pléistocene premieres observations a la Grand Grotte d’Arcy-sur-
Cure (Yonne), France. C. R. Acad. Sci. Paris 319, 147e151.
Boucot, A., Poinar Jr., G., 2010. Fossil Behavior Compendium. CRC Press, Boca Raton.
Blaxter, M., 2009. Nematodes (Nematoda). In: Hedges, B.S., Kumar, S. (Eds.), The Time-
tree of Life. University Press, Oxford, pp. 247e250.
Blaxter, M., Koutsovoulos, G., 2015. The evolution of parasitism in Nematoda. Parasitology
142 (Suppl. 1), 26e39.
Bryan, C.P., 1931. The Papyrus Ebers. Appleton and Company, New York.
Carroll, R.L., 1988. Vertebrate Paleontology and Evolution. W. H. Freeman and Co., New
York.
Chesunov, A.V., 1988. A case of nematode parasitism in nematodes. A new find and rede-
scription of a rare species Benthimermis australis Petter, 1983 (Nematoda: Marimermithida:
Benthimermithidae) in South Atlantic. Helminthologia 25, 115e128.
Cockburn, A., Barraco, R.A., Reyman, T.A., Peck, W.H., 1975. Autopsy of an Egyptian
mummy. Science 187, 1155e1160.
Crompton, D.W.T., 1999. How much human helminthiasis is there in the world? J. Para-
sitol. 85, 397e403.
Cruickshank, R.D., Ko, K., 2003. Geology of an amber locality in the Hukawng Valley,
northern Myanmar. J. Asian Earth Sci. 21, 441e455.
Da Silva, P., Borba, V.H., Dutra, J.M.F., Leles, D., da-Rosa, A.A.S., Ferreira, L.F.,
Araujo, A., 2014. A new ascarid species in cynodont coprolite dated of 240 million
years. Ann. Braz. Acad. Sci. 86, 265e269.
De Ley, P., Blaxter, M., 2004. A new system for Nematoda: combining morphological char-
acters with molecular trees, and translating clades into ranks and taxa. In: Cook, R.,
Hunt, D.J. (Eds.), Nematology Monographs and Perspectives, vol. 2. E.J. Brill, Leiden,
the Netherlands, pp. 633e653.
Disney, R.H.L., 1994. Scuttle Flies: The Phoridae. Chapman & Hall, London.
Dubinin, V.B., 1948. The presence of Pleistocene lice (Anoplura) and nematodes in the
cadaver of a fossilized gopher. Dokl. Akad. Nauk. SSSR 62, 417e420 (in Russian).
The Geological Record of Parasitic Nematode Evolution 87

Dubinina, M.N., 1972. The nematode Alfortia edentatus (Looss, 1900) from the intestine of an
upper Pleistocene horse. Parazitologia 6, 441e443 (in Russian).
Dupain, J., Nell, C., Petrzelkova, K.J., Garcia, P., Modry, D., Gordo, F.P., 2009. Gastroin-
testinal parasites of bonobos in the Lomako Forest, Democratic Republic of Congo. In:
Huffman, M.A., Chapman, C.A. (Eds.), Primate Parasite Ecology, the Dynamics and
Study of Host-Parasite Relationships. Cambridge University Press, Cambridge,
pp. 297e310.
Ebbell, B., 1937. The Papyrus Ebers, the Greatest Egyptian Medical Document. Levin and
Munksgaard, Copenhagen.
Eyualim-Abebe, Andrassy, I., Traunspurger, W., 2006. Freshwater Nematodes, Ecology and
Taxonomy. CABI Publishing, Wallingford, UK.
ujo, A., Confalonieri, U., Filho, C.M.R., 1987. The finding of hookworm
Ferreira, L.F., Ara
eggs in human coprolites from 7230  80 years BP, from Piauí, Brasil. An. Acad. Bras.
Cienc. 59, 280e281.
ujo, A., Confalonieri, U., Chame, M., Gomes, D.G., 1991. Trichuris eggs in
Ferreira, L.F., Ara
animal coprolites dated from 30,000 years ago. J. Parasitol. 73, 491e493.
Ferreira, L.F., Ara ujo, A., Duarte, A.N., 1993. Nematode larvae in fossilized
animal coprolites from lower and middle Pleistocene sites, central Italy. J. Parasitol.
79, 440e442.
Fortuner, R., Maggenti, A.R., 1987. A reappraisal of Tylenchina (Nemata). 4. The family
Anguinidae Nicoll, 1936 (1926). Rev. Nématol. 10, 163e176.
Gemmill, J.F., Von Linstow, O., 1902. Ichthyonema grayi Gemmill & von Linstow. Arch.
Naturg. 1, 113e118.
Gonçalves, M.L.C., Ara ujo, A., Ferreira, L.F., 2003. Human intestinal parasites in the past:
new findings and a review. Mem. Inst. Oswaldo Cruz 98, 103e118.
Grabenhorst, H., 1985. Eine zweite Bremse (Tabanidae) zusammen mit ihrem Parasiten
(Nematoda, Mermithoidae) aus dem oberplioz€an von Willershausen, Krs. Osterode.
Osterode Aufschluss 36, 325e328.
Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G., 2012. The Geologic Time Scale 2012.
Elsevier, Boston.
Hallett, S.L., Erséus, C., O’Donoghue, P.J., Lester, R.J.G., 2001. Parasitic fauna of Australian
Maine oligochaetes. Mem. Queensl. Mus. 46, 555e576.
Harwood, P.D., 1930. A new species of Oxysomatium (Nematoda) with some remarks on the
genera Oxysomatium and Aplectana and observations on the life history. J. Parasitol. 17,
61e73.
Helden, A.J., 2008. First extant records of mermithid nematode parasites of Auchenorrhyn-
cha in Europe. J. Invertebr. Pathol. 99, 351e353.
von Heyden, C., 1860. Mermis antiqua, ein fossiler Eigeweidewurm. Stettiner Entomol. Ztg.
21, 38.
Hope, W.D., Murphy, D.G., 1969. Syringonomous typicus, new genus, new species (Enoplida:
Leptosomatidae) a marine nematode inhabiting arenaceous tubes. Proc. Biol. Soc. Wash-
ington 82, 511e518.
Hope, W.D., Tchesunov, A.V., 1999. Smithsoninema inaequale n. g. and n. sp. (Nematoda:
Leptolaimidae) inhabiting the test of a foraminiferan. Invertebr. Biol. 118, 95e108.
Horne, P.D., 2002. First evidence of enterobiasis in ancient Egypt. J. Parasitol. 88, 1019e
1021.
Hugot, J.-P., Gardner, S.L., Borba, V., Araujo, P., Leles, D., Da-Rosa, A.A.S., Dutra, J.,
Ferreira, L.F., Araujo, A., 2014. Discovery of a 240 million year old nematode parasite
egg in a cynodont coprolite sheds light on the early origin of pinworms in vertebrates.
Parasites Vectors 7, 486e494.
Hunt, D.J., 1993. Aphelenchida, Longidoridae and Trichodoridae, Their Systematics and
Bionomics. CAB International, Wallingford.
88 George O. Poinar, Jr.

Inglis, W.G., 1965. Patterns of evolution in parasitic nematodes. In: Taylor, A.E.R. (Ed.),
Evolution of Parasites. Blackwell Scientific Publications, Oxford, pp. 79e124.
Iturralde-Vinent, M.A., 2001. Geology of the amber-bearing deposits of the Greater Antilles.
Caribbean J. Sci. 37, 141e167.
Iturralde-Vinent, M.A., MacPhee, R.D.E., 1996. Age and paleogeographic origin of
Dominican amber. Science 273, 1850e1852.
Jansen Jr., J., Over, H.J., 1962. Het voorkomen van parasieten in terpmateriaal uit Noordwest
Duitsland. Tijdschr. Diergeneeskd. 87, 1377e1379.
Jansen Jr., J., Over, H.J., 1966. Observations on helminth infections in a Roman army-camp.
In: Proceedings of the First International Congress of Parasitology, Rome, 1964, p. 791.
Jansson, H.-B., Poinar Jr., G.O., 1986. Some possible fossil nematophagous fungi. Trans. Br.
Mycol. Soc. 87, 471e474.
Jones, G.M., Hagen, N.T., 1987. Echinomermella matsi sp. n., an endoparasitic nematode
from the sea urchin Strongylocentrotus droebachiensis in Northern Norway. Sarsia 72,
203e212.
Kaur, T., Singh, J., 2009. Primate-parasitic zoonoses and anthropozoonoses: a literature
review. In: Huffman, M.A., Chapman, C.A. (Eds.), Primate Parasite Ecology, the
Dynamics and Study of Host-Parasite Relationships. Cambridge University Press, Cam-
bridge, pp. 199e230.
Klicks, M., 1990. Helminths as heirlooms and souveniers: a review of new world
paleoparasitology. Parasitol. Today 6, 93e100.
Larsson, S.G., 1978. Baltic amber- a Palaeobiological Study. Entomonograph 1, 1e192.
Lehmann, U., Hillmer, G., 1983. Fossil Invertebrates. Cambridge University Press,
Cambridge.
Littlewood, D.T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
Luc, M., 1987. A reappraisal of Tylenchina (Nemata). 7. The family Pratylenchidae Thorne,
1949. Rev. Nématol. 10, 203e218.
Maggenti, A.R., 1971. Nemic relationships and the origins of plant parasitic nematodes. In:
Zuckerman, B.M., Rhode, R.A. (Eds.), Plant Parasitic Nematodes, vol. 1. Academic
Press, New York, pp. 65e81.
McClelland, G., 2005. Nematodes (roundworms). In: Rohde, K. (Ed.), Marine Parasitology.
CABI Publishing, Wallingford, pp. 104e115.
Menge, A., 1866. Ueber ein Rhipidopteron und einige andere im Bernstein eingeschlossene
Tiere. Schriften der Naturforschenden Gesellschaft in Danzig, Neue Folge 1, 1e8.
Montenegro, A., Araujo, A., Eby, M., Ferreira, L.F., Hetherington, R., Weaver, A.J., 2006.
Parasites, Paleoclimate, and the Peopling of the Americas; using the hookworm to time
the Clovis Migration. Curr. Anthropol. 47, 193e200.
Myers, E.H., 1943. Life activities of Foraminifera in relation to marine ecology. Proc. Am.
Philos. Soc. 86, 439e458.
Osche, G., 1963. Morphological, biological, and ecological considerations in the phylogeny
of parasitic nematodes. In: Dougherty, E.C. (Ed.), The Lower Metazoa: Comparative
Biology and Phylogeny. University of California Press, Berkeley, pp. 283e302.
Paramonov, A.A., 1962. Principles of Plant Nematology, vol. 1. Akademii Nauk SSSR,
Moscow.
Patrucco, R., Tello, R., Bonavia, D., 1983. Parasitological studies of coprolites of pre-hispanic
Peruvian populations. Curr. Anthropol. 24, 393e394.
Petter, A.J., 1980. Une nouvelle famille de Nématodes parasites d’Invertébrés marins, les
Benthimermithidae. Ann. Parasitol. Hum. Comp. 55, 209e224.
Petter, A.J., 1983. Quelques nouvelles espéces du genre Benthimermis Petter, 1980 (Benthi-
mermithidae, Nematoda) du Sud de l’Océan Indien. Syst. Parasitol. 5, 4e15.
Poinar Jr., G.O., 1975. Entomogenous Nematodes. Brill, Leiden.
The Geological Record of Parasitic Nematode Evolution 89

Poinar Jr., G.O., 1977. Fossil nematodes from Mexican amber. Nematologica 23, 232e238.
Poinar Jr., G.O., 1978a. Mesidionema praecomasculatis gen. et sp. n.: Mesidionematidae Fam. N.
(Drilonematoidea: Rhabditida), a nematode parasite of earthworms. Proc. Helminthol.
Soc. Washington 45, 97e102.
Poinar Jr., G.O., 1978b. Thelastoma endoscolicum sp. n. (Oxyurida: Nematoda) a parasite of
earthworms (Oligochaeta: Annelida). Proc. Helminthol. Soc. Washington 45, 92e96.
Poinar Jr., G.O., 1979. Parasitodiplogaster sycophilon gen. n., sp. n. (Diplogasteridae: Nema-
toda), a parasite of Elisabethiella stuckenbergi Grandi (Agaionidae: Hymenoptera) in
Rhodesia. Proc. K. Ned. Akad. Wet. 82, 375e381.
Poinar Jr., G.O., 1983. The Natural History of Nematodes. Prentice Hall, Inc., Englewood
Cliffs.
Poinar Jr., G.O., 1984. First fossil record of parasitism by insect parasitic Tylenchida (Allan-
tonematidae: Nematoda). J. Parasitol. 70, 306e308.
Poinar Jr., G.O., 1985. Nematode parasites and infectious diseases of Tabanidae (Diptera).
Myia 3, 599e616.
Poinar Jr., G.O., 1991. The mycetophagous and entomophagous stages of Iotonchium califor-
nicum n. sp. (Iotonchidae: Tylenchida). Rev. Nématol. 14, 565e580.
Poinar Jr., G.O., 1992. Life in Amber. Stanford University Press, Stanford.
Poinar Jr., G.O., 1993. Origins and phylogenetic relationships of the entomophilic rhabdi-
tids, Heterorhabditis and Steinernema. Fundam. Appl. Nematol. 16, 333e338.
Poinar Jr., G.O., 2000. Heydenius araneus n.sp. (Nematoda: mermithidae), a parasite of a fossil
spider, with an examination of helminths from extant spiders (Arachnida: Araneae).
Invertebr. Biol. 119, 388e393.
Poinar Jr., G.O., 2001. Heydenius brownii (Nematoda: Mermithidae) parasitizing a planthop-
per (Achilidae: Homoptera) in Baltic amber. Nematology 3, 753e757.
Poinar Jr., G.O., 2002. First fossil record of nematode parasitism of ants; a 40 million year tale.
Parasitology 125, 457e459.
Poinar Jr., G.O., 2003. Fossil evidence of phorid parasitism (Diptera: Phoridae) by allantone-
matid nematodes (Tylenchida: Allantonematidae). Parasitology 127, 589e592.
Poinar Jr., G.O., 2011. The Evolutionary History of Nematodes. Brill Academic Publishers,
Leiden.
Poinar Jr., G.O., 2012. New fossil nematodes in Dominican and Baltic amber. Nematology
14, 483e488.
Poinar Jr., G.O., 2014. Evolutionary history of terrestrial pathogens and endoparasites as
revealed in fossils and subfossils. Adv. Biol. 2014, 1e29.
Poinar Jr., G.O., Boucot, A., 2006. Evidence of intestinal parasites of dinosaurs. Parasitology
133, 245e249.
Poinar Jr., G.O., Buckley, R., 2006. Nematode (Nematoda: Mermithidae) and hairworm
(Nematomorpha: Chordodidae) parasites in early cretaceous amber. J. Invertebr. Pathol.
93, 36e41.
Poinar Jr., G.O., Herre, E.A., 1991. Speciation and adaptive radiation in the Fig Wasp nem-
atode, Parasitodiplogaster (Diplogasteridae: Rhabditida), in Panama. Rev. Nématol. 14,
361e374.
Poinar Jr., G.O., Lane, R.S., 1978. Pheromermis myopis sp. n. (Nematoda: Mermithidae), a
parasite of Tabanus punctifer (Diptera: Tabanidae). J. Parasitol. 64, 440e444.
Poinar Jr., G.O., Milki, R., 2001. Lebanese Amber. Oregon State University Press, Corvallis.
Poinar Jr., G.O., Monteys, S.I., 2008. Mermithids (Nematoda: Mermithidae) of biting
midges (Diptera: Ceratopogonidae): Heleidomermis cataloniensis n. sp. from Culicoides cir-
cumscriptus Kieffer in Spain and a species of Cretacimermis Poinar, 2001 from a ceratopo-
gonid in Burmese amber. Syst. Parasitol. 69, 13e21.
Poinar Jr., G.O., Poinar, R., 1999. The Amber Forest. Princeton University Press, Princeton,
New Jersey.
90 George O. Poinar, Jr.

Poinar Jr., G.O., Poinar, R., 2003. Description and development of Gastromermis anisotis sp.
n. (Nematoda: Mermithidae) a parasite in a quadritrophic system involving a cyanobac-
terium, midge and virus. Nematology 5, 325e338.
Poinar Jr., G.O., Yanoviak, S.P., 2008. Myrmeconema neotropicum n. g., n. sp., a new tetrado-
nematid nematode parasitizing South American populations of Cephalotes atratus (Hyme-
noptera: Formicidae), with the discovery of an apparent parasite-induced host morph.
Syst. Parasitol. 69, 145e153.
Poinar Jr., G.O., Acra, A., Acra, F., 1994. Earliest fossil nematode (Mermithidae) in Creta-
ceous Lebanese amber. Fundam. Appl. Nematol. 17, 475e477.
Poinar Jr., G.O., Jaenke, J., Dombeck, I., 1997. Parasitylenchus nearcticus sp.n. (Tylenchida:
Allantonematidae) parasitizing Drosophila (Diptera: Drosophilidae) in North America.
Fundam. Appl. Nematol. 20, 187e190.
Poinar Jr., G.O., Latham, D.M., Poulin, R., 2002. Thaumamermis zealandica n. sp. (Mermi-
thidae: Nematoda) parasitising the intertidal marine amphipod Talorchestia quoyana (Tali-
tridae: Amphipoda) in New Zealand, with a summary of mermithids infecting
amphipods. Syst. Parasitol. 53, 227e233.
Poinar Jr., G.O., Lauchaud, J.-P., Castillo, A., Infante, F., 2006. Recent and fossil nematode
parasites (Nematoda: Mermithidae) of Neotropical ants. J. Invertebr. Pathol. 91, 19e26.
Poinar Jr., G.O., Kerp, H., Hass, H., 2008. Palaeonema phyticus gen. n., sp. n. (Nematoda:
Palaeonematidae fam. n.), a Devonian nematode associated with early land plants.
Nematology 10, 9e14.
Poinar Jr., G.O., Durate, D., Santos, M.J., 2009. Halomonhystera parasitica n. sp. (Nematoda:
Monhysteridae), a parasite of Talorchestia brito (Crustacea: Talitridae) in Portugal. Syst.
Parasitol. 75, 53e58.
Poinar Jr., G.O., Lewis, S., Hagen, N., Hyman, B., 2011. Systematic affinity of the sea urchin
parasite, Echinomermella matsi Jones & Hagen (Enoplida: Echinomermellidae). Nema-
tology 13, 747e753.
Poinar Jr., G.O., Weinstein, S.B., Garcia-Verdrenne, A.E., Kuris, A.M., 2014. First descrip-
tion of a nematode, Spinitectus gabata n. sp. (Spirurina: Cystidicolidae), from the deep sea
oarfish, Regalecus russelii (Regalecidae) in Japan. Int. J. Nematol. 24, 117e123.
Rasnitsyn, A.P., Quicke, D.L.J., 2002. History of Insects. Kluver Academic Publishers,
Dordrecht.
Reinhard, K.J., 1992. Parasitology as an interpretive tool in Archaeology. Am. Antiq. 57,
231e243.
Reinhard, K.J., Ara ujo, A., 2008. Archaeoparasitology. In: Pearsall, D.M. (Ed.), Encyclo-
pedia of Archaeology. Elsevier, Amsterdam, pp. 494e501.
Richardson, P.N., Hesling, J.J., Riding, I.L., 1977. Life cycle and description of Howardula
husseyi n. sp. (Tylenchida: Allantonematidae), a nematode parasite of the mushroom
phorid, Megaselia halterata (Phoridae, Diptera). Nematologica 23, 217e231.
Ringuelet, R.A., 1957. Restos de probables huevos de Nematodes en el estiercol del eden-
tado extinguido Mylodon listae (Ameghino). Ameghiniana 1, 15e16.
Rubtsov, I.A., 1977. A new genus and species of parasitic nematode, Ananus asteroideus
(Nematoda, Marimermithidae), from the asteroid Diplopteraster perigrinator. Bull. Mus.
Natl. His. Nat. 496, 1113e1117.
Rubtsov, I.A., 1980. The new marine parasitic nematode, Abos bathycola, from Priapulids and
the taxonomic position of the family Marimermithidae in the class Nematoda. Parasito-
logia 14, 177e181 (in Russian).
Rubstov, I.A., Platonova, T.A., 1974. Une nouvelle famille de Nematodes parasites marins.
Zool. Zh. 53, 1445e1458.
Rota-Stabelli, O., Daley, A.C., Pisani, D., 2013. Molecular timetrees reveal a Cambrian
colonization of land and a new scenario for ecdysozoan evolution. Curr. Biol. 23,
392e398.
The Geological Record of Parasitic Nematode Evolution 91

Sandison, A.T., Tapp, E., 1998. Disease in ancient Egypt. In: Cockburn, A., Cockburn, E.,
Reyman, T.A. (Eds.), Mummies, Disease and Ancient Cultures, Second ed. Cambridge
University Press, Cambridge, pp. 38e58.
Schlee, D., 1990. Das Bernstein-Kabinett. Stuttgart Beitr. Naturkd. C 28, 1e100.
Schlinger, E.I., 1975. Intertidal insects; order Diptera. In: Smith, R.I., Carlton, J.T. (Eds.),
Light’s Manual: Intertidial Invertebrates of the Central California Coast, third ed. Uni-
versity of California Press, Berkeley, pp. 436e446.
Schmidt, G.D., Duszynski, D.W., Martin, P.S., 1992. Parasites of the extinct Shasta ground
sloth, Nothrotheriops shastensis, in Rampart cave, Arizona. J. Parasitol. 78, 811e816.
Schwenk, J., 1927. Sobre un nematoid parasita de Oniscidae. Bol. Biol. 7, 78e80.
Siddiqi, M.R., 1983. Evolution of plant parasitism in nematodes. In: Stone, A.R.,
Platt, H.M., Khalil, L.F. (Eds.), Concepts in Nematode Systematics. Academic Press,
New York, pp. 113e129.
Sliter, W.V., 1971. Predation on benthic Foraminifers. J. Foraminifera Res. 1, 20e29.
Sudhaus, W., 2008. Evolution of insect parasitism in rhabditid and diplogastrid nematodes.
In: Makarov, S.E., Dimitrijevic, R.N. (Eds.), Advances in Arachnology and Develop-
mental Biology: Papers Dedicated to Professor Bozidar P.M. Curcic. Institute of
Zoology, Belgrade, pp. 117e130.
Sudhaus, W., 2010. Preadaptive plateau in Rhabditida (Nematoda) allow the repeated evo-
lution of zooparasites, with an outlook on evolution of life cycles with Spiroascarida.
Palaeodiversity 3, 117e130.
Sung, G.-H., Poinar Jr., G.O., Spatafora, J.W., 2008. The oldest fossil evidence of animal
parasitism by fungi supports a Cretaceous diversification of fungalearthropod
symbioses. Mol. Phylogenet. Evol. 49, 495e502.
Szadziewski, R., Poinar Jr., G.O., 2005. Additional biting midges (Diptera: Ceratopogoni-
dae) from Burmese amber. Pol. Pismo Entomol. 74, 349e362.
Tang, Y., Hominick, W.M., Killick-Kendrick, R., Killick-Kendrick, M., Page, A.M.,
1993. Didilia ooglypta n. gen., n. sp. (Tetradonematidae: Mermithoidea: Nematoda),
a parasite of phlebotomine sandflies in Afghanistan. Fundam. Appl. Nematol. 16,
325e331.
Tapp, E., Wildsmith, K., 1993. The autopsy and endoscopy of the Leeds Mummy. In:
David, A.R., Tapp, E. (Eds.), The Mummy’s Tale. St. Martins Press, New York,
pp. 132e153.
Taylor, A.L., 1955. Helminths in mediaeval remains. Vet. Rec. 67, 216e218.
Taylor, T.N., Krings, M., Taylor, E.L., 2015. Fossil Fungi. Elsevier, Amsterdam.
Tchesunov, A.V., Spiridonov, S.E., 1993. Nematimermis enoplivora gen. n., sp. n. (Nema-
toda: Mermithoidea) from marine free-living nematodes Enoplus spp. Russ. J. Nem-
atol. 1, 7e16.
Timm, R.W., Franklin, M.T., 1969. Two marine species of Aphelenchoides. Nematologica
15, 370e375.
Vanderburgh, D.J., Anderson, R.C., 1987. The relationship between nematodes of the genus
Cosmocercoides Wilkie, 1930 (Nematoda: Cosmocercoidea) in toads (Bufo americanus) and
slugs (Deroceras laeve). Can. J. Zool. 65, 1650e1661.
Ward, H.B., 1933. On Thalassonema ophioctinis, a nematode parasitic in the brittle star
Ophiocten amitinum. J. Parasitol. 19, 262e268.
Wilson, H.M., Anderson, L.I., 2004. Morphology and taxonomy of Paleozoic millipedes
(Diplopoda: Chilognatha: Archipolypoda) from Scotland. J. Paleontol. 78, 169e184.
Witenberg, G., Gerichter, C., 1944. The morphology and life history of Foleyella duboisi with
remarks on allied filariids of Amphibia. J. Parasitol. 30, 245e254.
Wood, J.R., Wilmshurst, J.M., Rawlence, N.J., Bonner, K.I., Worthy, T.H., Kinsella, J.M.,
Cooper, A., 2013. A megafauna’s microfauna: gastrointestinal parasites of New Zealand’s
extinct moa (Aves: Dinornithiformes). PLoS One 8, e57315.
92 George O. Poinar, Jr.

Yeates, G., Boag, B., 2006. Female size shows similar trends in all clades of the phylum
Nematoda. Nematology 8, 111e127.
Yeates, G.W., Spiridonov, S.E., Blakemore, R., 1998. Plesioungella kathleenae gen. n. et sp. n.
(Nematoda: Drilonematoidea) from the Australian endemic megascocecid earthworm
Fletcherodrilus unicus (Fletcher, 1899). N. Z. J. Zool. 25, 205e212.
CHAPTER THREE

Constraining the Deep Origin of


Parasitic Flatworms and Host-
Interactions with Fossil Evidence
Kenneth De Baets*, 1, Paula Dentzien-Diasx, Ieva Upeniece{,
Olivier Verneaujj, #, **, Philip C.J. Donoghuexx
*Fachgruppe Pal€aoUmwelt, GeoZentrum Nordbayern, Friedrich-Alexander-Universit€at
Erlangen-N€ urnberg, Erlangen, Germany
x
Nucleo de Oceanografia Geol ogica, Instituto de Oceanografia, Universidade Federal do Rio Grande,
Rio Grande, Brazil
{
Department of Geology, University of Latvia, Riga, Latvia
jj
Centre de Formation et de Recherche sur les Environnements Méditerranéens, University of Perpignan Via
Domitia, Perpignan, France
#
CNRS, Centre de Formation et de Recherche sur les Environnements Méditerranéens, Perpignan, France
**Unit for Environmental Sciences and Management, North-West University, Potchefstroom, South Africa
xx
School of Earth Sciences, University of Bristol, Life Science Building, Bristol, UK
1
Corresponding author: E-mail: kenneth.debaets@fau.de

Contents
1. Introduction 94
2. Assessment of the Flatworm Fossil Record 96
2.1 Devonian fossil hook circlets 97
2.2 Silurian blister pearls and calcareous concretions in bivalve shells 101
2.3 Permo-Carboniferous egg remains in shark coprolites 103
2.4 Cretaceous egg remains in terrestrial archosaur coprolites 105
2.5 Eocene shell pits in intermediate bivalve hosts 106
2.6 Eggs remains in a Pleistocene mammal coprolite 107
2.7 Holocene evidence for parasitic flatworms from ancient remains 107
2.8 Free-living flatworms 108
3. Interpolating or Extrapolating Extant ParasiteeHost Relationships and the 110
Assumption of ParasiteeHost Coevolution
4. Molecular Clock Studies 113
5. Conclusions and Future Prospects 119
Acknowledgements 121
References 122

Abstract
Novel fossil discoveries have contributed to our understanding of the evolutionary
appearance of parasitism in flatworms. Furthermore, genetic analyses with greater
coverage have shifted our views on the coevolution of parasitic flatworms and their
hosts. The putative record of parasitic flatworms is consistent with extant host
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.06.002 All rights reserved. 93
94 Kenneth De Baets et al.

associations and so can be used to put constraints on the evolutionary origin of the par-
asites themselves. The future lies in new molecular clock analyses combined with addi-
tional discoveries of exceptionally preserved flatworms associated with hosts and
coprolites. Besides direct evidence, the host fossil record and biogeography have the
potential to constrain their evolutionary history, albeit with caution needed to avoid
circularity, and a need for calibrations to be implemented in the most conservative
way. This might result in imprecise, but accurate divergence estimates for the evolution
of parasitic flatworms.

1. INTRODUCTION
Parasitic flatworms (Platyhelminthes: Neodermata) are a highly
diverse group containing many parasites of biomedical, veterinary and eco-
nomic importance (Olson and Tkach, 2005; Littlewood, 2006). Time con-
straints on the origin and evolution of parasitism in this group are still poorly
resolved due to their patchy and largely overlooked fossil record (Littlewood
and Donovan, 2003; Littlewood, 2006). The fossil record of parasitic
flatworms is often disregarded by parasitologists (Combes, 2001; Littlewood,
2006; Verneau et al., 2009a; Badets et al., 2011) and evolutionary (paleo)bi-
ologists (Labandeira, 2002; Erwin et al., 2011; Wey-Fabrizius et al., 2013)
alike. Most parasitologists have therefore focused on extrapolating or inter-
polating extant parasiteehost associations to infer information on the
evolution history of parasitic flatworms (Llewellyn, 1987; Brooks, 1989;
Brooks and McLennan, 1993; Boeger and Kritsky, 1997; Hoberg, 1999;
Hoberg et al., 1999; Littlewood et al., 1999a). Nevertheless, the last two
decades have seen several new fossil discoveries, which have extended the
record from certain lineages of parasitic flatworms deeper into the Cenozoic
( Jouy-Avantin et al., 1999; Todd and Harper, 2011) or from the Cenozoic
to the Mesozoic (Poinar and Boucot, 2006) or even the Paleozoic
(Upeniece, 2001, 2011; Dentzien-Dias et al., 2013).
Furthermore, molecular analyses have considerably shifted our views on
flatworm phylogeny (Lockyer et al., 2003a; Olson and Tkach, 2005;
Littlewood, 2008; Perkins et al., 2010; Laumer and Giribet, 2014) with
implications for older hypotheses of parasiteehost coevolution. Due to
the patchy fossil record, establishing the phylogeny of flatworms is particu-
larly important for establishing a timeline for the group. Traditionally, Mon-
ogenea (ectoparasitic with simple life cycles) and Cestoda (endoparasitic with
complex, trophically transmitted, life cycles) were often grouped based on
morphological similarities of their larval stages (Bychowsky, 1937; Brooks,
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 95

1989), sometimes referred to as the cercomer theory (Lockyer et al., 2003a).


However, most recent molecular-based analyses (Mollaret et al., 1997;
Lockyer et al., 2003a; Park et al., 2007; Perkins et al., 2010; Wey-Fabrizius
et al., 2013; Hahn et al., 2014; Egger et al., 2015) and those including alter-
native data such as microRNAs (Fromm et al., 2013) indicate a sister-group
relationship between cestodes and trematodes (rather than between Mono-
genea and Cestoda) with one possible exception (Laumer and Giribet, 2014;
Laumer et al., 2015), although no clear morphological characters support
this arrangement. Interestingly, the monophyly of Monogenea i.e. the sis-
ter-group relationship between Monopisthocotylea and Polyopisthocotylea
based on morphological data (Boeger and Kritsky, 2001), is not always sup-
ported by molecular phylogenetics (Mollaret et al., 1997; Justine, 1998;
Lockyer et al., 2003a) either. Based on recent results employing mitoge-
nomic data, primitive parasitic flatworms (Neodermata) were probably ecto-
parasitic with a simple life cycle on vertebrates (Park et al., 2007) and
engaged in epithelial feeding (Perkins et al., 2010). Subsequently, it has
been proposed they added an intermediate host to their life cycle, probably
first a crustacean intermediate host in Cestoda and a mollusc intermediate
host in Trematoda (Park et al., 2007) before switching to a blood diet
(Perkins et al., 2010). This differs from previous hypotheses (Brooks,
1989; Littlewood et al., 1999a), where a sister-group relation between Mon-
ogenea and Cestoda, and between these taxa and Trematoda was postulated.
This led to two now probably outdated hypotheses of interpreting the life
cycle evolution of Neodermata (Park et al., 2007). One hypothesis
(Littlewood et al., 1999a) suggested that proto-neodermatan first acquired
an endoparasitic association with vertebrates and that independent adoptions
of invertebrates by the Trematoda (molluscs) and Cestoda (crustaceans) as
well as ectoparasitism in Monogenea were subsequent acquisitions (verte-
brate first hypothesis). Another hypothesis (Cribb et al., 2001) suggested
that the association of common ancestor of the Trematoda with molluscan
hosts was primitive (acquiring its subsequent vertebrate hosts indepen-
dently), and that the vertebrates were involved in the life cycle of the
common ancestor of the Monogenea þ Cestoda clade as independent initial
hosts apart from that of trematodes, with the crustaceans as subsequent inter-
mediate hosts adopted by the Cestoda groups after the ancestral cestode
diverged from the monogeneans (mollusk first hypothesis).
The closest free-living relatives of helminths are also important in con-
straining divergence times in their evolutionary history (Littlewood et al.,
1999b; Near, 2002; Littlewood, 2006). However, the phylogeny of
96 Kenneth De Baets et al.

free-living flatworms has proven even more problematic (but see Littlewood
and Waeschenbach, 2015 for a review of recent advances) and has been
further complicated by the fact that some extant forms traditionally included
in Platyhelminthes have been excluded from the phylum based on molec-
ular analyses (Jondelius et al., 2002; Telford et al., 2003; Willems et al.,
2006; Wallberg et al., 2007; Hejnol et al., 2009) such as the Acoela
(Ruiz-Trillo et al., 1999; Mwinyi et al., 2010; Philippe et al., 2011) and Xen-
oturbella (Bourlat et al., 2003). Furthermore, platyhelminths have often been
grouped in the Platyzoa (Cavalier-Smith, 1998) together with various other
taxa including Acanthocephala, which have convergently evolved a parasitic
lifestyle with larval stages and have been shown to be closely related with
free-living Rotifera (Near, 2002; Weber et al., 2013). It remains unclear
whether Platyzoa is a clade or an artificial grouping generated by systematic
error and long-branch attraction artefacts (Edgecombe et al., 2011; Wey-
Fabrizius et al., 2013; Struck et al., 2014), since subsequent studies have
not only disagreed on the membership of the phyla, but also on the relation-
ships within this grouping. This makes new fossil discoveries of parasitic flat-
worms not only relevant in constraining the evolutionary origin of
flatworms, but also that of the Platyzoa as a whole.
Here we review the potential of fossil flatworm evidence with a view to
using these data to constrain the timescale for the evolutionary history of this
group and outline how they can be used to improve our understanding of
the evolutionary radiation of the Neodermata. We consider whether these
rare fossil finds are at least consistent with coevolution of parasitic flatworms
and their hosts, as well as how fossil finds and other geological evidence in
combination with molecular clock methodology can be best used to
constrain the temporal framework for the evolution of parasitic flatworms.
Such a temporal framework is a key to test evolutionary hypotheses
regarding the origin and diversification of parasitism and its coincidence
with certain biogeographic events, major environmental changes or key
ecological or evolutionary events in the evolution of their hosts.

2. ASSESSMENT OF THE FLATWORM FOSSIL RECORD


Fossil evidence for parasitic flatworms can be derived from (1) rare
exceptionally preserved body fossils, which can be isolated (Poinar and
Boucot, 2006; Dentzien-Dias et al., 2013) or remain associated with their
hosts (Upeniece, 2001, 2011) or (2) more commonly occur as characteristic
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 97

traces or skeletal pathologies in their (intermediate) hosts, which have the


potential to be traced back in the fossil record (Ruiz and Lindberg, 1989;
Ruiz, 1991; Ituarte et al., 2001, 2005; Huntley, 2007; Todd and Harper,
2011; Huntley and Scarponi, 2012; Huntley et al., 2014; Huntley and Scar-
poni, 2015; Huntley and De Baets, 2015). Parasite body fossils are scarce due
to their small size, lack of hard parts and the residence within the host and/or
isolation from their hosts (Conway Morris, 1981; Littlewood and Donovan,
2003; De Baets et al., 2011). The rarity of fossilized parasiteehost associa-
tions and the fact that culprits of traces or pathologies in the skeletons of their
hosts are often hard to identify can make it difficult to infer parasiteehost
associations from the fossil record. Nevertheless, it is the only direct evidence
for the presence of such associations in the geological past.

2.1 Devonian fossil hook circlets


Circlets of fossil hooks described from the Devonian of Latvia (Upeniece,
1996, 1998, 1999, 2001, 2011) are the oldest potential body fossil evidence
for parasitic flatworms. Upeniece (2001, 2011) discovered about 77 cir-
clets, which were mostly attached or closely associated with fossil gnathos-
tomes (16 juveniles of the antiarch placoderm Asterolepis ornata; 27
specimens of the acanthodian Lodeacanthus gaujicus: Figures 1(b, c, e, f)).
However, one isolated circlet was found close to a specimen of a clam
shrimp (Figure 1(g)) and two other circlets were found associated with
another crustacean arthropod (Mysidacea: Figure 1(e)), but the hooks are
too large to indicate parasitism based on the size of the crustaceans.
The location of these remains in fossils of their vertebrate hosts (Figure 1(b)
and (c)) and their similarity to the hooks of parasitic helminths, strongly sug-
gest a parasitic nature (Upeniece, 2011). In acanthodians (L. gaujicus), they
are associated with the gill regions, near the fin spines, and in the abdominal
region near the scapula (Figure 1(d) and (f)), while in placoderms (A. ornata)
their location is not so well determined (see Figure 1(a)). The length of
infested fishes varies between 1 and 4 cm (Figure 1(a) and (d)). Several of
them were infested with 2e9 parasites (7 hook circlets can be counted in
the specimen figured in Figure 1(b)). Most authors agree that they are the
remains of parasitic helminths, although their exact affinity remains the sub-
ject of debate (Upeniece, 2001; Littlewood and Donovan, 2003; Upeniece,
2011). These are reminiscent of hooks which are used by Neodermata
(Monogenea, Cestoda) and Acanthocephala to attach themselves to their
hosts. Differences in morphology and their location on the host body
(Upeniece, 2011) might even indicate that they belong to different groups
98 Kenneth De Baets et al.

(a) (b) (c)

(d)

(e)

(f) (g)
Figure 1 Fossil helminth remains in Middle Devonian gnathostomes (Upeniece, 2001,
2011) (Modified from Upeniece (2011) unless otherwise stated.): (a) Locations where hel-
minth remains were found on juveniles of the placoderm Asterolepis ornata. (b) Multiple
fossil circlets of parasitic helminth hooks (marked with ellipses in (a)) found inside the
acanthodian Lodeacanthus gaujicus, LDM 270/18c (Upeniece, 2001, Pl. 3, Figure 2); (c)
Close-up on the hook circlet found associated with L. gaujicus, specimen LDM 270/
33; (d) Locations where helminth remains were found in juvenile and adults of the
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 99

of parasites (Upeniece, 2001), including ectoparasites (e.g. monogeneans)


and endoparasites (e.g. cestodes and acanthocephalans).
The circular arrangement of the bilaterally symmetrically located hooks,
traces of cuticular disc as well as the maximum number of 16 hooks, which is
characteristic for early divergent monogeneans (Boeger and Kritsky, 1993),
indicates that at least some of them represent Monogenea (Combes, 2001;
Upeniece, 2011), although the larvae of Cestoda can also have radially
arranged chitinous hooks for attachment. The hooks are also considerably
larger (length: 0.02e0.40 mm) than those of extant flatworms, but this
does not necessarily rule out a monogenean affinity. Poulin (2005), for ex-
ample, could demonstrate an evolutionary trend towards decreasing body
size in extant ectoparasitic Monogenea, while this was less clear for derived
endoparasitic flatworms (Digenea). Interestingly, Upeniece (2011) observed
that small-sized acanthodians possessed small hook systems, while larger
acanthodians exhibited the greatest range in size of the hook systems. These
might indicate that parasites might have spent all their life in one host, which
might further corroborate a similarity to early divergent parasitic flatworms.
The lack of fossilized soft body parts, with the exception of traces of the disc
outline, further hampers a more precise taxonomic assignment. Upeniece
(1999, 2011) suggested that at least two morphological groups are presented
in both species of fish: hooks with a ‘handle’ and hooks without a ‘handle’.
She noticed that the hooks with ‘handles’ typically occur in the abdominal
region of acanthodians and placoderms, suggestive of a close affinity with
endoparasites such as Acanthocephala or Cestoda. The elongated tube-
like body of acanthocephalans typically possesses a thorny proboscis, which
is an anterior retractile organ bearing a large number of hooks (Bush et al.,
2001). This holdfast organ is only fully evaginated after death and resembles
the ‘introvert’ of Rotatoria, Priapulida, Kinorhyncha and Nematomorpha
larvae (Taraschewski, 2005). Larval forms of Cestoda typically bear hooks,
present even in basal extant Cestoda like Gyrocotylidae (Xylander, 2005)
and Amphilinidea (typically 10), which can be retained in the adult forms

=
acanthodian L. gaujicus; (e) Schematic drawing of circlets of fossil hooks (Modified from
Upeniece (1999).) found in juveniles of placoderm fish A. ornata (No 1e11), in juveniles
and adults of acanthodian L. gaujicus (No 12e23), and in/on crustacean Mysidacea (No
24); (f) Fossil hooks found in juvenile acanthodian body near the scapula (see also
Figure 1(e), No. 17), LDM 270/4a; (g) fossil hooks found near a clam shrimp. All speci-
mens derive from the Middle Devonian, Liepa (Lode) pit. Dark grey circles, squares e
possible endoparasites; Light grey circles, squares e possible ectoparasites.
100 Kenneth De Baets et al.

such as in Amphilinidea (Rohde, 2005). Derived Cestoda have a specialized


attachment device (the scolex) that has a highly variable morphology and
may have hooks: Diphyllidea typically possesses a scolex bearing a dorsal
and ventral set of apical hooks (Caira and Reyda, 2005), some Cyclophylli-
dea have a dome-shaped structure at the end of the scolex, the rostellum,
which may be armed with hooks arranged in one or more circles (Bush
et al., 2001), while others like Trypanorhyncha can have a scolex with
four retractable tentacles bearing hooks (Caira and Reyda, 2005). Most au-
thors agree that the fossil circlets of hooks probably belong to platyzoan hel-
minths (Upeniece, 2001, 2011; Poinar, 2003), although we cannot entirely
exclude the possibility that they belong to a now-extinct lineage of parasites.
A reinvestigation of these attachment structures with particular focus on
taphonomy, their composition (element analysis) and the three-dimensional
structure using computer tomography can be particularly useful for disen-
tangling phylogenetic affinity as it did for the elements of the now-extinct
conodonts (Purnell and Donoghue, 1997; 2005; Goudemand et al., 2011;
Murdock et al., 2013). Whatever the exact taxonomic affinity of the parasite
hook circlets, they remain the oldest direct evidence for the presence of
helminthegnathostome and helminthevertebrate association in the fossil
record (Boucot and Poinar, 2010).
Age: The finds of the Lode Formation were described initially as Early
Frasnian, Upper Devonian (Upeniece, 2001), although most authors now
assign this to the Upper Givetian, Middle Devonian (Mark-Kurik et al.,
1999; Jurina and Raskatova, 2012; Luksevics et al., 2012; Mark-Kurik and
P~oldvere, 2012; Luksevics et al., 2014). This corresponds with at least
381.9 Ma, the minimum age assigned to the GivetianeFrasnian boundary
(382.7 Ma  0.8 Myr: Becker et al., 2012). It is common practice in geolog-
ical and paleontological studies to use Ma for ‘Million years ago’, while XX
Myr is often used to refer to a duration of XX million years.
Luksevics et al. (2009) attributed various skeletal pathologies from Mid-
dle (Givetian) to Upper Devonian (Frasnian, Famennian) gnathostomes to
cestode and trematode infestations. However, the characteristics of these
(Luksevics et al., 2009) and other pathologies like skin lesions (Petit,
2010; Petit and Khalloufi, 2012) are insufficient to attribute them to partic-
ular group of parasites confidently, or even rule out potential other causes.
Interestingly, they also overlap temporally with the presence of blister pearls
in Devonian ammonoids (Rakoci nski, 2012), which might also have been
caused by parasitic flatworms, although no conclusive evidence for a parasitic
flatworm infestation was found (De Baets et al., 2011, 2013, 2015).
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 101

2.2 Silurian blister pearls and calcareous concretions in


bivalve shells
As (blister) pearls and volcano- to igloo-shaped concretions can be induced
by intermediate stages of parasitic flatworms (G€ otting, 1974, 1979; Lauckner,
1983; Campbell, 1985; Ituarte et al., 2001, 2005), their first occurrence in
the Silurian (Kríz, 1979; Liljedahl, 1985, 1994; De Baets et al., 2011) might
already indicate the presence of derived parasitic flatworms in the Silurian.
Pearls and blisters can, however, be caused by a variety of irritants, including
other parasites, shell burrowing organisms and inorganic particles (G€ otting,
1974; Lauckner, 1983). The earliest known fossil blister pearls from the Silu-
rian (Kríz, 1979; Liljedahl, 1985, 1994) and earliest known free pearls from
Triassic (Kutassy, 1937; Conway Morris, 1981; Combes, 2001; Geyer et al.,
2005; Rouse, 2005; Boucot and Poinar, 2010) are therefore not character-
istic for parasitism unless parasitic remains can be found inside of them
(De Baets et al., 2011). This is not straightforward since it has been demon-
strated that the remains of soft-bodied helminths can be destroyed during the
pearl formation process (Lauckner, 1983). However, other pathologies are
believed to be more characteristic for particular lineages of parasitic flat-
worms such as Gymnophallidae including shell pits (Ruiz and Lindberg,
1989; Ruiz, 1991; Huntley, 2007; Todd and Harper, 2011, Figure 1(e);
Huntley and Scarponi, 2012; Huntley et al., 2014; Huntley and Scarponi,
2015; Huntley and De Baets, 2015) and volcano- to igloo-shaped calcareous
concretions (Campbell, 1985; Ituarte et al., 2001, 2005; Figure 2(aec);
Huntley and De Baets, 2015). Ituarte et al. (2001, 2005) demonstrated a
link between igloo-shaped concretions and gymnophallid digenean flat-
worms, which these authors traced back to 6400 years in the Holocene. Su-
perficially, similar igloo-shaped concretions have, however, also been
reported from the Upper Silurian (Liljedahl, 1985, 1994; Figure 2(d)), but
the Paleozoic occurrence of this structure is not consistent with extant
host associations of Gymnophallidae, which typically have shorebirds (Char-
adriiformes) as final hosts (Ching, 1995), although some forms also infest
humans as final hosts (Lee and Chai, 2001). The earliest fossils that can be
confidently assigned to extant lineages of charadriiform birds are stem-group
representative of Alcidae from the Upper Eocene of North America (Mayr,
2011), although older charadriiform-like fossils have been reported from the
Lower Eocene of Denmark (Bertelli et al., 2010, 2013). Molecular clock es-
timates usually place the origin of shorebirds in the Cretaceous (Paton et al.,
2003; Baker et al., 2007), although this might be based on the incorrect
102 Kenneth De Baets et al.

(a) (b)

(c)

(d) (e)

Figure 2 Shell structures (igloo-shaped concretions, shell pits) which have been linked
with gymnophallid trematodes (Digenea) in extant and fossil bivalves. (a) Metacercaria
lodged in live position into an igloo-shaped calcareous covering of Gaimardia trapesina
(MLP 5659) from Beagle Channel in Ushuaia; note the noncalcified area around the
anterior end of the larva (photo: Cristian Ituarte; refigured from Ituarte et al. (2005).);
(b) Scanning electron micrograph of a left valve of Cyamiomactra sp. from a Holocene
sample of Río Varela (Tierra del Fuego) showing a single igloo-shaped covering just
below the anterodorsal margin (Photo: Cristia n Ituarte; refigured from Ituarte et al.
(2005).); (c) Upper view of an igloo-shaped covering in G. trapesina (MLP 5659) from
Beagle Channel in Ushuaia showing the non-calcified area in front of the igloo opening
(Photo: Cristian Ituarte; refigured from Ituarte et al. (2005).); (d) Igloo-shaped concretion
found close to posterior adductor muscle scar in the Silurian bivalve Nuculodonta got-
landica (SGU Type 1030) from the Halla Formation of Gotland (Modified from Liljedahl
(1994).); (e) Interior of right valve of Venericor clarendonensis (NHMUK PI TB 14236)
from the Eocene (subdivision B2 of the London Clay) with irregular shell deformations
and shell pits interpreted to have been produced in response to digenean trematode
infestation. Photo courtesy of Jon Todd; compare Todd and Harper (2011).
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 103

placement of fossil taxa (Dyke and Van Tuinen, 2004; Thomas et al., 2004;
Mayr, 2011) and/or other methodological artefacts (cf. Ksepka et al., 2014).
The appearance of shorebirds in the Cretaceous or Early Paleogene suggests
that the Paleozoic structures were probably caused by a different group of
parasites or even epizoa (Liljedahl, 1985, 1994) with similar behaviour,
although we cannot rule out the possibility that gymnophallids, or closely
related (now potentially extinct) taxa or their ancestors, had different life cy-
cles and host associations in the past. Shell pits have so far been traced back
only to the Eocene (Ruiz and Lindberg, 1989; Todd and Harper, 2011;
Huntley and De Baets, 2015), which is more or less consistent with the pres-
ence of gymnophallideshorebird associations (Figure 2).
Age: The Silurian occurrence of an igloo-shaped concretion was found
in the silicified M€ollboss 1 fauna from the Halla Formation (previously Halla
Beds) of Gotland (Liljedahl, 1985, 1994). Jeppsson et al. (2006) correlated
M€ ollboss 1 with the parvus graptolite biozone and the Ozarkodina bohemica
longa conodont subzone 2 as defined by Calner and Jeppsson (2003). The
top of the Pristiograptus dubius parvus e Gothographus nassa graptolite biozone
has been dated to 428.18  0.4796 Myr, yielded an minimum age of
427.7 Ma (Melchin et al., 2012). This is consistent with UePb age of
428.45  0.35 Myr obtained by Cramer et al. (2012) for the Gr€ otlingbo
bentonite at the nearby locality H€ orsne 3, which has been correlated with
M€ ollboss 1 (Jeppsson et al., 2006) and probably accounts for some of the sil-
ification of the strata there (Mikael Calner, personal communication 2014).

2.3 Permo-Carboniferous egg remains in shark coprolites


The earliest confident record of parasitic flatworms with complex parasite
life cycles are eggs attributed to Cestoda from the Rio do Rasto Formation
(Dentzien-Dias et al., 2013), which has been assigned to the Middle to Late
Permian (Holz et al., 2010). The eggs were obtained from a coprolite
(Figure 3(a)), which was isolated from its host and forms part of a set of
more than 800 coprolites of different shapes and sizes found in a geograph-
ically restricted area (Dentzien-Dias et al., 2012, 2013). It was identified as a
shark coprolite by its spiral structure and fossil content (Dentzien-Dias et al.,
2012, 2013). The eggs occur in a cluster (Figure 3(b)) and are ovoid, smooth
shelled and with a small operculum (polar swelling) suggesting that they are
nonerupted eggs (Figure 3(c) and (d)). Most eggs are filled with pyrite and
one egg is suggestive of containing a developing larva (Figure 3(c) and
(d)). The eggs vary little in size within the cluster, ranging from 145 to
155 mm in length and 88e100 mm in width. The morphological features
104 Kenneth De Baets et al.

(a) (b)

(c) (d)

Figure 3 Fossil evidence for the presence of derived parasitic flatworms (Cestoda) in
the Middle Permian (Modified from Dentzien-Dias et al. (2013).): (a) Picture of the spiral
heteropolar coprolite from the Rio do Rasto Formation, which has yielded the cestode
eggs, before destructive thin section analysis; (b) Thin section of the coprolite part con-
taining parasite eggs clustered in; (c) Cestode egg with a developing embryophore.
(d) Partial reconstruction of egg in (c) with interpretations of the observed structures.
Abbreviations: C ¼ capsule or shell; E ¼ embryophore (ochosphere); H ¼ putative
developing hooklets; I ¼ inner envelope; M ¼ oncospheral membrane; O ¼ outer
envelope; P ¼ putative polar thickening; Op ¼ operculum; S ¼ somatic cells.

of these eggs (operculum, egg shape and size: Figure 3(c) and (d)) as well as
their deposition together in an elongate arrangement (Figure 3(b)), which is
typical of modern tapeworm eggs deposited in mature segments or proglot-
tids, corroborates their cestode affinity.
Age: The coprolite derives from the upper member of the Rio do Rasto
Formation, which has also yielded a variety of vertebrate faunas (Dentzien-
Dias et al., 2012). The locality is located near Posto Queimado, where verte-
brate faunas indicate a Guadalupian (Late WordianeCapitanian) age (Cisneros
et al., 2012; Dias-Da-Silva, 2012). The coprolite should therefore be older
than the Guadalupian (Capitanian)eLopingian (Wuchiapingian) boundary
dated to at least 259.4 Ma (259.8  0.4 Myr: Henderson et al., 2012).
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 105

Older reports of potential cestode eggs (Zangerl and Case, 1976; Combes,
2001) within a Carboniferous coprolite are still controversial (Boucot, 1990;
Poinar, 2003; Dentzien-Dias et al., 2013). Strictly speaking, this coprolite
should be called a cololite as the fossilized faecal remains were found lodged
within its producer (Hunt et al., 2012). Boucot (1990) considered the report
as fairly speculative as the urea-rich environment of the dead shark would
probably lead to the rapid decomposition of such eggs, but the latest discovery
of cestode eggs in a Permian shark coprolite (Dentzien-Dias et al., 2013) illus-
trate that their interpretation as helminth eggs cannot be excluded. However,
further analyses of the morphology and the arrangement of these spherical
bodies are necessary to confirm their assignment to cestodes (Zangerl and
Case, 1976; Combes, 2001) or other parasitic helminths.
Age: The Cobelodus aculeatus specimen with the putative cestode eggs
derives from the Stark Shale, the core black shale member of the Dennis
Formation of the Missouri Series, near Fort Calhoun, Nebraska. It was
assigned to the Westphalian D by Zangerl and Case (1976) without any in-
formation corroborating this assignment. The Missouri series form part of
the local Missourian stage, which largely corresponds with the Kasimovian
(Falcon-Lang et al., 2011), which also includes the Stark Shale (Rosscoe,
2008; Rosscoe and Barrick, 2013). An age older than the Kasimoviane
Gzhelian boundary (303.6 Ma according to the 2012 Geological Timescale:
Davydov et al., 2012) can therefore be used as minimum constraint.
Pending reinvestigation of this fossil, it could extend range of cestode
eggs in shark coprolites by an additional 40 Myr.

2.4 Cretaceous egg remains in terrestrial archosaur


coprolites
The oldest fossil evidence for Trematoda is an egg which was recovered from
an Early Cretaceous isolated terrestrial vertebrate coprolite found near Bernis-
sart in Belgium (Poinar and Boucot, 2006). The producers of these coprolites
(and therefore also hosts of the parasites) are still debated (Chin et al., 1998;
Baele et al., 2012). Both Bertrand (1903) and Poinar and Boucot (2006) pre-
sented arguments that they could have been produced by theropod dinosaurs
(as opposed to crocodiles). However, theropod coprolites are rare (Hone and
Rauhut, 2010), particularly when compared with coprolites of aquatic verte-
brates (Chin, 2002) such as crocodylians in contemporary deposits (Hunt
et al., 2012). Their seemingly precise assignment to ‘Megalosaurus’ dunkeri is
therefore questionable, as it is based on the co-occurence of a single metarsal
in the same deposits. This metatarsal was originally assigned to M. dunkeri
106 Kenneth De Baets et al.

(now Altispinax dunkeri), but its morphology is only sufficient to assign it to


theropods at best (Pascal Godefroid, personal communication 2014). Never-
theless, it remains the oldest evidence for trematodes in terrestrial predatory
archosaurs (Poinar and Boucot, 2006).
Age: The fossil-bearing strata are now more precisely dated to be of Late
Barremian to Early Aptian age (Yans et al., 2005; Schnyder et al., 2009; Yans
et al., 2012), corresponding with the upper part of magnetochron M1n, M0r
and the basal part of M0n. This yields an approximate minimum age for
these strata of 125.93 Ma, the age attributed to the top of magnetochron
M0r by Ogg (2012), or a more conservative age of 125.7 Ma, the age
assigned to the base of the Tethyan Deshayesites deshayesi ammonoid biozone
(Schnyder et al., 2009) by Ogg et al. (2012).

2.5 Eocene shell pits in intermediate bivalve hosts


The oldest evidence for the presence of Gymnophallidae might lie in the
Eocene in the form of characteristic shell pits found in their intermediate
bivalve hosts (Ruiz and Lindberg, 1989, Figure 2(e), Todd and Harper,
2011). These characteristic pits have been reported throughout the Cenozoic
from the Eocene to the Holocene (Johannessen, 1973; Ruiz and Lindberg,
1989; Ruiz, 1991; Huntley, 2007; Todd and Harper, 2011; Huntley and
Scarponi, 2012, 2015; Huntley et al., 2014; Huntley and De Baets, 2015).
These pits have been commonly linked with Gymnophallidae (Ruiz and
Lindberg, 1989; Todd and Harper, 2011), although superficially similar
structures might also be caused by other digenetic trematodes such as Lep-
ocreadiidae (Ituarte et al., 2001; see review by Huntley and De Baets, 2015).
Age: The oldest precisely dated shells (Venericor clarendonensis) with pits
derive from subdivision B2 of the London Clay, Eocene (Todd and Harper,
2011). Berggren and Aubry (1996) assigned this unit to upper calcareous
nannofossil zone NP11, which corresponds with a minimum age of 53.9 Ma
according to the 2012 Geological Timescale (Vandenberghe et al., 2012).
Further support could come from the distribution of volcano-shaped
(Campbell, 1985) to igloo-shaped calcareous concretions (Ituarte et al.,
2001, 2005), which have been traced back to at least 6240  70 years BP
in the Holocene (Ituarte et al., 2005). Despite a certain degree of variability
in these structures in extant bivalves, they are believed to be characteristic for
gymnophallid trematodes (Campbell, 1985; Ituarte et al., 2001, 2005).
Studies on pathology have focused on invertebrate intermediate hosts
(Ruiz and Lindberg, 1989; Ituarte et al., 2001, 2005; Huntley, 2007;
Huntley and Scarponi, 2012; Huntley et al., 2014; Huntley and De Baets,
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 107

2015), although some pathologies in vertebrate intermediate hosts have also


been linked with parasitic flatworms. The best examples are probably the
teratological limb malformations in North American amphibians, which
have been linked with the trematode Ribeiroia on several occasions (Johnson
et al., 2001, 2002; Stopper et al., 2002; Johnson and Sutherland, 2003; John-
son and Chase, 2004; Koprivnikar et al., 2012) and could potentially be found
in the fossil record (cf. Fr€
obisch et al., 2014). Nevertheless, limb malforma-
tions e including supernumerary limbs and bone bridges e can have various
other causes, meaning their interpretation is not always straightforward
(Blaustein and Johnson, 2003; Lunde and Johnson, 2012).

2.6 Eggs remains in a Pleistocene mammal coprolite


The oldest Quaternary flatworm evidence is derived from an isolated Mid-
dle Pleistocene mammal coprolite ( Jouy-Avantin et al., 1999), which these
authors attributed to Ursidae based on its morphology and associated fossil
finds. The morphology of the eggs (asymmetrical shell, the brown colour
and the presence of an operculum) is characteristic of dicrocoelid flatworms,
which makes this the oldest direct evidence for the presence of dicrocoelid
flatworms (Digenea: Dicrocoelidae). They could not be assigned to a partic-
ular genus, although their dimensions are reminiscent of Dicrocoelium and
Eurythrema based on egg measurements.
Age: The coprolite (H13 HEN5 1526) derives from an archeological
layer at the Caune de l’Arago cave (Tautavel, Pyrénées-Orientales, France)
and could be dated to a minimum age of 550,000 years BP (Jouy-Avantin
et al., 1999) during a cold and dry climatic period (Lumley et al., 1984).

2.7 Holocene evidence for parasitic flatworms from ancient


remains
Other Quaternary parasitic flatworm fossils and subfossils derive from the
Holocene, mostly from archeological sites (see Gonçalves et al., 2003 for a
review, Searcey et al., 2013; Ara ujo et al., 2014; Beltrame et al., 2014),
with possible ages up 6368 years BP for Trematoda and ages up to
10,000 years BP for Cestoda. They can provide upper constraints for the
earliest appearance of various taxa of Cestoda and Trematoda, including
genera and species. The age assignment used in archeological publications
can be a bit confusing. Before Present (BP) stands for a timescale, which
starts at the 1st of January 1950 reflecting the fact that radiocarbon dating
became practicable around that time and also antedates large-scale nuclear
weapons testing altering the global ratio of carbon isotopes (Taylor,
108 Kenneth De Baets et al.

1985). Archeological publications often use BC (Before Christ) and AD


(Anno Domini). Hundred years BP is 100 years before 1950 (i.e. the year
AD 1850). At ages older than about 0.5 Ma, the difference between BP
and AD becomes negligible. We herein use the dates mentioned in the orig-
inal publications to avoid confusion. Note that age assignment might change
or differ according to the dating methods used (Iles, 1980). We recommend
using the most conservative age estimates using reliable methods.
Remains of the Cestoda Diphyllobothrium pacificum can be traced back to
about 10,000e4000 BP according to Reinhard (1992), although the exact
evidence for such an age were not discussed in this paper. The record of ano-
plocephalid cestodes can be traced back to at least 8920  200 years BP
based on eggs founds in coprolites attributed to humans (Fugassa et al.,
2010). Anoplocephalid remains which could be more specifically deter-
mined as Monoecocestus can be found in rodent coprolites dated as old as
6700  70 years BP (Sardella et al., 2010). The earliest reports of Hymenole-
pis were dated approximately from 4000 to 2000 years BC (Gonçalves et al.,
2003) and Taenia eggs from an Egyptian mummy attributed to about
3200 years BC (Reyman et al., 1977). According to Gonçalves et al.
(2003), trematodes (Fasciola as well as Opisthorchioidea) can be traced
back to at least 5400  40 to 5230  40 years BP (Roever-Bonnet et al.,
1979), Dicrocoelium can be traced back to 3384e3370 BC (Dommelier
Espejo, 2001), Schistosoma can be traced back to 3200 years BC based on
the discovery of Schistosoma haematobium antigen in the shin tissue of an
Egyptian predynasty mummy (Deelder et al., 1990) and Schistosoma ova in
another contemporary mummy (Reyman et al., 1977), and as Clonorchis
sinensis could be traced back to a mummy from Chu Dynasty (475e
221 years BC) with an age of at least 2171 years BP (Wen-yuan et al.,
1984). Eggs of S. haematobium were one of the earliest to be discovered in
Egyptian mummies.
Recently, ancient DNA of echinostomatid trematodes was extracted
from coprolites of the extinct ratite bird Megalapteryx from New Zealand
(Wood et al., 2013), which might range from about 6368 years BP to the
694  30 years BP, coincident with the time of their extinction (Wood
et al., 2012).

2.8 Free-living flatworms


The body fossil record of free-living flatworms is also of little help as it is
poorer or even more patchy than the fossil record of parasitic flatworms
(Poinar, 2003). The oldest free-living flatworm body fossils derive from
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 109

Eocene Baltic amber (Poinar, 2003) and calcareous nodules of Miocene age
(Pierce, 1960; Poinar, 2003), and those of flatworm egg capsules from Qua-
ternary lake sediments (Frey, 1964; Harmsworth, 1968; Gray, 1988). These
fossils considerably post-dated divergence time estimates for free-living flat-
worms derived from molecular clock studies. Older molecular clock studies
place the origin of (free-living) Platyhelminthes deep in the Precambrian
(Hausdorf, 2000; Otsuka and Sugaya, 2003), but more recent relaxed molec-
ular clock studies place their origins between the latest Precambrian (Edia-
caran) and the early Cambrian (Douzery et al., 2004; Peterson et al., 2004,
2008; Erwin et al., 2011). Some Ediacaran fossils have occasionally been
related to free-living flatworms (Allison, 1975; Palij et al., 1979; Fedonkin,
1985), although they cannot be confidently assigned to the phylum (Conway
Morris, 1981; Labandeira, 2002; Erwin et al., 2011). Even Dickinsonia was
assigned to flatworms at one point (Palij et al., 1979; Conway Morris,
1981). The taxonomic position of Dickinsonia has been heavily debated
(Retallack, 2007; Brasier and Antcliffe, 2008) and this taxon is often inter-
preted as one of the earliest divergent metazoans (Xingliang and Reitner,
2006; Sperling and Vinther, 2010). Various Permian to Triassic trace fossils
have also been attributed to turbellarians including Polycladida and Tricladida
(Alessandrello et al., 1988; Knaust, 2010). The assignment of trace fossils to
this phylum is also problematic as various worm-like groups with similar ecol-
ogy and mode of locomotion could also have produced these traces (Sei-
lacher, 2007). Curvolithus has also been attributed to flatworms (Seilacher,
2007), but could also have been produced by other taxa with similar behav-
iour (Buatois et al., 1998). Further studies of traces produced by extant forms
as well as fossil traces associated with body fossils are therefore important to
confidently assign them to the phylum (Collins et al., 2000; Knaust, 2010).
The oldest parasitic flatworm fossils are therefore not only important for
putting constraints on free-living flatworms, but also on the presence of
Platyzoa in the fossil record, a group currently containing both parasites
(Platyhelminthes, Acanthocephala) and free-living taxa (Wey-Fabrizius
et al., 2013). The assignment of Cambrian fossils (cambroclaves) to the Acan-
thocephala as suggested by some authors (Qian and Yin, 1984), which has also
been followed in some recent classifications (Amin, 2013), is highly question-
able and widely rejected (Conway Morris et al., 1997; Elicki and Wotte,
2003; Kouchinsky et al., 2012). These problematic Cambrian organisms
can be classified as Lophotrochozoa at best (Kouchinsky et al., 2012);
Compare Conway Morris and Crompton (1982) and Near (2002) for further
speculations and hypotheses on the origin and evolution of parasitism towards
110 Kenneth De Baets et al.

the rise of the Acanthocephala. Furthermore, the oldest accepted record of


Rotifera, the free-living relatives of Acanthocephala, derives from Dominican
amber deposits (Waggoner and Poinar, 1993), which is now more confi-
dently dated to the Miocene (Iturralde-Vinent and MacPhee, 1996;
Iturralde-Vinent, 2001). The oldest confidently assigned Acanthocephala re-
mains have been reported from archeological sites (Gonçalves et al., 2003);
some dating back to 9500 years BC according to Fry and Hall (1969). Consid-
ering that multiple authors (Upeniece, 2001, 2011; Littlewood and Donovan,
2003) have suggested that the Middle Devonian hook circlets might belong to
a platyzoan helminth, a putative reinvestigation of these fossils could be used
to constrain the evolutionary history of this entire group. Many studies on
extant flatworms have focused on the hook elements of particular groups
(Vignon and Sasal, 2010; Vignon, 2011) or taxa (Shinn et al., 2003). Informa-
tion and illustrations of these helminth structures can also be found in
comprehensive systematic treatments: Yamaguti (1959), Schmidt (1986)
and Khalil et al. (1994) for Cestoda; Yamaguti (1963a) for Monogenea and
Yamaguti (1963b) and Golvan (1969) for Acanthocephala.

3. INTERPOLATING OR EXTRAPOLATING EXTANT


PARASITEeHOST RELATIONSHIPS AND THE
ASSUMPTION OF PARASITEeHOST COEVOLUTION
Analysis of the range of current parasiteehost associations has often
been used to infer the evolutionary origin of parasitic organisms (Littlewood
and Donovan, 2003). With the exception of highly derived taxa, parasitic
flatworms do not parasitize hagfishes or lampreys (Littlewood, 2006).
This may suggest that parasitic flatworms evolved in basal gnathostomes
(Littlewood, 2006), which would lie somewhere around the Cambriane
Ordovician based on the host fossil record (Friedman and Sallan, 2012;
Donoghue and Keating, 2014). This considerably predates the oldest gener-
ally accepted fossil evidence for parasitic flatworms, but the fossil record
indicates that other groups like pentastomids, which parasitize vertebrates
today and might have done in the past, were already around at this time
(Walossek and M€ uller, 1994; Walossek et al., 1994; Maas and Waloszek,
2001; Waloszek et al., 2005; Sanders and Lee, 2010; Castellani et al.,
2011). If we map the fossils of parasitic flatworms on their host phylogenies
(Figure 4), they are at least consistent with extant parasitic flatwormehost
associations, which is not the case for all parasites (Figure 5). Pentastomids
mainly parasitize terrestrial vertebrates today (Christoffersen and De Assis,
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 111

Figure 4 Fossil finds of putative flatworm fossils discussed in the text mapped on their
host phylogeny, which was modified from Donoghue and Smith (2003) and Rowe
(2004) taking into account new phylogenetic hypotheses summarized in Donoghue
and Keating (2014).

2013), but in the CambrianeOrdovician (Walossek and M€ uller, 1994;


Walossek et al., 1994; Waloszek et al., 2005; Castellani et al., 2011), they
are found in marine deposits and there were no terrestrial vertebrates to serve
as hosts at this time. Their morphology indicates a parasitic lifestyle, but the

Figure 5 Comparisons of the consistency between extant and fossil host ranges of
parasitic flatworms and pentastomids. Host phylogeny was modified from
Goudemand et al. (2011). Note that the host gap would be considerably greater if
the position of conodonts would be more basal as postulated by some authors (Blieck
et al., 2010; Turner et al., 2010). Recent discoveries of Silurian pentastomids associated
with ostracods even further extend the host gap between fossil and extant pentasto-
mids (Siveter et al., 2015).
112 Kenneth De Baets et al.

exact host of these marine forms is still unknown as they were not found
directly associated with their hosts. Some have suggested their hosts might
have been conodonts (Walossek and M€ uller, 1994), of which tooth-like
remains have been commonly found in these deposits. Interestingly, the
CambrianeOrdovician fossil pentastomids resemble larvae of modern forms
that can infest fish, making it conceivable that these small pentastomids
represent adults that spent their entire life cycle on small fish-like vertebrates
(Sanders and Lee, 2010). The range of the host gap between Cambrian and
extant pentastomids (Figure 5) might therefore depend on the systematic
position of conodonts, which is still debated. Most authors agree that con-
odonts are chordates, probably either stem- or crown-vertebrates (Blieck
et al., 2010; Turner et al., 2010; Goudemand et al., 2011; Murdock et al.,
2013; Donoghue and Keating, 2014). Interestingly, putative pentastomid
remains were recently also reported from ostracods within the Silurian
Herefordshire Lagerst€atte (Siveter et al., 2015), which further increases the
host gap between extant and fossil pentastomids. This might suggest that in-
vertebrates might have been the initial hosts in the marine realm, if ostracods
are the final hosts as suggested by these authors and if no host switching
occurred between the Cambrian and the Silurian. However, pentasto-
midehost associations from the Cambrian and Ordovician lagerst€atten
would be necessary to further test this hypothesis.
Thus, it is not always possible to precisely constrain parasiteehost asso-
ciations in the fossil record. This is not only the case for isolated remains
of parasites not directly attached to their hosts (Castellani et al., 2011), but
also for parasite remains found in isolated coprolites not confidently assign-
able to precise host taxa (Jouy-Avantin et al., 1999; Poinar and Boucot,
2006; Dentzien-Dias et al., 2013).
Fossils of potential hosts might be common or present in the same layers,
although without direct evidence for a parasitic relationship (attached or
found within well-preserved body fossils of their hosts), appointing a poten-
tial host remains within the realm of speculation. In the case of intermediate
hosts, where the fossil evidence is often pathologies or traces, it is hard to be
certain of the identity of the culprits since various organisms with similar
behaviour can produce similar traces. Furthermore, it might also be hard
to identify the final host without finding remains of the putative parasite
associated with them. Predatoreprey relationships might provide a clue
(Ruiz and Lindberg, 1989; De Baets et al., 2011), but direct evidence for
predation from one taxon on another is rare in the fossil record too (Brett
and Walker, 2002; Walker and Brett, 2002). Nevertheless, age in itself might
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 113

be sufficient to rule out certain hosts if they had not yet evolved, they must
have had different intermediate and/or final hosts at that time. Although the
hosts of CambrianeOrdovician pentastomids remain speculative, hosts
assignable to taxa they currently parasitize were not around yet, so they
must have had additional parasiteehost associations in the past which are
now extinct.

4. MOLECULAR CLOCK STUDIES


Molecular clock (timetree) methods for calibrating phylogenies
have the potential to be useful to discriminate evolutionary scenarios in
parasite evolution or hosteparasite associations (Hypsa, 2006). However,
molecular clocks still need to be calibrated to obtain absolute age estimates,
which is not that straightforward among parasite groups with a patchy fossil
record. Thus, a cophylogenetic approach has been used (Hafner et al.,
1994; Page et al., 1998; Light and Hafner, 2007), which may in turn pro-
vide a robust evolutionary timescale for apparent cospeciating symbiotic
species when the timescale of the host lineage is (comparably) well con-
strained (Moran et al., 1993, 1995). Due to the lack of well-preserved spec-
imens in parasitic flatworms, most molecular clock studies have relied on
the host fossil record to inform divergence estimates (Verneau et al.,
2002, 2009a,b; Olson et al., 2010; Badets et al., 2011; Héritier et al.,
2015). More rarely biogeography (focusing on vicariance events) has
been invoked to constrain molecular clock estimates of parasitic flatworms
(Zietara and Lumme, 2002; Waltari et al., 2007; Badets et al., 2011;
Martínez-Aquino et al., 2014). Only one study (Perkins, 2010) used the
parasitic flatworm fossil record as a calibration, although it relied on the
Upper Devonian and Carboniferous putative flatworm fossils, whose taxo-
nomic assignment to Monogenea and Cestoda, respectively, is still debated
(as discussed above). Other studies have relied on molecular substitution
rates derived from other studies (Despres et al., 1992; Zietara et al.,
2002; Huyse and Volckaert, 2005), which is even more problematic
(Papadopoulou et al., 2010; Hipsley and M€ uller, 2014).
The best practice for using fossils for molecular clock calibration has been
discussed and reviewed extensively (Donoghue and Benton, 2007; Benton
et al., 2009; Parham et al., 2012). Most authors agree that the fossil speci-
mens can only directly provide a well-justified minimum constraint for
the origin of some particular lineages. According to Parham et al. (2012),
114 Kenneth De Baets et al.

fossil calibrations are well justified if the following criteria are fulfilled:
(1) listing of museum numbers of specimen(s) that demonstrate all the rele-
vant characters and provenance data, (2) availability of apomorphy-based
diagnoses of the specimen(s) or an explicit, up-to-data, phylogenetic analysis
that includes the specimen(s), (3) explicit statements on the reconciliation of
morphological and molecular data sets, (4) specification of the locality and
stratigraphic level (to the best current knowledge) from which the calibrat-
ing fossils are derived, (5) reference to a published radioisotopic age and/or
numeric timescale with details on its selection. Divergence time estimation is
not possible with minimum constraints alone, as the substitution rate is var-
iable and unknown; therefore, at least one point calibration or maximum
constraint is required to calculate the substitution rate and absolute diver-
gence times (Warnock et al., 2012).
It would therefore be more appropriate to use the oldest estimate (95%
confidence maximum) for the origin of the total group from a robust
molecular clock analysis as the maximum, and the oldest fossil assignable
with confidence to the crown group (as the minimum), to constrain the
evolutionary history of the host (and the parasite). For example, if the origin
of parasitic flatworms occurred during the early evolution of gnathostomes
as suggested by Littlewood (2006), this would mean they originated bet-
ween the Cambrian (643 Ma: oldest confidence intervals of Erwin et al.,
2011 for this node) and the earliest well-dated fossils that can be confidently
assigned to crown-group gnathostomes deriving from Ordovician
(w421.8 Ma: Benton et al., 2009; Donoghue and Keating, 2014). Alterna-
tively, the entire 95% confidence posterior interval (from robust molecular
clock studies) for diverging host clades could be used as priors on the clade
ages in the parasites.
A more conservative and less circular approach would be to use the lat-
est robust relaxed molecular clock estimates for the origin of their free-
living ancestors (744 Ma: oldest confidence of Erwin et al., 2011 for this
node) and the earliest certain appearance of parasitic flatworms in the fossil
record, which would be only Permian (>259.4 Ma as discussed above).
This would yield quite large confidence intervals (cf. Warnock et al.,
2012 for Parasitiformes), but such estimates would be more honest and
accurate (closer to reality) than seemingly precise estimates which are argu-
ably inaccurate, since the origin of parasitism falls outside the calibration
interval a priori (Warnock et al., 2011). Furthermore, reinvestigation of
putative flatworm trace or body fossils from Upper Silurian (>427.7 Ma)
or Middle Devonian (>381.9 Ma) deposits might further narrow this
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 115

time interval in the future. The fossil record also provides constraints on the
origin of Cestoda in shark hosts (Dentzien-Dias et al., 2013) and Trema-
toda in archosaur hosts (Poinar and Boucot, 2006). Furthermore, the fossil
record provides upper constraints on the origin of Gymnophallidae in the
form of characteristic pathologies, dicrocoelid flatworms derived from eggs
in coprolites as well as several genera and even species from archeological
sites (e.g. Dicrocoelium, Diphyllobothrium, Fasciola, Monoecocestus, Schistosoma,
Taenia).
The oldest fossil evidence for schistosomes are antigens deriving from a
3200 BC Egyptian mummy, but the origin of this group is believed to be
considerably older based on the evolutionary history of their intermediate
or final hosts (Lawton et al., 2011). Davis (1993) suggested that the genus
Schistosoma arose before the breakup of the supercontinent Gondwana
over 150 Ma based on the distribution of their snail hosts and that ancestors
of Asian schistosomes were carried to Asia via India after it separated from
Africa. More recent studies (Snyder and Loker, 2000) have suggested a
younger, ancestral Asian origin somewhere in the Miocene, which might
indicate that schistosomes only colonized Africa around 15e20 Ma (Lawton
et al., 2011). Performing a robust molecular clock analysis using dates of fos-
sil Schistosoma and their hosts might be a more formal way to test these
hypotheses. In some cases, additional historical dates might become avail-
able to constrain certain nodes such as the possible slave transport of Schis-
tosoma mansoni to South America (Lockyer et al., 2003b), which is so far not
contradicted by finds of older remains of S. mansoni in archeological sites of
South America (Gonçalves et al., 2003). Direct dating of samples yielding
ancient DNA (Wood et al., 2013) might also provide additional constraints
in such studies. However, a recent study by Mello et al. (2014) has demon-
strated that the assignment of calibration information to deeper phyloge-
netic nodes is more effective in obtaining more precise and accurate
divergence time estimates compared to analyses involving calibration at the
shallowest node.
Most authors agree that multiple, well-justified calibrations are the best
approach to obtain the most robust and accurate molecular clock estimates
(Warnock et al., 2011; Parham et al., 2012). Note that careful a priori selec-
tion of suitable calibration points cannot be replaced by using as a posteriori
cross-validation procedures (Near et al., 2005; And ujar et al., 2014) as these
only verify consistency (Clarke et al., 2011). In some cases, multiple inaccu-
rate calibrations might be consistent, which can result in erroneous rejection
of more reasonable calibrations. Furthermore, consistent calibrations may be
116 Kenneth De Baets et al.

redundant by definition, since they fail to correct for changes in rate varia-
tion (Clarke et al., 2011; Warnock et al., 2015). Furthermore, calibration
should be implemented in the most conservative way, which might result
in less precise, but ultimately more accurate divergence estimates (Warnock
et al., 2012).
In the absence of a suitable fossil record, one could resort to the use of
biogeographic events or calibrations related only to the host fossil record.
Nevertheless, both methods also have their problems and add an additional
component of circularity to calibration procedures depending on the
hypotheses being tested (Hipsley and M€ uller, 2014). Biogeographic calibra-
tions as they are currently implemented are problematic (Goswami and
Upchurch, 2010; Kodandaramaiah, 2011; De Baets and Donoghue, 2012;
Hipsley and M€ uller, 2014), and they should be implemented more con-
servatively. It should be established when a certain barrier, causal to a given
speciation event, actually occurred (De Baets and Donoghue, 2012;
Warnock, 2014).
Most importantly, there remains an assumption that biogeographic distri-
butions have not changed significantly in geological time, making it harder to
establish whether biogeographic barriers were coincident with speciation
events, and introduce an aspects of circularity (Crisp et al., 2011). Having
taxon-area relationships consistent or inconsistent with biogeographic
events, does not necessarily mean that these clades diversified at the same
time as these events, because older events might have led to similar distribu-
tions (pseudo-congruence) or younger events might have altered their dis-
tributions (pseudo-incongruence; see Donoghue and Moore, 2003). Even
for some of the classical examples of groups with current distributions
congruent with vicariance, such as onychophorans and cichlids, studies
have demonstrated that divergence might predate (Murienne et al., 2014)
or postdate (Friedman et al., 2013) the continental break-up of superconti-
nents, respectively.
Using the fossil record of hosts also introduces an aspect of circularity in
addition to other considerations related with fossil calibrations (Donoghue
and Benton, 2007; Parham et al., 2012) as discussed above. It assumes that
the current parasiteehost associations did not markedly change through
geological time, which is not necessarily true, particularly in groups which
are estimated to range several hundred million years into the past (e.g. the
pentastomid example we discussed above). Using hosts also leads to circular
reasoning when employing them to investigate hypotheses of parasiteehost
coevolution. In highly host-specific lineages with simple life cycles like the
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 117

Polystomatidae (Figure 6), it might work well, but for groups with common
host switches and/or complex life cycles this approach might be less suitable.
The most conservative way to implement them would be to use the oldest
reliable estimate for the origin of this group as a maximum and the oldest
from well-attributable fossil to this lineage as a minimum. Polystomatid flat-
worms are one of the most host-specific groups of parasitic flatworms and
their direct life cycle that involves a short free-living aquatic larval stage
(which means they are probably only passively disseminated by their hosts),
have made them an ideal model to test the use of constraints from biogeog-
raphy and the host fossil record (Bentz et al., 2001, 2006; Verneau et al.,
2002, 2009a,b; Badets et al., 2011). Verneau et al. (2002) used 425 Ma to
calibrate the split between Actinopterygii and Sarcopterygii (Figure 6),
although this event must have happened at the latest by about 419 Ma
(Zhu et al., 2009). However, it would be more conservative to use the oldest
robust estimate from relaxed molecular clock studies for the separation of
actinopterygian from sarcopterygians and the earliest fossil confidently
assigned to either tetrapods or lungfishes to constrain this node (Badets
et al., 2011). The oldest stem-group lungfish is generally considered to be
Diabolepis (Friedman, 2007; Qiao and Zhu, 2009), while one of the oldest
ingroup lungfishes might be Westollrhynchus (Qiao and Zhu, 2009). Badets
et al. (2011) also suggested that some dates might be consistent with the
break-up of the supercontinent Gondwana, although this needs to be
further tested with additional sampling.
Taxon sampling can play a large role in tree reconstruction and interpre-
tation with respect to biogeography (Trewick and Gibb, 2010) or host
switching (Hafner and Page, 1995). In the case of lineages or parasitee
host associations, which have been around for many hundreds of millions
of years, host range changes and extinction might contribute significantly
to missing taxa, making it hard to infer past biogeographical distribution
or parasiteehost associations, from extant data alone. There is at least
some evidence that extinction might also have played a role in parasitic flat-
worms and other helminths over longer timescales as several parasiteehost
associations documented in the (sub)fossil record are now evidently extinct
(Upeniece, 2001, 2011; Poinar and Boucot, 2006; Wood et al., 2013).
Furthermore, molecular studies with greater taxonomic coverage have
particularly focused on biomedically or economically important taxa such
as Schistosomatidae (Lockyer et al., 2003b; Orélis-Ribeiro et al., 2014) or
particular lineages with a high host specificity such as Polystomatidae (Bentz
et al., 2001, 2006; Badets et al., 2011, 2013; Héritier et al., 2015). To better
118 Kenneth De Baets et al.

Figure 6 Ultrametric tree of neobatrachian polystomes inferred from MULTIDIVTIME


(Modified after Verneau et al. (2009b).). Calibration points (black rectangles, nodes
1e3) were deduced from historical biogeographical scenarios suggested by Bentz
et al. (2001, 2006) and Badets et al. (2011). The divergence of the lineage associating
Metapolystoma, Eurasian and African Polystoma from their closest South and North
American relatives (nodes 1 and 2) was constrained between 65 and 56 Myr, reflecting
vertebrate exchanges between the two Americas in the Paleocene (Gayet et al., 1992)
and possible dispersal to Eurasia via Beringia. The divergence between the European
Polystoma species (i.e. Polystoma gallieni) and the lineage grouping Metapolystoma
and African Polystoma was constrained between 25 and 5 Myr, reflecting the
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 119

understand the evolutionary history using molecular methods, it is essential


to sample as many distinct lineages of parasites as possible as well as their free-
living relatives, focusing particularly on evolutionary important taxa which
have putative fossil records (e.g. Gymnophallidae, basal Monogenea, Ces-
toda) or reliable geological constraints derived from biogeography or the
evolutionary history of their hosts.
When no appropriate constraints are available, relative rates of uncali-
brated molecular clocks can be used to test the support or reject the temporal
congruence of parallel distributions or parasite-host evolution (Loader et al.,
2007; Hibbett and Matheny, 2009; Loss-Oliveira et al., 2012). Nevertheless,
it should be kept in mind when interpreting the results that the rate of
molecular evolution might be significantly different between parasites and
hosts (Page et al., 1998) or within and between lineages of parasites and/
or hosts (Thomas et al., 2006, 2010; Bromham, 2009), which can bias the
results of such studies (Hipsley and M€ uller, 2014). Furthermore, novel prob-
abilistic approaches make it possible to incorporate in biogeographic infer-
ence, estimates of the divergence time of lineages as well as external
sources of evidence such as climate, geography, their fossil record or ecolog-
ical tolerance (Sanmartín, 2012).

5. CONCLUSIONS AND FUTURE PROSPECTS


The earliest fossil evidence for the presence of helminths falls in the
Middle Devonian in the form of hooks, some of which are most reminis-
cent of extant Monogenea, although some could also belong to Acantho-
cephala or more derived flatworms (Cestoda). The oldest secure record of
parasitic flatworms with complex parasite life cycles lies in the Permian
which can be confidently assigned to cestodes, although the presence of

=
hypothesized ages of dispersal routes between Eurasia and Africa (Rage, 1988; Bentz
et al., 2001). Finally, the root prior was set at 160 Ma (sd  5 Myr), corresponding to an
initial divergence separating Asian and Australian polystomes from all other neobatra-
chian polystomes (Badets et al., 2011), hypothetically corresponding to a separation of
the western and eastern components of Gondwanaland. Divergence time estimates
(see Verneau et al., 2009b) are reported for two nodes that are relevant for understand-
ing the origin of the new Malagasy genus, i.e. Madapolystoma (see Du Preez et al.,
2010). According to Verneau et al. (2009b), Madapolystoma would have diverged
from Eupolystoma about 116 Ma (node A) and the first crown divergence in Madapo-
lystoma (node B) would have occurred about 63 Ma.
120 Kenneth De Baets et al.

igloo-shaped concretions reminiscent of those caused by gymnophallid


trematodes in extant bivalves might already indicate the presence of
derived parasitic flatworms with complex parasite life cycles in the Late
Silurian (>428 Ma). These Silurian occurrences are, however, not consis-
tent with evolutionary history of current gymnophallid hosts (shorebirds),
which are believed to have appeared somewhere between the Cretaceous
and Eocene. Characteristic pits in bivalves shells indicative for the presence
of digenetic trematodes (Gymnophallidae) appear already in the Eocene
(Ruiz and Lindberg, 1989; Todd and Harper, 2011; Huntley and De Baets,
2015), which is more or less consistent with the presence of their final host
in the fossil record. The first evidence for terrestrial parasitic flatworms and
trematodes was found in the form of eggs within a Lower Cretaceous
coprolite (Poinar and Boucot, 2006), which can be confidently attributed
to archosaurs (potentially theropod dinosaurs or crocodylians). The earliest
evidence for dicroelid trematodes (Jouy-Avantin et al., 1999) falls at about
0.55 Ma in the Middle Pleistocene. Various extant genera and species have
been described from younger archeological sites (Gonçalves et al., 2003;
Araujo et al., 2014). Several putative flatworm fossils need additional study
to confidently assign them to a certain lineage of flatworms including pla-
tyzoan helminth hooks in Middle Devonian gnathostomes, putative
cestode eggs in a Carboniferous shark coprolite and eggs in a Cretaceous
archosaur coprolite. Remarkably, the fossil record of parasitic flatworms
in considerably better than that of free-living flatworms (Poinar, 2003)
and Platyzoa in general (Conway Morris and Crompton, 1982; Wey-
Fabrizius et al., 2013) and it could therefore be used to constrain the evolu-
tionary origin of flatworms and other Platyzoa. Only rarely have studies
been performed to assess the evolution of these structures over larger scales;
e.g. see Malmberg (1990) for Monogenea, which is a rather controversial
study for different reasons (Gusev, 1992). A comparative analysis of hook
elements of acanthocephalans and parasitic flatworms (Monogenea, Ces-
toda) in a new molecular phylogenetic framework would therefore be in
order to more confidently assign the fossil hook circlets to a certain clade
or phylum. Furthermore, the study of eggs as well as hook circlets, which
is now largely done with destructive methods and in two dimensions,
would benefit from CT-scanning technologies to characterize their 3D-
morphology and structure in a nondestructive way and potentially reveal
additional details or fossils which otherwise might be destroyed by the sam-
ple preparation process (e.g. thin-sectioning, chemical sample preparation,
resedimentation procedures).
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 121

Fossil evidence can only provide minimum time constraints and is not avail-
able for all lineages of parasitic flatworms. Interpolations based on parasiteehost
associations or biogeographic events can potentially be used to supplement fos-
sil constraints. However, circularity in testing hypotheses should be avoided
and caution should be taken when multiple host-switching events are sus-
pected. Before using such calibrations it should be at least verified if this hypoth-
esis is robust to a wider sampling of extant and extinct taxa as well as the
evolutionary history of their hosts, where the sampling and fossil record might
be comparatively better. We therefore advise implementing such calibrations in
the most conservative way. For calibrations based on the evolutionary history of
parasites or their hosts, this would correspond with using the oldest estimate
based on relaxed molecular clock estimates as maximum and the oldest well-
attributable fossils of the parasite or its host as a minimum; for calibrations based
on biogeographic events, this would correspond to the using the oldest age of
the oldest geological event that could have influenced the distribution of the
parasites and their hosts as a maximum and the youngest age of the geological
events that could have influenced their distribution as a minimum.
Several recent discoveries indicate that exceptionally preserved gnathos-
tomes or their coprolites might yield additional finds of parasitic flatworm
fossils, particularly their attachment organs or their eggs (Littlewood and
Donovan, 2003; Poinar and Boucot, 2006; Dentzien-Dias et al., 2013),
which can be characteristic for certain lineages. Novel methods like exper-
imental decay studies or computer tomography might provide additional
insights into the phylogeny, 3D-morphology and ecology of such fossils.
The future of constraining the evolutionary history of Platyzoa and parasitic
flatworms lies in molecular clock methodology by combining information
from the geological record (particularly body fossils or eggs) and molecular
sequences with the fewest assumptions. Characteristic pathologies might also
put constraints on the evolutionary history of parasitic flatworm, although
this still needs to be further studied in extant and fossil hosts to establish a
robust relationship with a particular lineage of parasites (Campbell, 1985;
Ituarte et al., 2001, 2005).

ACKNOWLEDGEMENTS
Michael Calner (Lund University), Ervıns Luksevics (University of Latvia, Riga), Steve Ross-
coe (Hardin-Simmons University), Pascal Godefroit (Royal Belgian Institute of Natural Sci-
ences, Brussels) and Jon Todd (Natural History Museum, London) are thanked for pointing
us to literature with the latest stratigraphic assignment of the Silurian, Upper Devonian,
Carboniferous, Cretaceous and Eocene flatworm fossils, respectively. Jon Todd (Natural
122 Kenneth De Baets et al.

History Museum, London) and Cristian Ituarte (Museo Argentino de Ciencias Naturales,
Buenos Aires) kindly put pictures of bivalve pathologies linked to trematode parasites at
our disposal. This research was partially funded by an SNF grant (2012e141438) to Kenneth
De Baets. Rodney Bray (Natural History Museum, London, retired), John Huntley (Univer-
sity of Missouri, Columbia), Tim Littlewood (Natural History Museum, London) and Rachel
Warnock (Smithsonian National Museum of Natural History, Washington) kindly read and
commented on previous versions of this manuscript.

REFERENCES
Alessandrello, A., Pinna, G., Teruzzi, G., 1988. Land planarian locomotion trail from the
Lower Permian of Lombardian Pre-Alps (Tricladida Terricola). Atti Soc. Ital. Sci. Nat.
Mus. Civ. Stor. Nat. Milano 129.
Allison, C.W., 1975. Primitive fossil flatworm from Alaska: new evidence bearing on
ancestry of the Metazoa. Geology 3, 649e652.
Amin, O.M., 2013. Classification of the acanthocephala. Folia Parasitol. 60, 273e305.
And ujar, C., Soria-Carrasco, V., Serrano, J., Gomez-Zurita, J., 2014. Congruence test of mo-
lecular clock calibration hypotheses based on Bayes factor comparisons. Methods Ecol.
Evol. 5, 226e242.
Araujo, A., Reinhard, K., Fernando Ferreira, L., 2015. Paleoparasitology e human parasites
in ancient material. Adv. Parasitol. 90, 349e387.
Badets, M., Preez, L.D., Verneau, O., 2013. Alternative development in Polystoma gallieni
(Platyhelminthes, Monogenea) and life cycle evolution. Exp. Parasitol. 135, 283e286.
Badets, M., Whittington, I., Lalubin, F., Allienne, J.F., Maspimby, J.L., Bentz, S., Du
Preez, L.H., Barton, D., Hasegawa, H., Tandon, V., Imkongwapang, R., Ohler, A.,
Combes, C., Verneau, O., 2011. Correlating early evolution of parasitic platyhelminths
to Gondwana breakup. Syst. Biol. 60, 762e781.
Baele, J.-M., Godefroit, P., Spagna, P., Dupuis, C., 2012. Geological model and cyclic mass
mortality scenarios for the Lower Cretaceous Bernissart Iguanodon Bonebeds. In:
Godefroit, P. (Ed.), Bernissart Dinosaurs and Early Cretaceous Terrestrial Ecosystems.
Indiana University Press, Bloomington.
Baker, A.J., Pereira, S.L., Paton, T.A., 2007. Phylogenetic relationships and divergence times
of Charadriiformes genera: multigene evidence for the Cretaceous origin of at least 14
clades of shorebirds. Biol. Lett. 3, 205e210.
Becker, R.T., Gradstein, F.M., Hammer, O., 2012. Chapter 22-The Devonian Period. In:
Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G.M. (Eds.), The Geologic Time
Scale. Elsevier, Boston.
Beltrame, M.O., Vieira De Souza, M., Ara ujo, A., Sardella, N.H., 2014. Review of the
rodent paleoparasitological knowledge from South America. Quat. Int. 352, 68e74.
Benton, M.J., Donoghue, P.C.J., Asher, R.J., 2009. Calibrating and constraining molecular
clocks. In: Hedges, B.S., Kumar, S. (Eds.), The Timetree of Life. Oxford University
Press, Oxford.
Bentz, S., Leroy, S., Du Preez, L., Mariaux, J., Vaucher, C., Verneau, O., 2001. Origin and
evolution of African Polystoma (Monogenea: Polystomatidae) assessed by molecular
methods. Int. J. Parasitol. 31, 697e705.
Bentz, S., Sinnappah-Kang, N.D., Lim, L.-H.S., Lebedev, B., Combes, C., Verneau, O.,
2006. Historical biogeography of amphibian parasites, genus Polystoma (Monogenea:
Polystomatidae). J. Biogeogr. 33, 742e749.
Berggren, W.A., Aubry, M.-P., 1996. A late Paleocene-early Eocene NW European and
North Sea magnetobiochronological correlation network. Geol. Soc. Lond. Spec.
Publ. 101, 309e352.
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 123

Bertelli, S., Lindow, B.E.K., Dyke, G.J., Chiappe, L.M., 2010. A well-preserved ‘charadrii-
form-like’ fossil bird from the Early Eocene Fur Formation of Denmark. Palaeontology
53, 507e531.
Bertelli, S., Lindow, B.E.K., Dyke, G.J., Mayr, G., 2013. Another charadriiform-like bird
from the lower Eocene of Denmark. Paleontol. J. 47, 1282e1301.
Bertrand, C.-E., 1903. Les Coprolithes de Bernissart, Premiere Partie: Les Coprolithes qui
ont été attributés aux Iguanodons. Mém. Mus. R. His. Nat. Belg. 1, 1e154.
Blaustein, A.R., Johnson, P.T.J., 2003. The complexity of deformed amphibians. Front.
Ecol. Environ. 1, 87e94.
Blieck, A., Turner, S., Burrow, C.J., Schultze, H.-P., Rexroad, C.B., Bultynck, P.,
Nowlan, G.S., 2010. Fossils, histology, and phylogeny: why conodonts are not
vertebrates. Episodes-Newsmagazine Int. Union Geol. Sci. 33, 234.
Boeger, W., Kritsky, D., 1993. Phylogeny and a revised classification of the Monogenoidea
Bychowsky, 1937 (Platyhelminthes). Syst. Parasitol. 26, 1e32.
Boeger, W.A., Kritsky, D.C., 1997. Coevolution of the Monogenoidea (Platyhelminthes)
based on a revised hypothesis of parasite phylogeny. Int. J. Parasitol. 27, 1495e1511.
Boeger, W.A., Kritsky, D.C., 2001. Phylogenetic relationships of the Monogenoidea. In:
Littlewood, D.T.J., Bray, R.A. (Eds.), Interrelationships of the Platyhelminthes. Taylor
& Francis, London.
Boucot, A.J., 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier,
Amsterdam.
Boucot, A.J., Poinar, G.O., 2010. Host-parasite and host-parasitoid relationships and disease.
In: Boucot, A.J., Poinar, G.O. (Eds.), Fossil Behavior Compendium. CRC Press.
Bourlat, S.J., Nielsen, C., Lockyer, A.E., Littlewood, D.T.J., Telford, M.J., 2003. Xenoturbella
is a deuterostome that eats molluscs. Nature 424, 925e928.
Brasier, M.D., Antcliffe, J.B., 2008. Dickinsonia from Ediacara: a new look at morphology and
body construction. Palaeogeogr. Palaeoclimatol. Palaeoecol. 270, 311e323.
Brett, C.E., Walker, S.E., 2002. Predators and predation in Paleozoic marine environments.
Paleontol. Soc. Pap. 8, 93e118.
Bromham, L., 2009. Why do species vary in their rate of molecular evolution? Biol. Lett. 5,
401e404.
Brooks, D.R., 1989. The phylogeny of the cercomeria (Platyhelminthes: Rhabdocoela) and
general evolutionary principles. J. Parasitol. 75, 606e616.
Brooks, D.R., Mclennan, D.A., 1993. Parascript: Parasites and the Language of Evolution.
Smithsonian Institution Press, Washington.
Buatois, L.A., Mangano, M.G., Mikulas, R., Maples, C.G., 1998. The ichnogenus Curvoli-
thus revisited. J. Paleontol. 72, 758e769.
Bush, A.O., Fernandez, J.C., Esch, G.W., Seed, J.R., 2001. Parasitism: The Diversity and
Ecology of Animal Parasites. Cambridge University Press.
Bychowsky, B.E., 1937. Ontogenesis and phylogenetic interrelationships of parasitic flat-
worms. In: News of the Academy of Sciences, USSR Department of Mathematics
and Natural Sciences, vol. 4, pp. 1353e1383.
Caira, J.N., Reyda, F.B., 2005. Eucestoda (true tapeworms). In: Rohde, K. (Ed.), Marine
Parasitology. CABI Publishing, Oxon.
Calner, M., Jeppsson, L., 2003. Carbonate platform evolution and conodont stratigraphy
during the middle Silurian Mulde Event, Gotland, Sweden. Geol. Mag. 140, 173e203.
Campbell, D., 1985. The life cycle of Gymnophallus Rebecqui (Digenea: Gymnophallidae) and
the response of the bivalve Abra Tenuis to its metacercariae. J. Mar. Biol. Assoc. U.K. 65,
589e601.
Castellani, C., Maas, A., Waloszek, D., Haug, J.T., 2011. New pentastomids from the Late
Cambrian of Sweden e deeper insight of the ontogeny of fossil tongue worms. Palae-
ontogr. Abt. A 293, 95e145.
124 Kenneth De Baets et al.

Cavalier-Smith, T., 1998. A revised six-kingdom system of life. Biol. Rev. 73, 203e266.
Chin, K., 2002. Analyses of coprolites produced by carnivorous vertebrates. Paleontol. Soc.
Pap. 8, 43e50.
Chin, K., Tokaryk, T.T., Erickson, G.M., Calk, L.C., 1998. A king-sized theropod coprolite.
Nature 393, 680e682.
Ching, H.L., 1995. Evaluation of characters of the digenean family Gymnophallidae Moro-
zov, 1955. Can. J. Fish. Aquat. Sci. 52, 78e83.
Christoffersen, M.L., De Assis, J.E., 2013. A systematic monograph of the Recent Pentasto-
mida, with a compilation of their hosts. Zool. Meded. 87, 1e206.
Cisneros, J.C., Abdala, F., Atayman-G€ uven, S., Rubidge, B.S., Şeng€ or, A.M.C.,
Schultz, C.L., 2012. Carnivorous dinocephalian from the Middle Permian of Brazil
and tetrapod dispersal in Pangaea. Proc. Natl. Acad. Sci. 109, 1584e1588.
Clarke, J.T., Warnock, R.C.M., Donoghue, P.C.J., 2011. Establishing a time-scale for plant
evolution. New. Phytol. 192, 266e301.
Collins, A.G., Lipps, J.H., Valentine, J.W., 2000. Modern mucociliary creeping trails and the
bodyplans of Neoproterozoic trace-makers. Paleobiology 26, 47e55.
Combes, C., 2001. Parasitism: The Ecology and Evolution of Intimate Interactions. Univer-
sity of Chicago Press.
Conway Morris, S., 1981. Parasites and the fossil record. Parasitology 83, 489e509.
Conway Morris, S., Crampton, J., Bing, X., Chapman, A., 1997. Lower Cambrian cambro-
claves (incertae sedis) from Xinjiang, China, with comments on the morphological vari-
ability of sclerites. Palaeontology 40, 167e190.
Conway Morris, S., Crompton, D.W.T., 1982. The origins and evolution of Acanthocephala.
Biol. Rev. 57, 85e115.
Cramer, B.D., Condon, D.J., S€ oderlund, U., Marshall, C., Worton, G.J., Thomas, A.T.,
Calner, M., Ray, D.C., Perrier, V., Boomer, I., Patchett, P.J., Jeppsson, L., 2012. U-Pb
(zircon) age constraints on the timing and duration of Wenlock (Silurian) paleocommunity
collapse and recovery during the “Big Crisis”. Geol. Soc. Am. Bull. 124, 1841e1857.
Cribb, T.H., Bray, R.A., Littlewood, D.T.J., 2001. The nature and evolution of the associ-
ation among digeneans, molluscs and fishes. Int. J. Parasitol. 31, 997e1011.
Crisp, M.D., Trewick, S.A., Cook, L.G., 2011. Hypothesis testing in biogeography. Trends
Ecol. Evol. 26, 66e72.
Davis, G.M., 1993. Evolution of prosobranch snails transmitting Asian Schistosoma; coevolu-
tion with Schistosoma: a review. In: Sun, T. (Ed.), Progress in Clinical Parasitology.
Springer, New York.
Davydov, V.I., Korn, D., Schmitz, M.D., Gradstein, F.M., Hammer, O., 2012. Chapter
23-The Carboniferous Period. In: Gradstein, F.M., Ogg, J.G., Schmitz, M.D.,
Ogg, G.M. (Eds.), The Geologic Time Scale. Elsevier, Boston.
De Baets, K., Donoghue, P.C.J., 2012. Molecular clocks and tectonic blocks. Geol. Soc. Am.
Abstr. 44, 331.
De Baets, K., Keupp, H., Klug, C., 2015. Parasites of ammonoids. In: Klug, C., Korn, D., De
Baets, K., Kruta, I., Mapes, R.H. (Eds.), Ammonoid Paleobiology: From Anatomy to
Paleoecology. Springer, The Netherlands.
De Baets, K., Klug, C., Korn, D., 2011. Devonian pearls and ammonoid-endoparasite co-
evolution. Acta Palaeontol. Pol. 56, 159e180.
De Baets, K., Klug, C., Korn, D., Bartels, C., Poschmann, M., 2013. Emsian Ammonoidea
and the age of the Hunsr€ uck Slate (Rhenish Mountains, western Germany). Palaeontogr.
A 299, 1e113.
Deelder, A.M., Miller, R.L., De Jonge, N., Krijger, F.W., 1990. Detection of schistosome
antigen in mummies. Lancet 335, 724e725.
Dentzien-Dias, P.C., De Figueiredo, A.E.Q., Horn, B., Cisneros, J.C., Schultz, C.L., 2012.
Paleobiology of a unique vertebrate coprolites concentration from Rio do Rasto
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 125

formation (Middle/Upper Permian), Parana Basin, Brazil. J. South Am. Earth Sci. 40,
53e62.
Dentzien-Dias, P.C., Poinar Jr., G., De Figueiredo, A.E.Q., Pacheco, A.C.L., Horn, B.L.D.,
Schultz, C.L., 2013. Tapeworm eggs in a 270 million-year-old shark coprolite. PLoS
One 8, e55007.
Despres, L., Imbert-Establet, D., Combes, C., Bonhomme, F., 1992. Molecular evidence
linking hominid evolution to recent radiation of schistosomes (Platyhelminthes:
Trematoda). Mol. Phylogenet. Evol. 1, 295e304.
Dias-Da-Silva, S., 2012. MiddleeLate Permian tetrapods from the Rio do Rasto formation,
Southern Brazil: a biostratigraphic reassessment. Lethaia 45, 109e120.
Dommelier Espejo, S., 2001. Contribution a l’étude paléoparasitologique des sites néolithi-
ques en environnement lacustre dans les domaines jurassien et péri-alpin (Doctoral disser-
tation). Université de Reims Champagne-Ardenne.
Donoghue, M.J., Moore, B.R., 2003. Toward an integrative historical biogeography. Integr.
Comp. Biol. 43, 261e270.
Donoghue, P.C.J., Benton, M.J., 2007. Rocks and clocks: calibrating the Tree of Life using
fossils and molecules. Trends Ecol. Evol. 22, 424e431.
Donoghue, P.C.J., Keating, J.N., 2014. Early vertebrate evolution. Palaeontology 879e893.
Donoghue, P.C.J., Smith, M.P., 2003. The origin and early evolution of chordates: molec-
ular clocks and the fossil record. In: Donoghue, P.C.J., Smith, M.P. (Eds.), Telling the
Evolutionary Time: Molecular Clocks and the Fossil Record. CRC Press, London.
Douzery, E.J.P., Snell, E.A., Bapteste, E., Delsuc, F., Philippe, H., 2004. The timing of
eukaryotic evolution: Does a relaxed molecular clock reconcile proteins and fossils?
Proc. Natl. Acad. Sci. U.S.A. 101, 15386e15391.
Du Preez, L.H., Raharivololoniaina, L., Verneau, O., Vences, M., 2010. A new genus of pol-
ystomatid parasitic flatworm (Monogenea: Polystomatidae) without free-swimming life
stage from the Malagasy poison frogs. Zootaxa 2722, 54e68.
Dyke, G.J., Van Tuinen, M., 2004. The evolutionary radiation of modern birds (Neornithes):
reconciling molecules, morphology and the fossil record. Zool. J. Linn. Soc. 141, 153e177.
Edgecombe, G., Giribet, G., Dunn, C., Hejnol, A., Kristensen, R., Neves, R., Rouse, G.,
Worsaae, K., Sørensen, M., 2011. Higher-level metazoan relationships: recent progress
and remaining questions. Org. Divers. Evol. 11, 151e172.
Egger, B., Lapraz, F., Tomiczek, B., M€ uller, S., Dessimoz, C., Girstmair, J., Skunca, N.,
Rawlinson, K.A., Cameron, C.B., Beli, E., 2015. A Transcriptomic-Phylogenomic
Analysis of the Evolutionary Relationships of Flatworms. Curr. Biol. 25, 1347e1353.
Elicki, O., Wotte, T., 2003. Cambroclaves from the Cambrian of Sardinia (Italy) and Ger-
many: constraints for the architecture of western Gondwana and the palaeogeographical
and palaeoecological potential of cambroclaves. Palaeogeogr. Palaeoclimatol. Palaeoecol.
195, 55e71.
Erwin, D.H., Laflamme, M., Tweedt, S.M., Sperling, E.A., Pisani, D., Peterson, K.J., 2011.
The Cambrian conundrum: early divergence and later ecological success in the early his-
tory of animals. Science 334, 1091e1097.
Falcon-Lang, H.J., Heckel, P.H., Dimichele, W.A., Blake, B.M., Easterday, C.R.,
Eble, C.F., Elrick, S., Gastaldo, R.A., Greb, S.F., Martino, R.L., Nelson, W.J.,
Pfefferkorn, H.W., Phillips, T.L., Rosscoe, S.J., 2011. No major stratigraphic gap exists
near the Middle-Upper Pennsylvanian (Desmoinesian-Missourian) boundary in North
America. Palaios 26, 125e139.
Fedonkin, M., 1985. Sistematicheskoe opisanie vendskikh Metazoa. In: Sokolov, B.S.,
Iwanowski, A.B. (Eds.), Vendskaya Sistema 1: Istoriko-geologicheskoe i paleontologi-
cheskoe obosnovanie. Nauka, Moscow.
Frey, D.G., 1964. Remains of animals in quaternary lake and bog sediments and their
interpretation. Adv. Limnol. 2, 1e114.
126 Kenneth De Baets et al.

Friedman, M., 2007. The interrelationships of Devonian lungfishes (Sarcopterygii: Dipnoi) as


inferred from neurocranial evidence and new data from the genus Soederberghia Lehman,
1959. Zool. J. Linn. Soc. 151, 115e171.
Friedman, M., Keck, B.P., Dornburg, A., Eytan, R.I., Martin, C.H., Hulsey, C.D.,
Wainwright, P.C., Near, T.J., 2013. Molecular and fossil evidence place the origin of
cichlid fishes long after Gondwanan rifting. Proc. R. Soc. B Biol. Sci. 280.
Friedman, M., Sallan, L.C., 2012. Five hundred million years of extinction and recovery: a
phanerozoic survey of large-scale diversity patterns in fishes. Palaeontology 55, 707e742.
Fr€
obisch, N.B., Bickelmann, C., Witzmann, F., 2014. Early evolution of limb regeneration in
tetrapods: evidence from a 300-million-year-old amphibian. Proc. R. Soc. B Biol. Sci. 281.
Fromm, B., Worren, M.M., Hahn, C., Hovig, E., Bachmann, L., 2013. Substantial loss of
conserved and gain of novel MicroRNA families in flatworms. Mol. Biol. Evol. 30,
2619e2628.
Fry, G.F., Hall, H., 1969. Parasitological examination of prehistoric human coprolites from
Utah. Proc. Utah Acad. Sci. Arts Lett. 46, 102e105.
Fugassa, M.H., Beltrame, M.O., Sardella, N.H., Civalero, M.T., Aschero, C., 2010. Paleo-
parasitological results from coprolites dated at the PleistoceneeHolocene transition as
source of paleoecological evidence in Patagonia. J. Archaeol. Sci. 37, 880e884.
Gayet, M., Rage, J.-C., Sempere, T., Gagnier, P., 1992. Modalités des échanges de vertébrés
continentaux entre l’Amérique du nord et l’Amérique du sud au Crétacé supérieur et au
Paléocene. Bull. Soc. Géol. Fr. 163, 781e791.
Geyer, G., Hautmann, M., Hagdorn, H., Ockert, W., Streng, M., 2005. Well-preserved
mollusks from the Lower Keuper (Ladinian) of Hohenlohe (Southwest Germany). Pal-
aontologische Z. 79, 429e460.
Golvan, Y.J., 1969. Systématique des acanthocéphales (Acanthocephala Rudolphi 1801).
Mém. Mus. Natl. Hist. Nat. Sér. A LVII fasc. unique, 1e373.
Gonçalves, M.L.C., Ara ujo, A., Ferreira, L.F., 2003. Human intestinal parasites in the past:
new findings and a review. Mem. Inst. Oswaldo Cruz 98, 103e118.
Goswami, A., Upchurch, P., 2010. The dating game: a reply to Heads (2010). Zool. Scr. 39,
406e409.
G€otting, K.-J., 1974. Malakozoologie. Grundriss der Weichtierkunde. Fischer, Stuttgart.
G€otting, K.-J., 1979. Durch Parasiten induzierte Perlbildung bei Mytilus edulis L. (Bivalvia).
Malacologia 18, 563e567.
Goudemand, N., Orchard, M.J., Urdy, S., Bucher, H., Tafforeau, P., 2011. Synchrotron-
aided reconstruction of the conodont feeding apparatus and implications for the mouth
of the first vertebrates. Proc. Natl. Acad. Sci. 108, 8720e8724.
Gray, J., 1988. Evolution of the freshwater ecosystem: the fossil record. Palaeogeogr. Palae-
oclimatol. Palaeoecol. 62, 1e214.
Gusev, A., 1992. A response to Malmberg. Syst. Parasitol. 21, 167e168.
Hafner, M., Sudman, P., Villablanca, F., Spradling, T., Demastes, J., Nadler, S., 1994.
Disparate rates of molecular evolution in cospeciating hosts and parasites. Science
265, 1087e1090.
Hafner, M.S., Page, R.D.M., 1995. Molecular phylogenies and host-parasite cospeciation:
gophers and lice as a model system. Philos. Trans. R. Soc. Lond. Ser. B Biol. Sci. 349,
77e83.
Hahn, C., Fromm, B., Bachmann, L., 2014. Comparative genomics of flatworms (Platyhel-
minthes) reveals shared genomic features of ecto- and endoparastic neodermata. Genome
Biol. Evol. 6, 1105e1117.
Harmsworth, R.V., 1968. The developmental history of Blelham Tarn (England) as shown
by animal microfossils, with special reference to the Cladocera. Ecol. Monogr. 223e241.
Hausdorf, B., 2000. Early evolution of the bilateria. Syst. Biol. 49, 130e142.
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 127

Hejnol, A., Obst, M., Stamatakis, A., Ott, M., Rouse, G.W., Edgecombe, G.D.,
Martinez, P., Bagu~ na, J., Bailly, X., Jondelius, U., Wiens, M., M€ uller, W.E.G.,
Seaver, E., Wheeler, W.C., Martindale, M.Q., Giribet, G., Dunn, C.W., 2009. Assess-
ing the root of bilaterian animals with scalable phylogenomic methods. Proc. R. Soc. B
276, 4261e4270.
Henderson, C.M., Davydov, V.I., Wardlaw, B.R., Gradstein, F.M., Hammer, O., 2012.
Chapter 24-The Permian Period. In: Gradstein, F.M., Ogg, J.G., Schmitz, M.D.,
Ogg, G.M. (Eds.), The Geologic Time Scale. Elsevier, Boston.
Héritier, L., Badets, M., Du Preez, L.H., Aisien, M.S.O., Lixian, F., Combes, C.,
Verneau, O., 2015. Evolutionary processes involved in the diversification of chelonian
and mammal polystomatid parasites (Platyhelminthes, Monogenea, Polystomatidae)
revealed by palaeoecology of their hosts. Molecular Phylogenetics and Evolution 92,
1e10.
Hibbett, D., Matheny, P.B., 2009. The relative ages of ectomycorrhizal mushrooms and their
plant hosts estimated using Bayesian relaxed molecular clock analyses. BMC Biol. 7, 13.
Hipsley, C.A., M€ uller, J., 2014. Beyond fossil calibrations: realities of molecular clock prac-
tices in evolutionary biology. Fron. Genet. 5.
Hoberg, E., 1999. Systematics of the Eucestoda: advances toward a new phylogenetic para-
digm, and observations on the early diversification of tapeworms and vertebrates. Syst.
Parasitol. 42, 1e12.
Hoberg, E.P., Jones, A., Bray, R.A., 1999. Phylogenetic analysis among the families of the
Cyclophyllidea (Eucestoda) based on comparative morphology, with new hypotheses
for co-evolution in vertebrates. Syst. Parasitol. 42, 51e73.
Holz, M., França, A.B., Souza, P.A., Iannuzzi, R., Rohn, R., 2010. A stratigraphic chart of
the Late Carboniferous/Permian succession of the eastern border of the Parana Basin,
Brazil, South America. J. South Am. Earth Sci. 29, 381e399.
Hone, D.W.E., Rauhut, O.W.M., 2010. Feeding behaviour and bone utilization by
theropod dinosaurs. Lethaia 43, 232e244.
Hunt, A.P., Lucas, S.G., Milan, J., Spielmann, J.A., 2012. Vertebrate coprolite studies: status
and prospectus. N.M. Mus. Nat. Hist. Sci. Bull. 57, 5e24.
Huntley, J.W., 2007. Towards establishing a modern baseline for paleopathology: trace-
producing parasites in a bivalve host. J. Shellfish Res. 26, 253e259.
Huntley, J.W., De Baets, K., 2015. Trace fossil evidence of trematodeebivalve parasiteehost
interactions in deep time. Adv. Parasitol. 90, 201e231.
Huntley, J.W., F€ ursich, F.T., Alberti, M., Hethke, M., Liu, C., 2014. A complete Holocene
record of trematodeebivalve infection and implications for the response of parasitism to
climate change. Proc. Natl. Acad. Sci. 111, 18150e18155.
Huntley, J.W., Scarponi, D., 2012. Evolutionary and ecological implications of trematode
parasitism of modern and fossil northern Adriatic bivalves. Paleobiology 38, 40e51.
Huntley, J.W., Scarponi, D., 2015. Geographic variation of parasitic and predatory traces on
mollusks in the northern Adriatic Sea, Italy: implications for the stratigraphic paleobi-
ology of biotic interactions. Paleobiology 41, 134e153.
Huyse, T., Volckaert, F.A.M., 2005. Comparing host and parasite phylogenies: Gyrodactylus
flatworms jumping from goby to goby. Syst. Biol. 54, 710e718.
Hypsa, V., 2006. Parasite histories and novel phylogenetic tools: alternative approaches to
inferring parasite evolution from molecular markers. Int. J. Parasitol. 36, 141e155.
Iles, J.D., 1980. Autopsy of an Egyptian mummy (NakhteROM I). Can. Med. Assoc. J. 122,
512e513.
Ituarte, C., Cremonte, F., Zelaya, D.G., 2005. Parasite-mediated shell alterations in Recent
and Holocene sub-Antarctic bivalves: the parasite as modeler of host reaction. Invertebr.
Biol. 124, 220e229.
128 Kenneth De Baets et al.

Ituarte, C.F., Cremonte, F., Deferrari, G., 2001. Mantle-shell complex reactions elicited by
digenean metacercariae in Gaimardia trapesina (Bivalvia : Gaimardiidae) from the South-
western Atlantic Ocean and Magellan Strait. Dis. Aquat. Org. 48, 47e56.
Iturralde-Vinent, M.A., 2001. Geology of the amber-bearing deposits of the greater Antilles.
Carribean J. Sci. 37, 141e166.
Iturralde-Vinent, M.A., Macphee, R.D.E., 1996. Age and paleogeographical origin of
Dominican amber. Science 273, 1850e1852.
Jeppsson, L., Eriksson, M.E., Calner, M., 2006. A latest Llandovery to latest Ludlow high-res-
olution biostratigraphy based on the Silurian of Gotlandda summary. GFF 128, 109e114.
Johannessen, O.H., 1973. Deformations of the inner shell surface of Venerupis pullastra (Mon-
tagu) (Lamellibranghia) as a result of infection by a trematod metacercaria with a note of
Parasitism Leading to Parasitic Castration. Sarsia 52, 117e122.
Johnson, P.T., Lunde, K.B., Haight, R.W., Bowerman, J., Blaustein, A.R., 2001. Ribeiroia
ondatrae (Trematoda: Digenea) infection induces severe limb malformations in western
toads (Bufo boreas). Can. J. Zool. 79, 370e379.
Johnson, P.T.J., Chase, J.M., 2004. Parasites in the food web: linking amphibian malforma-
tions and aquatic eutrophication. Ecol. Lett. 7, 521e526.
Johnson, P.T.J., Lunde, K.B., Thurman, E.M., Ritchie, E.G., Wray, S.N., Sutherland, D.R.,
Kapfer, J.M., Frest, T.J., Bowerman, J., Blaustein, A.R., 2002. Parasite (Ribeiroia ondatrae)
infection linked to amphibian malformations in the western United States. Ecol.
Monogr. 72, 151e168.
Johnson, P.T.J., Sutherland, D.R., 2003. Amphibian deformities and Ribeiroia infection: an
emerging helminthiasis. Trends Parasitol. 19, 332e335.
Jondelius, U., Ruiz-Trillo, I., Bagu~ na, J., Riutort, M., 2002. The Nemertodermatida are
basal bilaterians and not members of the Platyhelminthes. Zool. Scr. 31, 201e215.
Jouy-Avantin, F., Combes, C., Lumley, H., Miskovsky, J.C., Moné, H., 1999. Helminth
eggs in animal coprolites from a Middle Pleistocene site in Europe. J. Parasitol. 85,
376e379.
Jurina, A., Raskatova, M., 2012. New data on the Devonian plant and miospores from the
Lode formation, Latvia. Sci. Pap. Univ. Latv. 783, 46e56.
Justine, J.-L., 1998. Non-monophyly of the monogeneans? Int. J. Parasitol. 28, 1653e1657.
Khalil, L.F., Jones, A., Bray, R.A., 1994. Keys to the Cestode Parasites of Vertebrates. CAB
International, Wallingford.
Knaust, D., 2010. Remarkably preserved benthic organisms and their traces from a Middle
Triassic (Muschelkalk) mud flat. Lethaia 43, 344e356.
Kodandaramaiah, U., 2011. Tectonic calibrations in molecular dating. Curr. Zool. 57,
116e124.
Koprivnikar, J., Marcogliese, D., Rohr, J., Orlofske, S., Raffel, T., Johnson, P.J., 2012. Mac-
roparasite infections of amphibians: what can they tell us? EcoHealth 9, 342e360.
Kouchinsky, A., Bengtson, S., Runnegar, B., Skovsted, C., Steiner, M., Vendrasco, M.,
2012. Chronology of early Cambrian biomineralization. Geol. Mag. 149, 221e251.
Kríz, J., 1979. Silurian Cardiolidae (Bivalvia). Sb. Geol. Ved Paleontol. 22, 1e157.
Ksepka, D.T., Ware, J.L., Lamm, K.S., 2014. Flying rocks and flying clocks: disparity in fossil
and molecular dates for birds. Proc. R. Soc. B 281, 20140677.
Kutassy, E., 1937. Die €alteste fossile Perle und Verletzungsspuren an einem triadischen
Megalodus. Math. Naturwiss. Anz. Ung. Akad. Wiss. 55, 1005e1023.
Labandeira, C.C., 2002. Paleobiology of predators, parasitoids, and parsites: death and
accomodation in the fossil record of continental invertebrates. Paleontol. Soc. Pap. 8,
211e250.
Lauckner, G., 1983. Diseases of Mollusca: Bivalvia. In: Kinne, O. (Ed.), Diseases of Marine
Animals, Volume II: Introduction, Bivalvia to Scaphopoda. Biologische Anstalt Helgo-
land, Hamburg.
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 129

Laumer, C.E., Giribet, G., 2014. Inclusive taxon sampling suggests a single, stepwise origin of
ectolecithality in Platyhelminthes. Biol. J. Linn. Soc. 111, 570e588.
Laumer, C.E., Hejnol, A., Giribet, G., 2015. Nuclear genomic signals of the ‘microturbellar-
ian’roots of platyhelminth evolutionary innovation. eLife 4, e05503.
Lawton, S., Hirai, H., Ironside, J., Johnston, D., Rollinson, D., 2011. Genomes and geog-
raphy: genomic insights into the evolution and phylogeography of the genus
Schistosoma. Parasites Vectors 4, 131.
Lee, S.-H., Chai, J.-Y., 2001. A review of Gymnophalloides seoi (Digenea: Gymnophallidae)
and human infections in the Republic of Korea. Korean J. Parasitol. 39, 85e118.
Light, J.E., Hafner, M.S., 2007. Cophylogeny and disparate rates of evolution in sympatric
lineages of chewing lice on pocket gophers. Mol. Phylogenet. Evol. 45, 997e1013.
Liljedahl, L., 1985. Ecological aspects of a silicified bivalve fauna from the Silurian of Gotland.
Lethaia 18, 53e66.
Liljedahl, L., 1994. Silurian nuculoid and modiomorphid bivalves from Sweden. Fossils Strata
33, 1e89.
Littlewood, D.T.J., 2006. The evolution of parasitism in flatworms. In: Maule, A.G.,
Marks, N.J. (Eds.), Parasitic Flatworms: Molecular Biology, Biochemistry, Immunology
and Physiology.
Littlewood, D.T.J., 2008. Platyhelminth systematics and the emergence of new characters.
Parasite J. Soc. Fr. Parasitol. 15, 333e341.
Littlewood, D.T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
Littlewood, D.T.J., Rohde, K., Bray, R.A., Herniou, E.A., 1999a. Phylogeny of the Platy-
helminthes and the evolution of parasitism. Biol. J. Linn. Soc. 68, 257e287.
Littlewood, D.T.J., Rohde, K., Clough, K.A., 1999b. The interrelationships of all major
groups of Platyhelminthes: phylogenetic evidence from morphology and molecules.
Biol. J. Linn. Soc. 66, 75e114.
Littlewood, D.T.J., Waeschenbach, A., 2015. Evolution: A Turn Up for the Worms. Curr.
Biol. 25, R457eR460.
Llewellyn, J., 1987. Phylogenetic inference from platyhelminth life-cycle stages. Int. J. Para-
sitol. 17, 281e289.
Loader, S.P., Pisani, D., Cotton, J.A., Gower, D.J., Day, J.J., Wilkinson, M., 2007. Relative
time scales reveal multiple origins of parallel disjunct distributions of African caecilian
amphibians. Biol. Lett. 3, 505e508.
Lockyer, A.E., Olson, P.D., Littlewood, D.T.J., 2003a. Utility of complete large and small
subunit rRNA genes in resolving the phylogeny of the Neodermata (Platyhel-
minthes): implications and a review of the cercomer theory. Biol. J. Linn. Soc. 78,
155e171.
Lockyer, A.E., Olson, P.D., Ostergaard, P., Rollinson, D., Johnston, D.A., Attwood, S.W.,
Southgate, V.R., Horak, P., Snyder, S.D., Le, T.H., Agatsuma, T., Mcmanus, D.P.,
Carmichael, A.C., Naem, S., Littlewood, D.T.J., 2003b. The phylogeny of the Schisto-
somatidae based on three genes with emphasis on the interrelationships of Schistosoma
Weinland, 1858. Parasitology 126, 203e224.
Loss-Oliveira, L., Aguiar, B.O., Schrago, C.G., 2012. Testing synchrony in historical
biogeography: the case of new world primates and hystricognathi rodents. Evol. Bio-
inform. 8, 127.
Luksevics, E., Lebedev, O., Mark-Kurik, E., Karataj€ ute-Talimaa, V., 2009. The earliest
evidence of host-parasite interactions in vertebrates. Acta Zool. 90, 335e343.
Luksevics, E., Stinkulis, Ģ., Murnieks, A., Popovs, K., 2012. Geological evolution of the
Baltic Artesian Basin. In: Delina, A., Kalvans, A., Saks, T., Bethers, U., Vircavs, V.
(Eds.), Highlights of Groundwater Research in the Baltic Artesian Basin. University
of Latvia.
130 Kenneth De Baets et al.

Luksevics, E., Stinkulis, Ģ., Saks, T., Popovs, K., Jatnieks, J., 2014. The Devonian strati-
graphic succession and evolution of the Baltic sedimentary basin. In: Rocha, R.,
Pais, J., Kullberg, J.C., Finney, S. (Eds.), STRATI 2013. Springer International
Publishing.
Lumley, H.D., Fournier, A., Park, Y., Yokoyama, Y., Demouy, A., 1984. Stratigraphie du
remplissage Pléistocene moyen de la Caune de l’Arago a Tautavel. E  tude de huit carot-
tages effectués de 1981 a 1983. L’Anthropologie 88, 5e18.
Lunde, K.B., Johnson, P.T.J., 2012. A practical guide for the study of malformed amphibians
and their causes. J. Herpetol. 46, 429e441.
Maas, A., Waloszek, D., 2001. Cambrian derivatives of the early arthropod stem lineage,
pentastomids, tardigrades and lobopodians e an ‘Orsten’ perspective. Zool. Anz. 240,
451e459.
Malmberg, G., 1990. On the ontogeny of the haptor and the evolution of the Monogenea.
Syst. Parasitol. 17, 1e65.
Mark-Kurik, E., Blieck, A., Loboziak, S., Candilier, A.-M., 1999. Miospore assemblage from
the Lode Member (Gauja Formation) in Estonia and the MiddleeUpper Devonian
boundary problem. Proc. Acad. Sci. Est. Geol. 48, 86e98.
Mark-Kurik, E., P~ oldvere, A., 2012. Devonian stratigraphy in Estonia: current state and
problems. Est. J. Earth Sci. 61, 33e47.
Martínez-Aquino, A., Ceccarelli, F.S., Eguiarte, L.E., Vazquez-Domínguez, E., De
Leon, G.P.-P., 2014. Do the historical biogeography and evolutionary history of the
digenean Margotrema spp. across central Mexico mirror those of their freshwater fish hosts
(Goodeinae)? PLoS One 9, e101700.
Mayr, G., 2011. The phylogeny of charadriiform birds (shorebirds and allies) e reassessing the
conflict between morphology and molecules. Zool. J. Linn. Soc. 161, 916e934.
Melchin, M.J., Sadler, P.M., Cramer, B.D., Cooper, R.A., Gradstein, F.M., Hammer, O.,
2012. Chapter 21-The Silurian Period. In: Gradstein, F.M., Ogg, J.G.,
Schmitz, M.D., Ogg, G.M. (Eds.), The Geologic Time Scale. Elsevier, Boston.
Mello, B., Schrago, C.G., Mello, B., Schrago, C.G., 2014. Assignment of calibration infor-
mation to deeper phylogenetic nodes is more effective in obtaining precise and accurate
divergence time estimates. Evol. Bioinform. 10, 79e85.
Mollaret, I., Jamieson, B.G.M., Adlard, R.D., Hugall, A., Lecointre, G., Chombard, C.,
Justine, J.-L., 1997. Phylogenetic analysis of the Monogenea and their relationships
with Digenea and Eucestoda inferred from 28S rDNA sequences. Mol. Biochem. Para-
sitol. 90, 433e438.
Moran, N.A., Munson, M.A., Baumann, P., Ishikawa, H., 1993. A molecular clock in endo-
symbiotic bacteria is calibrated using the insect hosts. Proc. R. Soc. Lond. Ser. B Biol. Sci.
253, 167e171.
Moran, N.A., Von Dohlen, C.D., Baumann, P., 1995. Faster evolutionary rates in endosym-
biotic bacteria than in cospeciating insect hosts. J. Mol. Evol. 41, 727e731.
Murdock, D.J.E., Dong, X.-P., Repetski, J.E., Marone, F., Stampanoni, M.,
Donoghue, P.C.J., 2013. The origin of conodonts and of vertebrate mineralized
skeletons. Nature 502, 546e549.
Murienne, J., Daniels, S.R., Buckley, T.R., Mayer, G., Giribet, G., 2014. A living fossil tale
of Pangaean biogeography. Proc. R. Soc. B Biol. Sci. 281.
Mwinyi, A., Bailly, X., Bourlat, S.J., Jondelius, U., Littlewood, D.T.J., Podsiadlowski, L.,
2010. The phylogenetic position of Acoela as revealed by the complete mitochondrial
genome of Symsagittifera roscoffensis. BMC Evol. Biol. 10.
Near, T.J., 2002. Acanthocephalan phylogeny and the evolution of parasitism. Integr. Comp.
Biol. 42, 668e677.
Near, T.J., Meylan, P.A., Shaffer, H.B., 2005. Assessing concordance of fossil calibration
points in molecular clock studies: an example using turtles. Am. Nat. 165, 137e146.
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 131

Ogg, J.G., 2012. Chapter 5-Geomagnetic Polarity Time Scale. The Geologic Time Scale.
Elsevier, Boston.
Ogg, J.G., Hinnov, L.A., Huang, C., 2012. Chapter 27-Cretaceous. In: Gradstein, F.M.,
Ogg, J.G., Schmitz, M.D., Ogg, G.M. (Eds.), The Geologic Time Scale. Elsevier, Boston.
Olson, P.D., Caira, J.N., Jensen, K., Overstreet, R.M., Palm, H.W., Beveridge, I., 2010.
Evolution of the trypanorhynch tapeworms: parasite phylogeny supports independent
lineages of sharks and rays. Int. J. Parasitol. 40, 223e242.
Olson, P.D., Tkach, V.V., 2005. Advances and trends in the molecular systematics of the
parasitic Platyhelminthes. Adv. Parasitol. 60, 165e243.
Orélis-Ribeiro, R., Arias, C.R., Halanych, K.M., Cribb, T.H., Bullard, S.A., 2014. Chapter
one - diversity and ancestry of flatworms infecting blood of nontetrapod craniates “fishes”.
In: Rollinson, D., Stothard, J.R. (Eds.), Advances in Parasitology. Academic Press.
Otsuka, J., Sugaya, N., 2003. Advanced formulation of base pair changes in the stem regions
of ribosomal RNAs; its application to mitochondrial rRNAs for resolving the phylogeny
of animals. J. Theor. Biol. 222, 447e460.
Page, R.D.M., Lee, P.L.M., Becher, S.A., Griffiths, R., Clayton, D.H., 1998. A different
tempo of mitochondrial DNA evolution in birds and their parasitic lice. Mol. Phyloge-
net. Evol. 9, 276e293.
Palij, V., Posti, E., Fedonkin, M., 1979. Myagkotelye metazoa i iskopaemye sledy zhivot-
nykh venda i rannego kembriya. Paleontol. Verkhnedokembrijskikh kembrijskikh otloz-
henij Vost. Evr. platformy 49e82.
Papadopoulou, A., Anastasiou, I., Vogler, A.P., 2010. Revisiting the insect mitochondrial
molecular clock: the mid-Aegean trench calibration. Mol. Biol. Evol. 27, 1659e1672.
Parham, J.F., Donoghue, P.C.J., Bell, C.J., Calway, T.D., Head, J.J., Holroyd, P.A.,
Inoue, J.G., Irmis, R.B., Joyce, W.G., Ksepka, D.T., Patané, J.S.L., Smith, N.D.,
Tarver, J.E., Van Tuinen, M., Yang, Z., Angielczyk, K.D., Greenwood, J.M.,
Hipsley, C.A., Jacobs, L., Makovicky, P.J., M€ uller, J., Smith, K.T., Theodor, J.M.,
Warnock, R.C.M., Benton, M.J., 2012. Best practices for justifying fossil calibrations.
Syst. Biol. 61, 346e359.
Park, J.K., Kim, K.H., Kang, S., Kim, W., Eom, K.S., Littlewood, D.T.J., 2007. A common
origin of complex life cycles in parasitic flatworms: evidence from the complete mito-
chondrial genome of Microcotyle sebastis (Monogenea : Platyhelminthes). BMC Evol.
Biol. 7, 11.
Paton, T.A., Baker, A.J., Groth, J.G., Barrowclough, G.F., 2003. RAG-1 sequences resolve
phylogenetic relationships within Charadriiform birds. Mol. Phylogenet. Evol. 29,
268e278.
Perkins, E., 2010. Family Ties: Molecular Phylogenetics, Evolution and Radiation of Flat-
worm Parasites (Monogenea: Capsalidae) (Ph.D. thesis), University of Adelaide. https://
digital.library.adelaide.edu.au/dspace/handle/2440/60948.
Perkins, E.M., Donnellan, S.C., Bertozzi, T., Whittington, I.D., 2010. Closing the mito-
chondrial circle on paraphyly of the Monogenea (Platyhelminthes) infers evolution in
the diet of parasitic flatworms. Int. J. Parasitol. 40, 1237e1245.
Peterson, K.J., Cotton, J.A., Gehling, J.G., Pisani, D., 2008. The Ediacaran emergence of
bilaterians: congruence between the genetic and the geological fossil records. Phil. Trans.
R. Soc. B Biol. Sci. 363, 1435e1443.
Peterson, K.J., Lyons, J.B., Nowak, K.S., Takacs, C.M., Wargo, M.J., Mcpeek, M.A., 2004.
Estimating metazoan divergence times with a molecular clock. Proc. Natl. Acad. Sci.
U.S.A. 101, 6536e6541.
Petit, G., 2010. Skin nodules in fossil fishes from Monte Bolca (Eocene, Northern Italy).
Geodiversitas 32, 157e163.
Petit, G., Khalloufi, B., 2012. Paleopathology of a fossil fish from the Solnhofen Lagerst€atte
(Upper Jurassic, southern Germany). Int. J. Paleopathol. 2, 42e44.
132 Kenneth De Baets et al.

Philippe, H., Brinkmann, H., Copley, R.R., Moroz, L.L., Nakano, H., Poustka, A.J.,
Wallberg, A., Peterson, K.J., Telford, M.J., 2011. Acoelomorph flatworms are deutero-
stomes related to Xenoturbella. Nature 470, 255e258.
Pierce, W., 1960. Silicified turbellaria from Calico Mountains nodules. Bull. South. Calif.
Acad. Sci. 59, 138e143.
Poinar, G., 2003. A rhabdocoel turbellarian (Platyhelminthes, Typhloplanoida) in Baltic amber
with a review of fossil and sub-fossil platyhelminths. Invertebr. Biol. 122, 308e312.
Poinar, G., Boucot, A.J., 2006. Evidence of intestinal parasites of dinosaurs. Parasitology 133,
245e249.
Poulin, R., 2005. Evolutionary trends in body size of parasitic flatworms. Biol. J. Linn. Soc.
85, 181e189.
Purnell, M.A., Donoghue, P.C., 2005. Between death and data: biases in interpretation of the
fossil record of conodonts. Spec. Pap. Palaeontol. 73, 7e25.
Purnell, M.A., Donoghue, P.C.J., 1997. Architecture and functional morphology of the skel-
etal apparatus of ozarkodinid conodonts. Phil. Trans. R. Soc. Lond. Ser. B Biol. Sci. 352,
1545e1564.
Qian, Y., Yin, G., 1984. Zhijinitae and its stratigraphical significance. Acta Palaeontol. Sin.
23, 216e223.
Qiao, T., Zhu, M., 2009. A new tooth-plated lungfish from the Middle Devonian of
Yunnan, China, and its phylogenetic relationships. Acta Zool. 90, 236e252.
Rage, J.-C., 1988. Gondwana, Tethys, and terrestrial vertebrates during the Mesozoic and
Cainozoic. Geol. Soc. Lond. Spec. Publ. 37, 255e273.
Rakoci nski, M., 2012. The youngest Devonian record of “Housean pits” in ammonoids.
Geol. Q. 56, 387e390.
Reinhard, K.J., 1992. Parasitology as an interpretive tool in archaeology. Am. Antiq. 57,
231e245.
Retallack, G.J., 2007. Growth, decay and burial compaction of Dickinsonia, an iconic Edi-
acaran fossil. Alcheringa Australas. J. Palaeontol. 31, 215e240.
Reyman, T.A., Zimmerman, M.R., Lewin, P.K., 1977. Autopsy of an Egyptian mummy. 5.
Histopathologic investigation. Can. Med. Assoc. J. 117, 470e472.
Roever-Bonnet, H.D., Rijpstra, A., Van Renesse, M., Peen, C., 1979. Helminth eggs and
gregarines from coprolites from the excavations at Swifterbant. Swifterbant Contribution
10. Helinium Wetteren 19, 7e12.
Rohde, K., 2005. Amphilinidea (unsegmented tapeworms). In: Rohde, K. (Ed.), Marine
Parasitology. CABI Publishing, Oxon.
Rosscoe, S.J., 2008. Idiognathodus and Streptognathodus Species from the Lost Branch to
Dewey Sequences (Middle-Upper Pennsylvanian) of the Midcontinent Basin, North
America (Doctor of Philosophy). Texas Tech University. https://repositories.tdl.org/
ttu-ir/handle/2346/9102.
Rosscoe, S.J., Barrick, J.E., 2013. North American species of the conodont genus Idiogna-
thodus from the Moscovian-Kasimovian boundary composite sequence and correlation
of the Moscovian-Kasimovian stage boundary. N.M. Mus. Nat. Hist. Sci. Bull. 60,
354e371.
Rouse, G.W., 2005. Fossil parasites. In: Rohde, K. (Ed.), Marine Parasitology. CABI Pub-
lishing, Oxon.
Rowe, T.B., 2004. Chordate phylogeny and development. In: Donoghue, M.J., Cracraft, J.
(Eds.), Assembling the Tree of Life. Oxford University Press, Oxford.
Ruiz-Trillo, I., Riutort, M., Littlewood, D.T.J., Herniou, E.A., Baguna, J., 1999. Acoel flat-
worms: earliest extant bilaterian metazoans, not members of Platyhelminthes. Science
283, 1919e1923.
Ruiz, G.M., 1991. Consequences of parasitism to marine invertebrates: host evolution? Am.
Zool. 31, 831e839.
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 133

Ruiz, G.M., Lindberg, D.R., 1989. A fossil record for trematodes: extent and potential uses.
Lethaia 22, 431e438.
Sanders, K.L., Lee, M.S.Y., 2010. Arthropod molecular divergence times and the Cambrian
origin of pentastomids. Syst. Biodivers. 8, 63e74.
Sanmartín, I., 2012. Historical biogeography: evolution in time and space. Evol. Educ.
Outreach 5, 555e568.
Sardella, N.H., Fugassa, M.H., Rindel, D.D., Go~ ni, R.A., 2010. Paleoparasitological results
for rodent coprolites from Santa Cruz Province, Argentina. Mem. Inst. Oswaldo Cruz
105, 33e40.
Schmidt, G.D., 1986. CRC Handbook of Tapeworm Identification. CRC Press, Boca Raton.
Schnyder, J., Dejax, J., Keppens, E., Nguyen, T.U.,T.T., Spagna, P., Boulila, S., Galbrun, B.,
Riboulleau, A., Tshibangu, J.-P., Yans, J., 2009. An Early Cretaceous lacustrine record:
organic matter and organic carbon isotopes at Bernissart (Mons Basin, Belgium). Palae-
ogeogr. Palaeoclimatol. Palaeoecol. 281, 79e91.
Searcey, N., Reinhard, K.J., Egarter-Vigl, E., Maixner, F., Piombino-Mascali, D.,
Zink, A.R., Van Der Sanden, W., Gardner, S.L., Bianucci, R., 2013. Parasitism of the
Zweeloo Woman: Dicrocoeliasis evidenced in a Roman period bog mummy. Int. J.
Paleopathol. 3, 224e228.
Seilacher, A., 2007. Trace Fossil Analysis. Springer.
Shinn, A.P., Bron, J.E., Sommerville, C., Gibson, D.I., 2003. Comments on the mecha-
nism of attachment in species of the monogenean genus Gyrodactylus. Invertebr.
Biol. 122, 1e11.
Siveter, D.J., Briggs, D.E.G., Siveter, D.J., Sutton, M.D., 2015. A 425-million-year-old Silu-
rian pentastomid parasitic on ostracods. Curr. Biol. 25, 1632e1637.
Snyder, S.D., Loker, E.S., 2000. Evolutionary relationships among the Schistosomatidae
(Platyhelminthes: Dignea) and an Asian origin for Schistosoma. J. Parasitol. 86, 283e288.
Sperling, E.A., Vinther, J., 2010. A placozoan affinity for Dickinsonia and the evolution of late
Proterozoic metazoan feeding modes. Evol. Dev. 12, 201e209.
Stopper, G.F., Hecker, L., Franssen, R.A., Sessions, S.K., 2002. How trematodes cause limb
deformities in amphibians. J. Exp. Zool. 294, 252e263.
Struck, T.H., Wey-Fabrizius, A.R., Golombek, A., Hering, L., Weigert, A., Bleidorn, C.,
Klebow, S., Iakovenko, N., Hausdorf, B., Petersen, M., K€ uck, P., Herlyn, H.,
Hankeln, T., 2014. Platyzoan paraphyly based on phylogenomic data supports a non-
coelomate ancestry of spiralia. Mol. Biol. Evol. 31, 1833e1849.
Taraschewski, H., 2005. Acanthocephala (thorny or spiny-headed worms). In: Rohde, K.
(Ed.), Marine Parasitology. CABI Publishing, Oxon.
Taylor, R.E., 1985. The beginnings of radiocarbon dating in American antiquity: a historical
perspective. Am. Antiq. 50, 309e325.
Telford, M.J., Lockyer, A.E., Cartwright-Finch, C., Littlewood, D.T.J., 2003. Combined
large and small subunit ribosomal RNA phylogenies support a basal position of the acoe-
lomorph flatworms. Proc. R. Soc. B Biol. Sci. 270, 1077e1083.
Thomas, G., Wills, M., Szekely, T., 2004. A supertree approach to shorebird phylogeny.
BMC Evol. Biol. 4, 28.
Thomas, J.A., Welch, J.J., Lanfear, R., Bromham, L., 2010. A generation time effect on the
rate of molecular evolution in invertebrates. Mol. Biol. Evol. 27, 1173e1180.
Thomas, J.A., Welch, J.J., Woolfit, M., Bromham, L., 2006. There is no universal molecular
clock for invertebrates, but rate variation does not scale with body size. Proc. Natl. Acad.
Sci. 103, 7366e7371.
Todd, J.A., Harper, E.M., 2011. Stereotypic boring behaviour inferred from the earliest
known octopod feeding traces: Early Eocene, southern England. Lethaia 44, 214e222.
Trewick, S.A., Gibb, G.C., 2010. Vicars, tramps and assembly of the New Zealand avifauna:
a review of molecular phylogenetic evidence. Ibis 152, 226e253.
134 Kenneth De Baets et al.

Turner, S., Burrow, C.J., Schultze, H.-P., Blieck, A., Reif, W.-E., Rexroad, C.B.,
Bultynck, P., Nowlan, G.S., 2010. False teeth: conodont-vertebrate phylogenetic rela-
tionships revisited. Geodiversitas 32, 545e594.
Upeniece, I., 1996. Lodeacanthus gaujicus ng et sp. (Acanthodii: Mesacanthidae) from the Late
Devonian of Latvia. Mod. Geol. 20, 383e398.
Upeniece, I., 1998. The first finds of fossil parasitic flatworms (Platyhelminthes). Ichthyolith
Issues Spec. Publ. 4, 53e55.
Upeniece, I., 1999. Fossil record of parasitic helminths in fishes. In: 5th International Sym-
posium on Fish Parasites: Abstracts, 154.
Upeniece, I., 2001. The unique fossil assemblage from the Lode Quarry (Upper Devonian,
Latvia). Foss. Rec. 4, 101e119.
Upeniece, I., 2011. Palaeoecology and Juvenile Individuals of the Devonian Placoderm and
Acanthodian Fishes from Lode Site, Latvia (Unpublished Doctoral thesis). University of
Latvia. http://www.lu.lv/zinas/t/7361/
Vandenberghe, N., Hilgen, F.J., Speijer, R.P., Ogg, J.G., Gradstein, F.M., Hammer, O.,
Hollis, C.J., Hooker, J.J., 2012. Chapter 28-The Paleogene Period. In: Gradstein, F.M.,
Schmitz, J.G.O.D., Ogg, G.M. (Eds.), The Geologic Time Scale. Elsevier, Boston.
Verneau, O., Bentz, S., Sinnappah, N.D., Preez, L.D., Whittington, I., Combes, C., 2002. A
view of early vertebrate evolution inferred from the phylogeny of polystome parasites
(Monogenea: Polystomatidae). Proc. R. Soc. B Biol Sci. 269, 535e543.
Verneau, O., Du Preez, L., Badets, M., 2009a. Lessons from parasitic flatworms about
evolution and historical biogeography of their vertebrate hosts. C. R. Biol. 332,
149e158.
Verneau, O., Du Preez, L.H., Laurent, V., Raharivololoniaina, L., Glaw, F., Vences, M.,
2009b. The double odyssey of Madagascan polystome flatworms leads to new insights
on the origins of their amphibian hosts. Proc. Biol. Sci. 276, 1575e1583.
Vignon, M., 2011. Putting in shape e towards a unified approach for the taxonomic descrip-
tion of monogenean haptoral hard parts. Syst. Parasitol. 79, 161e174.
Vignon, M., Sasal, P., 2010. The use of geometric morphometrics in understanding shape
variability of sclerotized haptoral structures of monogeneans (Platyhelminthes) with in-
sights into biogeographic variability. Parasitol. Int. 59, 183e191.
Waggoner, B.M., Poinar Jr., G.O., 1993. Fossil habrotrochid rotifers in Dominican amber.
Experientia 49, 354e357.
Walker, S.E., Brett, C.E., 2002. Post-Paleozoic patterns in marine predation: was there a
Mesozoic and Cenozoic marine predatory revolution? Paleontol. Soc. Pap. 8, 119e194.
Wallberg, A., Curini-Galletti, M., Ahmadzadeh, A., Jondelius, U., 2007. Dismissal of Acoe-
lomorpha: Acoela and Nemertodermatida are separate early bilaterian clades. Zool. Scr.
36, 509e523.
Walossek, D., M€ uller, K.J., 1994. Pentastomid parasites from the Lower Paleozoic of
Sweden. Trans. R. Soc. Edinburgh Earth Sci. 85, 1e37.
Walossek, D., Repetski, J.E., Muller, K.J., 1994. An exceptionally preserved parasitic
arthropod, Heymonsicambria taylori n.sp. (Arthropoda incertae sedis: Pentastomida),
from Cambrian-Ordovician boundary beds of Newfoundland, Canada. Can. J. Earth
Sci. 31, 1664e1671.
Waloszek, D., Repetski, J.E., Maas, A., 2005. A new Late Cambrian pentastomid and a re-
view of the relationships of this parasitic group. Trans. R. Soc. Edinburgh Earth Sci. 96,
163e176.
Waltari, E., Hoberg, E.P., Lessa, E.P., Cook, J.A., 2007. Eastward Ho: phylogeographical
perspectives on colonization of hosts and parasites across the Beringian nexus. J. Bio-
geogr. 34, 561e574.
Warnock, R., 2014. Molecular clock calibration. In: Rink, W.J., Thompson, J. (Eds.), Ency-
clopedia of Scientific Dating Methods. Springer, Netherlands.
Constraining the Deep Origin of Parasitic Flatworms and Host-Interactions with Fossil Evidence 135

Warnock, R.C.M., Parham, J.F., Joyce, W.G., Lyson, T.R., Donoghue, P.C.J., 2015. Cali-
bration uncertainty in molecular dating analyses: there is no substitute for the prior eval-
uation of time priors. Proc. R. Soc. B Biol Sci. 282.
Warnock, R.C.M., Yang, Z., Donoghue, P.C.J., 2011. Exploring uncertainty in the calibra-
tion of the molecular clock. Biol. Lett. 8, 156e159.
Warnock, R.C.M., Yang, Z., Donoghue, P.C.J., 2012. Exploring uncertainty in the calibra-
tion of the molecular clock. Biol. Lett. 8, 156e159.
Weber, M., Wey-Fabrizius, A.R., Podsiadlowski, L., Witek, A., Schill, R.O., Sugar, L.,
Herlyn, H., Hankeln, T., 2013. Phylogenetic analyses of endoparasitic Acanthocephala
based on mitochondrial genomes suggest secondary loss of sensory organs. Mol. Phylo-
genet. Evol. 66, 182e189.
Wen-Yuan, Y., De-Xiang, W., Guang-Fang, S., Zhong-Bi, W., Ren-Sheng, T., 1984. Par-
asitologische Untersuchung einer alten Leiche aus der Chu-Dynastie der Streitenden
Reiche aus dem Mazhuan-Grab Nr. 1, Kreis Jiangling, Provinz Hubei. Acta Acad.
Med. Wuhan 4, 23e27.
Wey-Fabrizius, A.R., Podsiadlowski, L., Herlyn, H., Hankeln, T., 2013. Platyzoan mito-
chondrial genomes. Mol. Phylogenet. Evol. 69, 365e375.
Willems, W.R., Wallberg, A., Jondelius, U., Littlewood, D.T.J., Backeljau, T.,
Schockaert, E.R., Artois, T.J., 2006. Filling a gap in the phylogeny of flatworms: rela-
tionships within the Rhabdocoela (Platyhelminthes), inferred from 18S ribosomal
DNA sequences. Zool. Scr. 35, 1e17.
Wood, J.R., Wilmshurst, J.M., Rawlence, N.J., Bonner, K.I., Worthy, T.H., Kinsella, J.M.,
Cooper, A., 2013. A megafauna’s microfauna: gastrointestinal parasites of New Zealand’s
extinct moa (Aves: Dinornithiformes). PLoS One 8, e57315.
Wood, J.R., Wilmshurst, J.M., Wagstaff, S.J., Worthy, T.H., Rawlence, N.J., Cooper, A.,
2012. High-resolution coproecology: using coprolites to reconstruct the habits and hab-
itats of New Zealand’s extinct upland Moa (Megalapteryx didinus). PLoS One 7, e40025.
Xingliang, Z., Reitner, J., 2006. A fresh look at Dickinsonia: removing it from Vendobionta.
Acta Geol. Sin. Engl. Ed. 80, 636e642.
Xylander, W., 2005. Gyrocotylidea (unsegmented tapeworms). In: Xylander, W. (Ed.), Ma-
rine Parasitology. CABI Publishing, Oxon.
Yamaguti, S., 1959. Systema Helminthum. In: The Cestodes of Vertebrates, vol. II. Inter-
science Publishers, New York.
Yamaguti, S., 1963a. Systema Helminthum. In: Monogenea and Aspidocotylea, vol. IV.
Interscience Publishers, New Work.
Yamaguti, S., 1963b. Systema Helminthum. In: Acanthocephala, vol. V. Interscience Pub-
lishers, New York.
Yans, J., Dejax, J., Pons, D., Dupuis, C., Taquet, P., 2005. Implications paléontologiques et
géodynamiques de la datation palynologique des sédiments a facies wealdien de Bernissart
(bassin de Mons, Belgique). C. R. Palevol 4, 135e150.
Yans, J., Dejax, J., Schnyder, J., 2012. On the age of the Bernissart Iguanodons. In:
Godefroit, P. (Ed.), Bernissart Dinosaurs and Early Cretaceous Terrestrial Ecosystems.
Zangerl, R., Case, G.R., 1976. Cobelodus aculeatus (Cope), an anacanthous shark from Penn-
sylvanian black shales of North America. Palaeontogr. Abt. A 154, 107e157.
Zhu, M., Zhao, W., Jia, L., Lu, J., Qiao, T., Qu, Q., 2009. The oldest articulated
osteichthyan reveals mosaic gnathostome characters. Nature 458, 469e474.
Zietara, M., Huyse, T., Lumme, J., Volckaert, F., 2002. Deep divergence among subgenera
of Gyrodactylus inferred from rDNA ITS region. Parasitology 124, 39e52.
Zietara, M.S., Lumme, J., 2002. Speciation by host switch and adaptive radiation in a fish
parasite genus Gyrodactylus (Monogenea, Gyrodactylidae). Evolution 56, 2445e2458.
CHAPTER FOUR

From Fossil Parasitoids to Vectors:


Insects as Parasites and Hosts
Christina Nagler, Joachim T. Haug1
Department of Biology II, Functional Morphology Group, University of Munich (LMU),
Planegg-Martinsried, Germany
1
Corresponding author: E-mail: jhaug@bio.lmu.de

Contents
1. Introduction 139
1.1 Insects as parasites and hosts 139
1.2 Insects in the fossil record 140
2. Insect Parasitism sensu stricto (s. str.) e Paraneoptera 142
2.1 Phthiraptera 142
2.1.1 General aspects 142
2.1.2 Phylogenetic inference of appearance and molecular estimations of early 143
evolution
2.1.3 Fossil representatives 145
2.2 Hemiptera 146
2.2.1 General aspects 146
2.2.2 Phylogenetic inference of appearance and molecular estimations of early 146
evolution
2.2.3 Fossil representatives 146
3. Insect Parasitism s.str. e Antliophora 147
3.1 Siphonaptera 148
3.1.1 General aspects 148
3.1.2 Phylogenetic inference of appearance and molecular estimations of early 148
evolution
3.1.3 Fossil representatives 149
3.2 Diptera 152
3.2.1 General aspects 152
3.2.2 Phylogenetic inference of appearance and molecular estimations of early 153
evolution
3.2.3 Fossil representatives 154
4. Insect Parasitism s.str. e Neuropteroida 159
4.1 Neuroptera (Mantispidae) 159
4.1.1 General aspects 159
4.1.2 Phylogenetic inference of appearance and molecular estimations of early 160
evolution
4.1.3 Fossil representatives 160

Advances in Parasitology, Volume 90


© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.09.003 All rights reserved. 137
138 Christina Nagler and Joachim T. Haug

4.2 Coleopterida (Coleoptera, Meloidae) 161


4.2.1 General aspects 161
4.2.2 Phylogenetic inference of appearance and molecular estimations of early 162
evolution
4.2.3 Fossil representatives 162
5. Parasitoids 163
5.1 Hymenoptera 163
5.1.1 General aspects 163
5.1.2 Phylogenetic inference of appearance and molecular estimations of early 166
evolution
5.1.3 Fossil representatives 167
5.2 Strepsiptera 169
5.2.1 General aspects 169
5.2.2 Phylogenetic inference of appearance and molecular estimations of early 170
evolution
5.2.3 Fossil representatives 171
6. Plant Parasitism (versus Phytophagy) 172
6.1 General aspects 172
6.2 Phylogenetic inference of appearance and molecular estimations of early 173
evolution
6.3 Fossil representatives 173
7. Insects as Hosts 174
7.1 Nematoida 174
7.1.1 General aspects 174
7.1.2 Phylogenetic inference of appearance and molecular estimations of early 174
evolution
7.1.3 Fossil representatives 175
7.2 Mites 175
7.2.1 General aspects 175
7.2.2 Phylogenetic inference of appearance and molecular estimations of early 176
evolution
7.2.3 Fossil representatives 176
7.3 Pseudoscorpions 177
7.3.1 General aspects 177
7.3.2 Phylogenetic inference of appearance and molecular estimations of early 178
evolution
7.3.3 Fossil representatives 179
8. Insects as Vectors 179
8.1 General aspects 179
8.2 Phylogenetic inference of appearance and molecular estimations of early 181
evolution
8.3 Fossil representatives 182
9. Conclusion 182
Insects as Parasites and Hosts 139

10. Outlook 183


Acknowledgements 184
References 184

Abstract
Within Metazoa, it has been proposed that as many as two-thirds of all species are para-
sitic. This propensity towards parasitism is also reflected within insects, where several
lineages independently evolved a parasitic lifestyle. Parasitic behaviour ranges from
parasitic habits in the strict sense, but also includes parasitoid, phoretic or kleptopara-
sitic behaviour. Numerous insects are also the host for other parasitic insects or meta-
zoans. Insects can also serve as vectors for numerous metazoan, protistan, bacterial and
viral diseases. The fossil record can report this behaviour with direct (parasite associated
with its host) or indirect evidence (insect with parasitic larva, isolated parasitic insect,
pathological changes of host). The high abundance of parasitism in the fossil record
of insects can reveal important aspects of parasitic lifestyles in various evolutionary lin-
eages. For a comprehensive view on fossil parasitic insects, we discuss here different
aspects, including phylogenetic systematics, functional morphology and a direct com-
parison of fossil and extant species.

1. INTRODUCTION
1.1 Insects as parasites and hosts
Insecta is often considered to be the largest animal group comprising
over half of all metazoan species (Mayhew, 2007). Also, it has been estimated
that only about 20% of all existing insect species have been described so far
(Grimaldi and Engel, 2005). Considering that over 60% of all insect species
have been proposed to be parasites (Price 1980), we could estimate that
there should be at least 600,000 parasitic insect species in the modern fauna
(although one could ask how reliable these estimations are).
Ectoparasitism has been suggested to have evolved not less than 30 times
in different insect groups (Grimaldi and Engel, 2005). Ectoparasitism and
especially haematophagy occurs, for example, in fleas (Siphonaptera), lice
(Phthiraptera), midges (Diptera), some beetles (Coleoptera) and bugs (Hem-
iptera) (Pe~
nalver and Pérez-de la Fuente 2014, Lukashevich and Mostovki,
2003). Other insects, such as mantid lacewings (Mantispidae) and blister bee-
tles (Meloidae) are endoparasites or kleptoparasites.
While parasites (in the strict sense) do not kill their host, parasitoids do
(Kathirithamby, 2009). Parasitoidism is an important feeding strategy for
different wasps and relatives (Hymenoptera), at least in their larval stages,
140 Christina Nagler and Joachim T. Haug

but also for twisted-wing ‘parasites’ (Strepsiptera) and occasionally also for
other insects. In contrast to most parasites, the size relation of host and para-
sitoid is oddly skewed. While parasites are usually significantly smaller than
their hosts, parasitoids reach about the same size as their host. Often we
also see parasitoids that have parasitoids as host, a phenomenon known as
hyperparasitoidism (Kathirithamby, 2009). Besides parasitoidism, another
important parasitic mode (in the widest sense) is plant parasitism, which
occurs in several insect groups, e.g., hemipterans, coleopterans, dipterans
or hymenopterans.
Finally, insects do not only act as parasites themselves, but many serve as
hosts also for other insects, nematode and nematomorph worms and mites.
They also are hosts to a plethora of bacteria, viruses and protists. Many
notable pathogens, including Trypanosoma, Plasmodium and Dengue virus
use insects as vectors, to become widely distributed (e.g. Poinar, 2014a).

1.2 Insects in the fossil record


The indirect fossil record of insects dates back to the Silurian, 420 million
years ago (ma), based on possible insect coprolites (Labandeira, 2006 and
references therein). Molecular estimations argue for an even earlier origin
(e.g. Misof et al., 2014). Yet, no definitive insect body fossil has been found
that old. The oldest body fossil that was interpreted as an insect is a single
specimen of a supposed insect nymph from the Late Devonian, ca. 425 ma
(Garrouste et al., 2012). However, H€ ornschemeyer et al. (2013) rejected
this interpretation and interpreted it as ‘poorly preserved Devonian
arthropod’ based on a reinvestigation of the specimen. Other putative
Devonian insects tend to be rather fragmentary (e.g. Engel and Grimaldi,
2004; 400 ma).
Definitive and more complete insect body fossils have been described
from the Carboniferous (350 ma) and they already appear quite diversified
at that time (e.g. Shear and Kukalova-Peck, 1990). Many major modern in-
sect groups make their appearance later in the Triassic, 250 ma and many
minor modern extant insect groups (‘families’) find their origins in the
Cretaceous, 120 ma (Grimaldi and Engel, 2005).
The oldest fossil evidence for arthropod parasitism in general is
provided by pentastomids from the Cambrian (495 ma; Castellani et al.,
2011 and references therein). However, while pentastomids are arthropods
they are not insects. Parasitism in insects originated several times
(Labandeira, 2002) and existed definitely since the Triassic (250 ma; see
below).
Insects as Parasites and Hosts 141

Parasitized and parasitic insects are preserved as fossils in rocks and in


numerous occasions in amber. Amber, fossilized tree resin, preserves soft-
bodied specimens as well as organisms with a cuticle, to an exceptionally
detailed degree (Pohl et al., 2010; Lewis and Grimaldi, 1997) providing
unique snapshots of life as it was millions of years ago.
We can roughly distinguish four different types of fossil, which allow
parasitism to be inferred in the fossil record:
1. The most direct cases are parasites directly associated with their hosts.
Typical examples are mites on insects, or nematodes emerging from an
insect (see below). Inclusions in amber are perfect examples of this
kind of preservation (e.g. Poinar, 2014a).
2. Also quite direct is finding isolated parasites, i.e., parasites without their
hosts. In such cases, a morphological comparison to extant relatives can
give clues; a fossil flea will be as parasitic as a modern one. In cases where
we have no closely related extant form, a functional morphological com-
parison to modern parasitic forms can still be a strong indicator for a parasitic
lifestyle e presumably because their morphology is known to be modified
directly as a result of their parasitic lifestyle. More rarely, the analysis of the
contents of the digestive or circulatory system might also be informative.
3. More indirect evidence for parasitism, applicable to many insect groups,
is the finding of free-living developmental stages of a parasite. This type
of evidence relies heavily on comparison with modern forms.
4. Another type of indirect evidence is the influence on morphology,
growth and development of the host as a reaction to parasitism. This is
especially telling (but not only) in cases of plant parasitism indicated by
galls, leaf mines and asymmetric growth. However, it can be difficult
to identify the responsible organisms for these pathologies. Also here
inference relies heavily on comparisons with extant parasiteehost reac-
tions (e.g. McNaughton, 1983).
In the following, we provide examples of fossil insect parasitism in various
insect groups. Parasitism is meant here in its broadest sense, parasitism sensu
lato. This includes, for example, parasitism sensu stricto (traditional type of
parasite), parasitoidism (host about the same size and is killed) plant parasitism
(plant is ‘forced’ to help the insect) and kleptoparasitism (a parasite taking
prey or other useful material from the host; Rozen, 2003; Sivinski et al.,
1999). Furthermore, we consider cases where insects are hosts or vectors for
other organisms. Due to the enormous diversity of insects, we will not be
able to cover every possible fossil example insects involved in parasitic re-
lationships, but we tried our best to cover all major fossil examples.
142 Christina Nagler and Joachim T. Haug

2. INSECT PARASITISM SENSU STRICTO (S. STR.) e


PARANEOPTERA
Not all insect groups are known to have parasitic representatives. As
far as we could find, there seem to be no unequivocal parasitic representa-
tives of Entognatha, Archaeognatha, Zygentoma, Odonatoptera, Ephemer-
optera, Palaeodictyoptera or Polyneoptera. Hence active parasitism appears
to be restricted to eumetabolan insects (Paraneoptera þ Holometabola).
Paraneoptera includes the groups Phthiraptera (chewing lice and sucking
lice), Psocoptera (book lice, bark lice), Hemiptera and Thysanoptera.
Phthiraptera is most likely an in-group of Psocoptera (e.g. Beutel et al.,
2014), traditionally both have been placed together in the group Psocodea.
Psocodea or Psocoptera (if simply accepting that Phthiraptera are psocopter-
ans) is sister group to Hemiptera plus Thysanoptera, both being again sister
groups (Light et al., 2010; Johnson and Clayton, 2003). Active parasitism
occurs in Phthiraptera, which are obligate ectoparasites of birds and mam-
mals, and also in some Hemiptera.

2.1 Phthiraptera
2.1.1 General aspects
The group Phthiraptera, or true lice, currently comprises 4900 extant species.
They are traditionally split into two subgroups, ‘Mallophaga’, the chewing
lice, and Anoplura, the sucking lice (Grimaldi and Engel, 2005). The chewing
lice are named for their chewing mouthparts and include Amblycera, Ischno-
cera and Rhynchophthirina (Johnson and Clayton, 2003).
All phthirapterans are obligate ectoparasites of birds or mammals. They
are extremely modified for their parasitic lifestyle and behaviourally adapted
for particular microhabitats on the hosts (Johnson and Clayton, 2003). All
stages (including the adults) are wingless and dorsoventrally flattened.
They posses claws for clinging to hair or feathers and have modified
mouthparts for feeding (Light et al., 2010). The true lice are highly host-
and site-specific and spend their entire life on the host (as the only group
of ectoparasitic arthropods except mites). The females lay large eggs and atta-
ch them to the hair or feathers of their host with a special secretion (Martill
and Davis, 1998). Due to the dependence of their life cycle to the host and
their host specificity, Phthiraptera is a candidate for being the most special-
ized group of parasites among insects. They are a model system for phyloge-
netic studies including co-speciation and coevolution (Wappler et al., 2004;
Johnson and Clayton, 2003).
Insects as Parasites and Hosts 143

After hatching, phthirapterans develop through three nymphal stages to


the adult. The lice feed on keratin in hair or feathers, secretions or blood.
Their ancestors most likely had simple chewing mouthparts and were free
living in the nests or burrows of vertebrates. Later in their evolutionary his-
tory, they apparently became more dependent on their hosts and adapted
from associates to parasites. They fed directly from their hosts and developed
iteratively more modified mouthparts, dependent on where on the hosts
they lived and their chosen diets (Light et al., 2010).
Amblycera (traditionally classified as ‘Mallophaga’) is the sister group to
all remaining phthirapterans. They are obligate parasites mostly of birds, but
about 12% parasitize mammals such as rodents (Grimaldi and Engel, 2005;
Johnson and Clayton, 2003).
The group Ischnocera is the largest group of ‘Mallophaga’ with approx-
imately 3080 species. About 2700 feed on birds and 380 on mammals
(Yoshizawa and Lienhard, 2010).
Rhynchophthirina includes only three species; they parasitize elephants
and wild pigs. All of them possess an elongated ‘rostrum’ which bears the
mandibles in a 180 angle from the ‘normal’ position to break through
the skin and feed on blood. They are the sister group to Anoplura and it
is likely that Anoplura evolved from a Rhynchophthirina-like ancestor
(Light et al., 2010; Johnson and Clayton, 2003).
Anoplurans, the sucking lice, parasitize only mammals. The mouthparts
and the head of the 550 extant species are extremely modified and adapted
to the mammalian host.

2.1.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Liposcelidae, traditionally classified as ‘Psocoptera’ or book lice are the sister
group to Phthiraptera (Yoshizawa and Johnson 2010). Due to a discovery of
a representative of Liposcelidae in mid-Cretaceous amber of Myanmar
(about 100 ma), the minimal divergence time of true lice and Liposcelidae
is 100 ma. Cretoscelis burmitica (Figure 1(a)) from amber of Myanmar
(Grimaldi and Engel, 2006), although not parasitic itself, provides an impor-
tant calibration point for the appearance of true lice.
Recent findings of troctomorphs, another in-group of Psocoptera, show
a minimal age for the entire group Psocodea to be 135 ma (Early Creta-
ceous; Azar et al., 2015a). Grimaldi and Engel (2005) estimated the evolu-
tionary appearance of lice to be around 145 ma, where the host of lice could
be an early mammal, bird, another feathered theropod dinosaur or a haired
144 Christina Nagler and Joachim T. Haug

Figure 1 Fossil representatives of Psocodea (Paraneoptera). (a) Cretoscelis burmitica


(Liposcelidae, Psocodea) (modified from Grimaldi and Engel (2006, Figure 2) with kind
permission of the Royal Society). (b) Saurodectes vranskyi (modified from Grimaldi and
Engel (2005, Figure 38.12) with kind permission of Cambridge University Press and David
Grimaldi). (c) Megamenopon rasnitsyni (Phthiraptera, Psocodea) (modified from the fossil
pictured in Wappler et al. (2004), Figure 1(a) with kind permission of the Royal Society). (d)
Psittacobrosus bechsteini (Phthiraptera, Psocodea) (modified from Mey (2005), Figure 1
with kind permission of Eberhard Mey).

pterosaur. Based on molecular dating techniques, Smith et al. (2011) sup-


ported the notion that the early hosts of lice were feathered theropod dino-
saurs. The appearance of Amblycera (the earliest offshoot of Phthiraptera)
has been suggested to be around 120 ma, with the ancestral host again being
an early feathered dinosaur (Wappler et al., 2004).
Molecular data indicate a slightly older origin of lice, at least in the Early
Cretaceous (Yoshizawa and Johnson 2003), which would be more or less
Insects as Parasites and Hosts 145

congruent with the fossil record. Yet, also even earlier origins have been
suggested (e.g. Martill and Davis, 1998).
Molecular data support an evolutionary scenario from free-living detri-
tivorous generalistic psocodeans over liposcelid-like forms to highly special-
ized obligate ectoparasitic phthirapterans (Yoshizawa and Johnson, 2003).
Representatives of Liposcelidae have been found in nests of birds, mammals
or in the plumage of birds and fur of mammals, feeding on faeces, shed fur
and feathers (Grimaldi and Engel, 2006). This association may have given
the input to a permanent parasitism in lice (Yoshizawa and Lienhard, 2010).
Based on molecular data, Anoplura supposedly appeared in the Late
Cretaceous (77 ma) and soon after the CretaceousePalaeogene boundary
(before 65 ma), they radiated and adapted to different mammalian in-groups
(Smith et al., 2011; Light et al., 2010).

2.1.3 Fossil representatives


Due to the close connection between these ectoparasites and their hosts, fossils
of Phthiraptera are rare. To date, six possible cases have been discovered: Voigt
(1952) reported louse eggs on hair in Baltic amber, which has been recognized
by various authors. Grimaldi and Engel (2005) and Wappler et al. (2004)
considered this discovery has very little systematic value, but Mey (2005)
emphasized its importance and suggested a reinvestigation of the material.
Rasnitsyn and Zherikhin (1999) described Saurodectes vrsanskyi (Figure
1(b)) from the Zaza Formation of Baissa (130 ma), and interpreted it as an
ischnoceran louse. Yet, its systematic attribution was based on an elimination
process of other insect groups. Grimaldi and Engel (2005) stated that it is an
‘ectoparasite with phthirapteran affinities’ and later (Grimaldi and Engel,
2006) as a ‘putative louse or close relative, but also an exceedingly bizarre
insect’. Mey (2005) followed this statement. Dalgleish et al. (2006) proposed
that Saurodectes vrsanskyi is neither a phthirapteran nor could be assigned to
any other insect ‘order’.
Cuticular remains from the Triassic of Saptura Badin (215 ma) were
described as a louse by Kumar and Kumar (2001). Subsequent authors inter-
preted these remains as that of an oribatid mite (Dalgleish et al., 2006; Mey,
2005; Wappler et al., 2004).
The single specimen of Amblyceropsis indica from the Triassic Bagra For-
mation (Kumar, 2001) was also described as a putative louse. Yet, it is likely
also a mite, probably a representative of Prostigmata (Dalgleish et al., 2006;
Mey, 2005; Wappler et al., 2004).
Megamenopon rasnitsyni is 44 million years old from the Messel Lagerst€atte
in Germany and resembles modern menoponid lice (Figure 1(c)). The
146 Christina Nagler and Joachim T. Haug

specimen is likely an ectoparasite of aquatic birds, as in the gut of this para-


sitic louse remains of feathers have been found (Wappler et al., 2004). The
newest discovery was Psittacobrosus bechsteini by Mey (2005). It seems now
extinct, but is not really a fossil (Figure 1(d)); its host is a specimen of the
bird Ara tricolor found in the 1980s.
To conclude: only the last two specimens, of M. rasnitsyni and P. bech-
steini, can be assigned to Phthiraptera without any doubts. There is hence
only one direct phthirapteran fossil, namely, M. rasnitsyni with its last meal
in it. The phthirapteran eggs should be reinvestigated to clarify their system-
atic position.

2.2 Hemiptera
2.2.1 General aspects
Hemiptera includes Fulgoromorpha, Cicadomorpha (together forming
Auchenorrhyncha), Coleorrhyncha, Sternorrhyncha (all four groups
together formerly referred to as ‘Homoptera’) and Heteroptera (Beutel
et al., 2014). A prominent haematophagous group within Hemiptera is
Reduviidae; some species are commonly known as kissing bugs (Grimaldi
and Engel, 2005). Two other groups, Cimicidae and Polyctenidae, contain
also several haematophagous species (Yao et al., 2014). The exact relation-
ship of the hemipteran subgroups is still under discussion (Beutel et al., 2014;
Li et al., 2012).

2.2.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Their high diversity is often explained as having diverged with the radiation
of angiosperms coincident possibly with their phytophagous behaviour.
Phylogenetic analyses indicate that the origin of Hemiptera lies in the
Late Permian (Li et al., 2012). They have been considered to have evolved
by means of a phytophagous strategy giving way to arthropod predation and
finally becoming haematophagous, around 110e32 ma (Otalora-Luna et al.,
2015). The haematophagy should have evolved at least three times in these
bugs (Ribeiro et al., 2012). A recent finding extends the record of haema-
tophagous hemipterans to the Early Cretaceous (145 ma) (Yao et al., 2014).

2.2.3 Fossil representatives


There is one questionable compression fossil ascribed to Reduviidae from
the Early Cretaceous (140e120 ma) and some reduviids were discovered
in Burmese (Early Cretaceous), Canadian (Late Cretaceous) and Dominican
Insects as Parasites and Hosts 147

(Miocene) amber (Poinar and Poinar, 2005; Poinar, 1992). Remarkable


fossil representatives from the Early Cretaceous in China (145 ma) have
been described by Yao et al. (2014). They provided evidence for haema-
tophagy on mammals, birds or avian-rafted dinosaurs with the aid of
geochemical data.

3. INSECT PARASITISM S.STR. e ANTLIOPHORA


The exact relationships within Holometabola have been in flux. Yet,
the current (more or less stable) hypothesis suggests that wasps, ants and bees
(Hymenoptera) are the sister group to all remaining holometabolans. These
form two sister groups Neuropteroidea and Mecopteroidea (based on larval
morphological characters and on molecular data, Peters et al., 2014;
Labandeira, 2011). Active parasites (in the strict sense) are known in both
of the latter groups.
The monophyletic group Mecopteroidea (Panorpida of some authors)
consists of Mecoptera, Diptera (together Antliophora) and Trichoptera
and Lepidoptera (together Amphiesmenoptera; Beutel et al., 2014; Peters
et al., 2014). The ancestor of Mecopteroidea most likely existed in the
Permian already.
Within Antliophora, Diptera is the sister group to Mecoptera (Peters
et al., 2014; Beutel et al., 2011; Wiegmann et al., 2009). Siphonaptera (fleas),
traditionally considered to be the sister group of Mecoptera, has been repeat-
edly resolved as an in-group of Mecoptera (e.g. Whiting, 2002, and below).
Antliophorans are characterized by an elongation of the mouthparts.
This appears to be part of their ground pattern. This might be seen as a
possible preadaptation to a blood-sucking (haematophagous) type of ecto-
parasitism. As will be outlined in detail below, haematophagy is a ground
pattern feature for dipterans and fleas, but also some mecopterids appear
to have fed on blood. While most modern mecopterids feed on flowers,
this source of food was not yet readily available when the group diversified.
Hence, we might even speculate that the elongate mouthparts in the stem
species of Mecopteroidea were already used for blood sucking and ectopar-
asitism could represent an autapomorphy of Mecopteroidea. Mecopterans
feeding on flowers could then represent a novelty, similar to evolutionary
changes in Diptera (see below for details).
Most mecopterans are insects with extremely elongated mouthparts and
an elongate body. Larvae are generally scavengers and adults feed on nectar
148 Christina Nagler and Joachim T. Haug

(Poinar, 2012b). Certain scorpionflies from the Early Cretaceous (100 ma)
have long pointed mouthparts with fine serrations, this indicates that these
early mecopterans fed on blood (Boucot and Poinar, 2010).
Some fossils of the Late Jurassic (150e140 ma) indicate a Mesozoic age
for the sister group to the fleas (Grimaldi and Engel, 2005) and that this
type of fossil scorpionflies may have led to Siphonaptera (Huang et al.,
2012; Poinar, 2012b). It has also been suggested that Siphonaptera is either
an in-group (Friedrich and Beutel 2010) or the sister group (Whiting et al.,
2008) to the mecopteran in-group Boreidae. Despite the fact the exact
origin and relationships of fleas and other scorpionflies remains partly un-
clear, we can still state that with the radiation of mammals in the Palaeocene
and the obligate ectoparasitism of fleas, Siphonaptera radiated far more
recently than other mecopteran groups (Grimaldi and Engel, 2005).

3.1 Siphonaptera
3.1.1 General aspects
Siphonaptera, the true fleas, are laterally compressed, wingless obligate ecto-
parasites. They feed exclusively on the blood of mammals (94%) and birds
(6%) (Poinar, 2012b). About 2500 living species are known and they are
evolutionarily successful as they are a long-lived lineage that has radiated suc-
cessfully across multiple hosts (Poulin and Morand, 2000). Like many other
insect parasites, fleas have reduced eyes, reduced antennae, no wings, large
claws and modified mouthparts to pierce the skin of the host and suck blood
(Lukashevich and Mostovki, 2003; Rasnitsyn and Quicke, 2002). Interest-
ingly all insect ectoparasites have a dorsoventrally flattened body, except
fleas, where it is laterally compressed. In contrast to lice, fleas do not generally
spend their whole life on the host. Larvae often live in or near the nest of the
host and feed on host faeces, exfoliated skin; others are predators or even can-
nibals (Grimaldi and Engel, 2005; Rasnitsyn and Quicke, 2002).

3.1.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Based on molecular studies, Siphonaptera is considered to be monophyletic;
Tungidae, with its species retaining numerous plesiomorphic characters, likely
represents the sister group to all other fleas (Beutel et al., 2014; Whiting et al.,
1997, 2008). It is still controversial as to which fossils are indeed all part of the
early evolutionary history of fleas. The systematic and phylogenetic position
of some Mesozoic fossils is still not settled (Gao et al., 2013a; Perrichot et al.,
2012; Poinar, 2012b; Vransky et al., 2010; see below).
Insects as Parasites and Hosts 149

Flea-like ancestors have been considered to have shifted step by step


from pterosaurs to mammalian hosts (Gao et al., 2012; Huang et al.,
2012; Poinar, 2012b). Mesozoic flea-like fossils indicate a possible diversifi-
cation of Siphonaptera at least in the Late Jurassic (150 ma) (Gao et al.,
2013a; Huang et al., 2012).

3.1.3 Fossil representatives


Due to the close connection between siphonapterans and their hosts, fleas
are rare in the fossil record just like lice (Lukashevich and Mostovki, 2003).
• Extinct adult fossil specimens of Siphonaptera (Figure 2(b)) ascribed to
extant genera have been found in Eocene Baltic amber and in Miocene
Dominican amber (40e50 ma) (Perrichot et al., 2012; Poinar, 2012b;
Lukashevich and Mostovki, 2003; Lewis and Grimaldi, 1997). These
species found in amber indicate the evolution for modern species-groups
in the Middle Caenozoic (Perrichot et al., 2012).
While such clear in-group fossils are easy to recognize as such, there are also
some more challenging Mesozoic findings:
• Strashila incredibilis (Figure 2(f )) from the Jurassic of Transbaikalia in East
Siberia (Rasnitsyn, 1992). It has long, slender legs to grasp its host, but is
thought to have been unable to walk on the ground, which would indi-
cate a permanent life on the host (Grimaldi and Engel, 2005). Rasnitsyn
and Quicke (2002) supposed S. incredibilis as possible sister group to Tar-
winia and Saurophthirus (see next points). However, recent morphological
study indicates that Strashilidae are highly specialized dipterans (flies)
(Huang et al., 2013).
• Saurophthirus longipes (Figure 2(e)) from the Early Cretaceous Zaza For-
mation of Baissa, central Siberia (Ponomarenko, 1976). This large ecto-
parasitic insect with long, slender legs closely resembles certain Mesozoic
giant fleas (Huang et al., 2012; see below). Some extant ectoparasitic bat
flies show similar body proportions and legs (Grimaldi and Engel, 2005;
Lukashevich and Mostovski, 2003). Rasnitsyn and Quicke (2002) pro-
posed Saurophthirus as sister group to Tarwinia and true fleas. Gao et al.
(2013a) assigned S. longipes to the ‘family’ Saurophthiridae (representing
an empty, useless bracket) and supposed it to represent the closest relative
to modern fleas.
• A specimen of Tarwinia australis (Figure 2(c)) and three other (similar
appearing) specimens from the Early Cretaceous Koonwarra sediments
of Victoria (Jell and Duncan, 1986). These are thought (by Grimaldi
and Engel, 2005) to be ‘the early close relative of fleas from the
150 Christina Nagler and Joachim T. Haug

Figure 2 Fossil representatives of Mecopterida (Mecopteroidea). (a) Mosquito-like scor-


pionfly (Boreidae, Mecoptera) (modified from Grimaldi and Engel (2005, Figure 12.3) with
kind permission of Cambridge University Press and David Grimaldi). (b) Eospilopsyllus kob-
berti (Siphonaptera, Mecopterida) (modified from Perrichot et al. (2012, Figure 1) with
kind permission of Magnolia Press). (c) Tarwinia australis (Siphonaptera, Mecopterida)
(modified with permission from Elsevier from Huang et al. (2014, Figure 4)). (d) Pseudopu-
licidae sp. (Pseudopulicidae, Siphonaptera) (modified with permission from Macmillan
Publishers Ltd from Huang et al. (2012, Figure 2(a))). (e) Saurophthirus longipes (Sauroph-
thiridae, Siphonaptera) (modified from Rasnitsyn (1992, Figure 5) with kind permission of
Alexander Rasnitsyn). (f) Strashila incredibilis (Siphonaptera, Mecopterida) (modified from
Rasnitsyn (1992, Figure 2) with kind permission of Alexander Rasnitsyn).

Mesozoic’. After a current reinterpretation of the holotype, T. australis is


a definitive early representative of the lineage to Siphonaptera, and has
some particular features that are closely related to another Mesozoic giant
flea group, Pseudopulicidae (see below; Huang, 2015).
Insects as Parasites and Hosts 151

• Mesozoic giant fleas: two species (Figure 2(d)) from the Middle Jurassic
(165 ma) and one species from the Early Cretaceous (125 ma) in China.
These specimens share morphological characters with extant fleas, but
retain plesiomorphic features like nonjumping hind legs. Also they
have similar mouthparts in siphonate mecopterids. These giant fleas
might document an early diversification towards haematophagy (Huang
et al., 2012), assuming this is not a plesiomorphic feature. All three spe-
cies were ascribed subsequently to Pseudopulicidae (see below; Huang
et al., 2013)
• Pseudopulex jurassicus from the Middle Jurassic and Pseudopulex magnus from
the Early Cretaceous (Gao et al., 2012). These specimens share some
morphological characters with extant fleas, but could represent an extinct
lineage in early flea evolution (Gao et al., 2012). The morphological data
indicate that Pseudopulicidae (including these two) represent an early
offshoot of the siphonapteran lineage (Gao et al., 2012). Subsequently
Gao et al. (2013a) suggested Tarwinia þ Pseudopulicidae as sister group
to Saurophthirus þ the modern fleas based on morphological characters.
• Saurophthiridae. With Saurophthirus exquisitus from the Lower Creta-
ceous Yixian Formation in China et al. (Gao et al., 2013a). This species
has short and slender piercingesucking stylet mouthparts. Hence Sau-
rophthiridae (including another species S. longipes see above) resembles
the true fleas more than any other Mesozoic flea or flea-like fossil.
This indicates a divergence of Saurophthiridae and true fleas in the
Late Jurassic (Gao et al., 2013a). However, Grimaldi and Engel (2005)
saw no features that indicate special relationships to true fleas, but may
be because at the point of their review the only known Saurophthirus
was S. longipes, where the mouthparts were not preserved.
All flea-like Mesozoic fossils are definitely not true fleas in the strict sense
(Poinar, 2012b), but are more likely early offshoots of the evolutionary line-
age towards them (some may also represent other groups, see above). One
could simply consider modern fleas (including most amber fossils) as Siphon-
aptera sensu stricto, while using Siphonaptera sensu lato for the Mesozoic flea-
like fossils plus Siphonaptera s.str., instead of discussing at which point the
term ‘flea’ should be applied. Hence, we can state that early representatives
of Siphonaptera sensu lato (before the node of Siphonapetra sensu stricto) were
likely parasitizing feathered dinosaurs, pterosaurs or early mammals as ecto-
parasites grasping with their long legs (Gao et al., 2013a; Rasnitsyn, 1992;
Ponomarenko, 1976; Jell and Duncan, 1986). Their habitus indicates a strict
ectoparasitic lifestyle, adapted more for attachment than jumping (as in
152 Christina Nagler and Joachim T. Haug

modern forms). The early representatives of Siphonaptera sensu lato appear


not to have entangled themselves in hair or feathers. They could have either
pierced through scales or fed on the skin between the scales on nonfeathered
dinosaurs (Poinar, 2012b) or they could have parasitized the membrane of
the wings of pterosaurs (Gao et al., 2013a, 2012; Huang et al., 2012; Vransky
et al., 2010).

3.2 Diptera
3.2.1 General aspects
The superdiverse insect group Diptera is cosmopolitan and with 154,000
described species represents 10e12% of all described animal species
(Lambkin et al., 2013). Dipterans are easily recognizable as their hind wings
are reduced to halteres.
The ancestral feeding mode for Diptera is blood-sucking. Females of
groups which retain this mode have mandibles, which function as lancets
(Grimaldi and Engel, 2005). Haematophagy is obligate for females (of
blood-sucking species) to complete each gonadotrophic cycle and produce
fertile eggs (Greenwalt et al., 2013). Early divergent representatives of
Diptera, like midges and craneflies, resemble mecopterans in basic
morphology, with their long wings and long, dangling legs. Haematophagous
groups within the Diptera are numerous: Phlebotominae, Sycorinae, Culici-
dae, Corethrellidae, Ceratopogonidae, Simuliidae, Rhagionidae, Tabanidae,
Carnidae, Muscidae and Glossinidae (Grimaldi and Engel, 2005).
Extant adults of Chironomidae feed on nectar and honeydew, but early
representatives of Chironomidae from the Triassic and Jurassic were still
haematophagous. Such forms had biting mandibles, like Aenne triassica
from the Late Triassic of England (210 ma; Azar and Nel, 2012).
Extant representatives of Culicidae are blood suckers of almost all verte-
brates. Hence, likely fossil hosts could have been early mammals, lizards,
snakes, turtles, crocodilians, but also nonavian dinosaurs (Poinar et al., 2000).
Species of Corethrellidae are known as frog-biting midges. These tiny
midges are host-specific and parasitize frogs. 105 extant and 7 fossil species
are known (Borkent, 2014). Female frog-biting midges find their hosts
by listening to the mating calls of the host frogs (De Silva and Bernal, 2013).
Representatives of Ceratopogonidae or biting midges are small, but have
well-developed wings. Representatives of older lineages feed on vertebrates,
while more derived species feed on insects as ectoparasites (Borkent et al.,
2013; Pérez-de la Fuente et al., 2011). Ceratopogonidae is very species
rich, with 6180 extant species (Borkent, 2014; Greenwalt et al., 2013).
Insects as Parasites and Hosts 153

Simuliidae or blackflies are a group with 2100 extant species (Craig et al.,
2012). The claws of fossil forms indicate that early representatives of
Simuliidae fed on feathered nonavian dinosaurs or early birds (Currie and
Grimaldi, 2000).
Psychodidae with about 3000 extant species has received special scienti-
fic attention due to their medical significance. Within Psychodidae,
representatives of Sycorinae and Phlebotominae are parasitic blood feeders
(Grimaldi and Engel, 2005).
The parasitic groups Tabanidae and Rhagionidae are in-groups of
Tabanomorpha (Kerr, 2010). Species of Rhagionidae are known as water
snipe flies and are a relatively small group with about 500 extant species
(Zhang, 2012). About 50 extant species are parasitic on birds, mammals or
amphibians, mainly species of Symphoromyia (Zhang, 2012). In other groups,
plesiomorphic are parasitic while derived forms have abandoned this strat-
egy. An exception might be Rhagionidae, which is thought to have evolved
a predatory lifestyle, while parasitic forms evolved de novo from this strategy
(Grimaldi and Engel, 2005).
The elongate mouthparts of the 4500 species of Tabanidae (Zhang,
2012) are well suited for blood- and nectar-feeding at the same time (Karoyli
et al., 2014) and they have been considered to be the first pollinators of early
angiosperms (Labandeira, 2010).

3.2.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Dipterans are thought to have evolved in the latest Permian to the earliest
Triassic, 247 ma, when other insect lineages became extinct. After the
TriassiceJurassic mass extinction event (about 200 ma), early brachycerans
(e.g. Tabanidae) radiated and new dipteran larval types evolved. These
larvae lived in soil, likely due to the drying of aquatic and semiaquatic
habitats. Hence brachycerans became one of the most dominant insect
groups. In the Early Cretaceous (145 ma), the more derived brachyceran
in-groups, Cyclorrhapha and Schizophora, radiated (Lambkin et al.,
2013; Wiegmann et al., 2011; Reidenbach et al., 2009). In this period,
dipteran in-groups evolved all kind of new diets: phytophagy, coprophagy,
necrophagy, larval or adult predators, leaf mining and different types of
parasitism (Grimaldi and Engel, 2005; Rasnitsyn and Quicke, 2002).
Hence while blood sucking and haematophagy seem to be the ancient
diet of dipterans, this type of diet must have been lost (but also regained)
independently several times. Most often it was displaced by nectar feeding
154 Christina Nagler and Joachim T. Haug

multiple times during the radiation of angiosperms in the Lower


Cretaceous (Choufani et al., 2013; Greenwalt et al., 2013; Azar and Nel,
2012; Krenn and Asp€ ock, 2012; Zhang, 2012).
Modern dipterans do not use their mandibles to chew. It is supposed that
the first ‘non-mandibulate’ forms occurred after the Lower Cretaceous (Azar
and Nel, 2012). Yet, due to the less patchy geographic distribution of
culicids they have been considered to have existed before the separation
of Pangaea, 300e200 ma (Harbach and Greenwalt, 2012).
Ceratopogonidae þ Chironomidae is the sister group to Simuliidae þ
Thaumaleidae. Fossils ascribed to Simuliidae (176 ma) and Thaumaleidae
(146 ma) indicate that biting midges (Ceratopogonidae) should have evolved
at the latest in the Late Jurassic (Beutel et al., 2014; Borkent et al., 2013; Chou-
fani et al., 2013). The earliest direct fossil evidence comes from the Lebanese
amber (122 ma; Borkent et al., 2013; Pérez-de la Fuente et al., 2011).
Psychodidae has been thought to have evolved by the Triassic (248 ma)
before the breakup of Pangaea (Andrade Filho et al., 2009, 2007; Azar and
Nel, 2003). The protozoan Leishmania is supposed to have coevolved with psy-
chodidans (Andrade Filho and Brazil, 2003). It is likely that psychodidans
evolved from a blood-sucking form parasitic on nonavian dinosaurs or ptero-
saurs gradually to a nectar- and pollen-eating form, since only the old lineages
Sycoracinae and Phlebotominae are still parasitic (Azar and Nel, 2003).
Tabanomorph-like forms are thought to be the oldest representatives of
Brachycera and radiated probably 200 ma (Lambkin et al., 2013; Wiegmann
et al., 2011). Tabanidae (Figure 3(a)) are considered to have evolved quite
recently in the Early and Middle Cretaceous (Beutel et al., 2014).
Culicomorpha with Nymphomyiidae, Culicoidea and Chironomoidea
are thought to represent early radiations (Lambkin et al., 2013). This
interpretation is more or less congruent with a recent molecular study
(Wiegmann et al., 2011). Culicidae (mosquitoes) should have evolved
during the Middle Jurassic, 191 ma (Reidenbach et al., 2009), based on
molecular data. Still all these inferences are dependent on certain relation-
ship assumptions and the exact relationships within Diptera are still not satis-
fyingly clarified (Beutel et al., 2014).

3.2.3 Fossil representatives


Dipteran fossils are numerous especially in amber. A number of examples
include the following:
• The oldest representative of Diptera has been considered to be Grauvogelia
arzvilleriana from the Middle Triassic Voltzia-Buntsandstein (240 ma) (Gall
Insects as Parasites and Hosts 155

Figure 3 Fossil representatives of Diptera (Mecopteroidea) (a and b); Mantispidae


(Neuroptera, Neuropteroida) (c and e); Meloidae (Coleoptera, Coleopterida) (f). (a)
Laiyangitabanus formosus (Tabanidae, Diptera) (modified with permission from Elsevier
from Zhang (2012, Figure 2(a))). (b) Burmaculex antiquus (Culicidae, Diptera) (modified
from Grimaldi et al. (2002, Figure 36(b)) with kind permission of the American Museum
of Natural History). (c) Dicromantispa electromexicana (Mantispidae, Neuroptera)
(modified from Engel and Grimaldi (2007, Figure 5) with kind permission of the American
Museum of Natural History). Body length 10.4 mm. (d) Micromantispa cristata
(Mantispidae, Neuroptera) (modified with permission from Elsevier from Shih et al.
(2015, Figure 5(a))). (e) Mantispa styriaca (Mantispinae, Mantispidae) larva (modified
from Ohl (2011, Figure 2(b)) with kind permission from Springer Science and Business
Media). (f) Epispasta abbreviata (Meloidae, Coleoptera) larva (modified with permission
from Elsevier from Bologna et al. (2008, Figure 1(f))).
156 Christina Nagler and Joachim T. Haug

and Grauvogel, 1966). As haematophagy is thought to be ancestral for


Diptera, we might assume a similar mode of life for this species.
• Only slightly younger is the fossil of a nematoceran (mosquito-like form)
in 230 ma Triassic amber from Italy (Schmidt et al., 2012).
• Specimens of dipterans are the most common and most diverse forms in
amber inclusions; in Spanish amber dipterans are in 38% of all biological
inclusions (Delclos et al., 2007; Grimaldi et al., 2002). Due to their abun-
dance in lake sediments and their swarm behaviour, dipterans in general
are very frequent in the fossil record (Grimaldi and Engel, 2005).
• Fossil chironomids are abundant not only as inclusions in amber but also
as impressions from the Triassic to the Cretaceous (Lukashevich and
Przhiboro, 2015; Azar and Nel, 2010; Azar et al., 2008; Jarzembowski
et al., 2008; Veltz et al., 2007; Ansorge, 1996; Kalugina and Kovalev,
1985; Kalugina, 1974, 1976, 1980, 1993). Yet, only Triassic and Jurassic
forms appear to have been haematophagous, while the younger forms
were most likely already feeding on nectar and other flower products.
• Culicidae (mosquitoes) is represented in the fossil record by at least 25
species; 11 are known in amber, 14 are preserved as compression or
impression fossils (Hulden and Hulden, 2014; Greenwalt et al., 2013).
Due to their larvae, which have an obligate connection to water, most
fossil mosquitoes are preserved in lake or other lacustrine sediments
(Briggs, 2013). The fossils range from the Middle Cretaceous to the
Oligocene. Some Eocene fossils have been ascribed to modern mosquito
species (Harbach and Greenwalt, 2012; Poinar et al., 2000; Grimaldi
et al., 2002; Briggs, 2013). The oldest fossil mosquito Burmaculex antiquus
(Figure 3(b)) from Burmese amber (100 ma) has been proposed to repre-
sent the sister group to all remaining culicids. This is based on the reten-
tion of several plesiomorphic characters that show an intermediate
condition between modern mosquitoes and other midges (Harbach
and Greenwalt, 2012; Borkent and Grimaldi, 2004). Grimaldi et al.
(2002) stated that the gut of this specimen contains granular material,
possibly representing a blood meal. This statement has not been further
supported.
• One of the most spectacular findings is a modern-type mosquito, Culiseta
sp., from the Kishenehn shale in Montana (46 ma). With the aid of time-
of-flight secondary ion mass spectrometry, very high level of iron in the
abdomen of the midge could be detected (Greenwalt et al., 2013). This
was interpreted as a fossilized blood meal providing a direct evidence for
haematophagy in fossil representatives of Culicidae.
Insects as Parasites and Hosts 157

• The oldest representative of Corethrellidae is Corethrella cretacea, 122 ma


from Lower Cretaceous Lebanese amber (Szadziewski, 1995; Borkent,
2008).
• The fossil record of Ceratopogonidae is one of the best known among
insects with 274 fossil species (Borkent, 2014) and with countless spec-
imens in all 15 major amber deposits currently known. This abundance
is connected to (1) the close association of their habitats and resin sour-
ces, (2) their small size and (3) their swarming behaviour (Szadziewski
et al., 2015; Pérez-de la Fuente et al., 2011). Amber specimens have
been recorded from Lebanese amber (122 ma) to Dominican amber
(15 ma; Szadziewski et al., 2015; Choufani et al., 2014; Borkent,
2000; Sontag and Szadziewski, 2011). Compression fossils of biting
mites are rare. The oldest compression fossil is Archiaustroconops besti
of Southern England from the Purbeck Limestone group (146 ma), rep-
resenting a single wing (Borkent et al., 2013). The oldest specimen
ascribed to Leptoconops, and hence a presumed representative of Leba-
noculicoidinae (sister group to all other Ceratopogonidae), was found
in Lebanese amber (122 ma) (Borkent et al., 2013; Borkent, 2001,
2000). Leptoconops also has extant representatives. This was used as an
indication of morphological stability in Ceratopogonidae for million
of years, which has been related to a lack of change in diet of Cerato-
pogonidae during the radiation of the angiosperms (Borkent and Craig,
2004; Choufani et al., 2013).
• Archicnephia ornithoraptor from the Cretaceous amber (90 ma) of New Jer-
sey is the only amber inclusion of a representative of Simuliidae known
so far (Grimaldi and Engel, 2005). The oldest definitive representatives of
Simuliidae are Kovalevimyia lacrimosa from the Late Jurassic of Siberia (176
ma) and a Simulimima grandis pupa from the Middle Jurassic of Siberia
(Currie and Grimaldi, 2000).
• Representatives of Psychodidae are diverse and abundant in the fossil
record with 74 species (Azar and Nel, 2003). Psychodidans occur in
Lower Cretaceous Burmese, Lebanese, France and Spain amber (120e
135 ma) to Miocene Mexican and Dominican amber, but also in the
Early Jurassic of Germany (180 ma) and in the Late Triassic of Virginia
(210 ma; Azar et al., 2015b; Ibanez-Bernal et al., 2014; Petrulevicius
et al., 2011; Andrade Filho et al., 2009; Andrade Filho et al., 2008;
Andrade Filho et al., 2007; Blagoderov et al., 2007; Poinar, 2007;
Grimaldi and Engel, 2005; Andrade Filho and Brazil, 2003; Azar and
Nel, 2003; Brazil and Andrade Filho, 2002; Fraser et al., 1996; Ansorge,
158 Christina Nagler and Joachim T. Haug

1994). Sycoracinae, a more derived psychodid in-group, occurred first in


Cretaceous amber of France and New Jersey (Azar and Salame, 2015;
Petrulevicius et al., 2011; Azar, 2007). Two discoveries give direct evi-
dence for the haematophagy of Psychodidae. Penalver and Grimaldi
(2005) reported a swarm of specimens of Phlebotominae associated
with mammalian hair from the Miocene Dominican amber. The more
spectacular one is the observation of nucleated erythrocytes with parasi-
tophorous vacuoles in the gut of an amber-embedded sand fly (Poinar
and Poinar, 2004a).
• Due to the close connection between the water snipe flies (Tabanomor-
pha) and trees, they are a well-represented group in fossil deposits; 80
species have been described (Solorzano Kraemer and Nel, 2009). Gallia
alsatica, one of the oldest fossils among snipe flies, comes from the Jurassic
of France (200 ma; Krzeminski and Krzeminska, 2003). Snipe flies have
been considered to be abundant during the Jurassic and declined during
the Cretaceous (Mostovski, 2008). However, numerous Cretaceous fos-
sils have been found, which indicate that snipe flies were in fact still
abundant and widespread (Solorzano Kraemer and Nel, 2009). A recent
finding from China, Qiyia jurassica, an aquatic parasitic fly larva from the
Jurassic (150 ma) could be a representative of the early lineage of
Tabanomorpha and would support their supposedly Jurassic age (Chen
et al., 2014).
• Horse flies (an in-group of Tabanomorpha) have been reported from
Miocene Mexican amber (20 ma; Strelow et al., 2013) from Dominican
amber, Baltic amber (Trojan, 2002) and the Cretaceous New Jersey
amber (Grimaldi et al., 2011) with 28 species (Strelow et al., 2013).
The oldest representatives of Tabanidae are impression fossils from the
Lower Cretaceous of China (Zhang, 2012), Europe (Mostovski et al.,
2003) and America (Grimaldi et al., 2011).
Summing up, first representatives of Diptera are as old as the Triassic.
Their evolutionary history shows a gradual evolution from ancestral
blood-feeding, hence ectoparasitism, to a predatory or nectar-feeding life-
style, sometimes also with reversal to blood sucking. Several indications of
haematophagy (Greenwalt et al., 2013; Poinar, 2011b, 2007, 2005a,b,
2004a; Poinar and Telford, 2005; Poinar and Poinar, 2004a) and two asso-
ciations with the remains of their host (Vullo et al., 2010; Penalver and
Grimaldi, 2005) give also more evidence for their parasitic lifestyle and func-
tion as vectors in the past. Within the brachyceran lineages of Diptera several
groups evolved parasitoidism. Yet this behaviour is quite scattered
Insects as Parasites and Hosts 159

(phylogenetically) and not associated with functional morphological special-


izations. Hence identifying this behaviour in a fossil is more than challenging
and therefore not further followed here.

4. INSECT PARASITISM S.STR. e NEUROPTEROIDA


Comparably recent molecular- and morphologically based phyloge-
netic analyses indicate that Coleopterida and Neuropterida are sister groups,
the parent group named Neuropteroida (Beutel et al., 2011; Wiegmann
et al., 2009). Neuropterida consists of Neuroptera þ Megaloptera (Asp€ ock
and Asp€ ock, 2008) and its sister group Raphidioptera (Beutel et al., 2014;
Wiegmann et al., 2009).

4.1 Neuroptera (Mantispidae)


Within Neuroptera, there are approximately 6000 described species (Beutel
et al., 2014). Neuroptera has been considered to have originated in the
Permian (Grimaldi and Engel, 2005). The only known parasitic in-group
of Neuroptera is Mantispidae.

4.1.1 General aspects


Mantispidae (mantid lacewings) is a group within Neuroptera with 400
species known globally (Engel and Grimaldi, 2007). They are distinct by a
mantodean-like appearance e they have prominent raptorial forelegs, elon-
gated ‘necks’ and well-developed eyes (Grimaldi and Engel, 2005). There
are generally four in-groups recognized: Drepanicinae, Calomantispinae,
Mantispinae and Symphrasinae. All larvae are all highly specialized obligate
parasites of spiders or hymenopterans (Grimaldi and Engel, 2005); most
larvae of Mantispinae are parasitic on spider eggs.
The first instar larva is active and very mobile, whereas the next two
stages are immobile. Larvae of Mantispinae are obligate parasites of spiders’
egg sacs. Some larvae are ectoparasitic on spiders, before these produce an
egg sac. Hence they feed on haemolymph until they finally can parasitize
the egg sac and feed on the eggs (Ohl, 2011; Engel and Grimaldi, 2007).
It is necessary that the larvae are agile, to find and reach a spider’s egg sac.
If a mantispid larva attaches to a juvenile spider, the larva remains within
the chambers of the book lungs of the spider until the next year, when
the spider will become adult and produce an egg sac. Thus, the host spider
induces and terminates a diapause of the mantispid larvae (Ohl, 2011). The
160 Christina Nagler and Joachim T. Haug

larva of the fossil Dicromantispa moronei has been proposed as a parasitoid of


spiders, yet without further explanation (Engel and Grimaldi, 2007).

4.1.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Grimaldi and Engel (2005) suggested a radiation within Mantispinae during
the Palaeogene (60 ma) hence a rather young age. Haring and Asp€ ock
(2004) predicted an origin in the Late Triassic (202 ma) or the Early Jurassic
(200 ma) based on molecular data. Also adult forms provide evidence since
the Early Cretaceous (Shih et al., 2015; Poinar and Buckley, 2011). A fossil
larva in amber found by Ohl (2011) (see below) is the first direct fossil record
of a parasitic larva of Mantispinae and provides their minimum age (Eocene,
44 ma). Two fossil species ascribed to Drepanicinae demonstrate an appear-
ance of these at least in the Late Cretaceous (100e66 ma; Shi et al., 2015).
Hence also Mantispidae is at least that old.

4.1.3 Fossil representatives


• Ohl (2011) found one spider-boarding larva ascribed to Mantispinae in
Middle Eocene Baltic amber (44 ma) on a juvenile clubionid spider in
a typical position, anterior on the opisthosoma (Figure 3(e)). The larva
has a flattened, elongated body and bears strong thoracic legs, that imply
active movement of the larva. This morphology indicates that it is likely
the first instar larva of a species of Mantispinae. This represents the first
and so far only discovery of a fossil mantispid larva (Ohl, 2011).
• Adult mantispidans, which would indirectly indicate the presence of
parasitic larvae, are rare in the fossil record (Engel and Grimaldi, 2007;
Wedmann and Makarkin, 2007). Only 14 species have been ascribed
to Mantispidae. Five of them were found in amber: Micromantispa cristata
(Figure 3(d)) in Cretaceous amber of Myanmar (Shih et al., 2015),
Doratomantispa burmanica in Burmese amber (Poinar and Buckley,
2011), Dicromantispa electromexicana in Mexican amber (Engel and
Grimaldi, 2007), D. moronei in Dominican amber (Engel and Grimaldi,
2007), Feroseta prisca in Dominican amber (Poinar, 2006).
• The others were found in deposits dating from the Jurassic (Wedmann
and Makarkin, 2007; Ansorge and Schl€ uter, 1990; Panfilov, 1980), the
Cretaceous (Makakrin, 1996; Makarkin, 1990), the Eocene (Wedmann
and Makarkin, 2007; Cockerell, 1921) and the Oligocene (Nel, 1988).
Two additional fossils ascribed to Mantispidae are two species of Mantis-
pidiptera, described by Grimaldi (2000) and Whalfera venatrix, described by
Insects as Parasites and Hosts 161

Engel (2004). The assignment of these two fossils to Mantispidae is


doubtful at best (Wedmann and Makarkin, 2007).

4.2 Coleopterida (Coleoptera, Meloidae)


Coleoptera and Strepsiptera are now considered to represent sister groups
within the monophyletic group Coleopterida (Beutel et al., 2014). This
sister group relationship has been established rather recently based on molec-
ular- and morphologically-based phylogenetic analysis (Beutel et al., 2014,
2011; Niehus et al., 2012; Wiegmann et al., 2009).
Approximately 355,000 extant species and 600 fossil species of
Coleoptera have been described (Grimaldi and Engel, 2005). The lineage
of Coleoptera is extremely diverse and abundant on all continents (Beutel
et al., 2014). Archostemata is thought to be the (extant?) sister group of all
remaining coleopterans. They have been considered to have originated in
the Permian (Beutel et al., 2014). Yet, Bethoux (2009) interpreted the
Carboniferous fossil Adiphlebia lacoana as a possible early coleopteran
implying an older age of the coleopteran lineage. This interpretation has
been questioned by Kukalova-Peck and Beutel (2012). In the Late Triassic,
the large coleopteran in-groups Myxophaga and Adephaga radiated, also
first representatives of Polyphaga appeared (Beutel et al., 2014; Grimaldi
and Engel, 2005). Another radiation of Coleoptera took place in the
Jurassic. A dramatic increase in species richness occurred in the Late Creta-
ceous, possibly coupled to the radiation of the angiosperm plants (Beutel
et al., 2014).

4.2.1 General aspects


Blister beetles (Meloidae, Coleoptera) are a monophyletic in-group of Ten-
ebrionoidea with almost 3000 species (Bologna et al., 2008). Meloidae have
two exciting evolutionary modifications: the terpenoid cantharidin and
hypermetabolous development (Grimaldi and Engel, 2005). Due to the
damaging effect on the digestive tract of animals, the terpenoid cantharidin
is effective against predators. Thus it protects meloids and their eggs from
predators. It also acts as a sexual attractor and is linked to the sexual behav-
iour in the in-group Meloinae (Bologna et al., 2008; Capinera, 2003).
Cantharidin causes blisters on the skin of sensitive animals (some humans
or horses), which led to their common name ‘blister beetles’ (Capinera,
2003). Hypermetabolous development includes seven larval stages, where
the first so-called triungulin phase is the most important one, when we
look on parasitism (Grimaldi and Engel, 2005). The triungulin larva
162 Christina Nagler and Joachim T. Haug

(Figure 3(f)) is often highly specialized for ectoparasitic behaviour in the


form of phoresy (Chmielewski, 1977 named phoresy ‘transportation para-
sitism’) and is adapted to reach its host. One of their strategies is to wait
on flowers and attach to the host with their grasping claws on the thoraco-
pods. Thus, the host takes them in its nest and they parasitize the clutch
(Engel, 2005). They parasitize Hymenoptera (Apoidea) and sometimes
grasshoppers (Acridoidea; Bologna et al., 2008; Capinera, 2003). Once
locating the host’s egg pod, the larvae feed on eggs and moult briefly in
the second stage from larval instars IIeV (Capaneira, 2003). After two
more stages, the larva will be a pupa and is found in the soil.

4.2.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
The triungulin larva evolved once in the stem-group Meloidae (Meloinae,
Tetraonycinae, Nemognathinae) around 112 million years old. It has been
suggested that phoresy appeared at least seven times independently within
Meloidae (Bologna et al., 2008), but appears unlikely when applying the
concept of Occam’s razor. Molecular clock studies indicate an even earlier
appearance of Meloidae, during the Early Cretaceous (125e135 ma).
They appear to have radiated fast with the acquisition of parasitism and
hypermetabolism (Bologna et al., 2008).

4.2.3 Fossil representatives


There are three fossils of meloid triungulins.
• Poinar (1992) described a larva of Meloidae sp. on the pronotal dorsal
surface of the extinct bee, Proplebeia dominicana, from Early Miocene
Dominican amber. A similar relationship between Meloidae and stingless
bees is not known nowadays, thus this fossil indicates a loss of this hoste
parasite relationship (Engel, 2005).
• Larsson (1978) described an isolated meloid triungulin from Middle
Eocene Baltic amber, yet this interpretation has been questioned (Engel,
2005).
• The oldest and definitive direct observation of a hosteparasite relation-
ship has been found in Middle Eocene Baltic amber (45 ma; Engel,
2005). Engel (2005) described meloid triungulins on the setae of the
bee Protolithurgus ditomeus (Megalichidae, Hymenoptera). If you consider
that bees are quite uncommon in amber, it was very lucky to find one
with meloid triungulins on it.
Insects as Parasites and Hosts 163

5. PARASITOIDS
5.1 Hymenoptera
5.1.1 General aspects
Hymenoptera is one of the most diverse groups among insects. Currently
there are 125,000 extant described species, traditionally divided in two
groups: ‘Symphyta’ (‘plant wasps’) and Apocrita (‘waisted wasps’) (Delcl os
et al., 2007). Parasitoidism, hyperparasitoidism and kleptoparasitism occur
in numerous in-groups: Orussidae (a group of ‘plant wasps’), Stephanoidea,
Ephialtioidea, Megalyroidea, Trigonalyoidea, Evanoidea, Proctotrupo-
morpha, Ichneumonoidea (all of these Apocrita, formerly also together
called Parasitica) and some groups of Aculeata (Grimaldi and Engel,
2005). Surprisingly the hosts of most extant species appear to be unknown
(Engel and Perrichot, 2014).
The only symphytan parasitoid group is Orussidae, which is the likely sis-
ter group to Apocrita (Grimaldi et al., 2002). Given this pattern of character
distribution, parasitoidism appears to have evolved in or slightly before the
stem species of Orussidae þ Apocrita (and was lost in apocritan in-groups).
Orussidans were diverse and abundant in the Jurassic and Cretaceous and
were associated with trees due to their ectoparasitoidism of wood-boring
beetles and wood wasps. The wasp waist is restricted to Apocrita. This
morphological character appears to provide them a better mobility of the
abdomen and hence the ovipositor. This appears to have been an essential
novelty for effective parasitoidism.
Like species of Orussidae, species of apocritan groups branching off early,
Stephanoidea and Megalyroidea, are ectoparasitoids of wood-boring beetles
and symphytan wasps (Grimaldi and Engel, 2005). Stephanoideans have a
very similar biology to species of Orussidae. The females use their elongated
ovipositor to pierce through wood into the host that is later eaten by the
parasitoid larvae (Engel et al., 2013a). This type of biology is also retained
the next group branching of the lineage towards the more derived apocri-
tans, Ephialtioidea (Figure 4(a)).
Representatives of Megalyroidea parasitize not only wood-boring bee-
tles (as the groups before) but also spheciform wasps. Their close relatives,
trigonalyoideans, are hyperparasitoids, hence are parasitoids on parasitoid
larvae (Perrichot, 2013).
Representatives of Evanoidea are diverse. The supposedly ancestral
appearing representatives of Aulacidae (Figure 4(b)) are parasitoids, but the
164 Christina Nagler and Joachim T. Haug

Figure 4 Fossil representatives of Hymenoptera. (a) Praeproapocritus flexus (Ephialtiti-


dae, Ephialtioidea) (modified with permission from John Wiley & Sons from Li et al.
(2013, Figure 3)). (b) Aulacus eocenicus (Aulacidae, Evanoidea) (modified after Nel and
€eg (2004, Figure 1)). (c) Tagsmiphron spiculum (Stigmaphronidae, Ceraphronidea)
Plo
(modified with permission from Elsevier from McKellar and Engel (2011, Figure 2A)). (d)
Necrobythus pulcher (Scolebythidae, Chrysidoidea) (modified with permission from
Elsevier from Engel et al. (2013b, Figure 4(b))). (e) Deinodryinus areolatus (Dryinidae,
Chrysidoidea) (modified after Gugliemino and Olmi (2011), Figure 2).

more derived forms such as evaniids and gasterupiids are larval predators of
cockroach oothecae (Grimaldi and Engel, 2005). Their inferred oldest rela-
tives are species of the extinct group Praeaulacidae; species of this group
were abundant in the Jurassic and Early Cretaceous of Australia and Asia
(Li et al., 2014).
Proctotrupomorpha is a group of usually very small and tiny parasitoids
(except the giant Pelicinidae). In-groups are Stigmaphronidae, Ceraphroni-
dea, Platygastroidea (Grimaldi and Engel, 2005). These wasps are
Insects as Parasites and Hosts 165

endoparasitoids of wood-boring or cone-boring beetles, flies, symphytans,


lepidopterans or green lacewings and some are hyperparasitoids on other
apocritan species (Shih et al., 2013). Species of Platygastridae and Scelionidae
(in-groups of Platygastroidea) are both endoparasitoids of insect and spider
eggs. Scelionidan species do not complete their development until the
host has reached adultness (Grimaldi et al., 2002). The known representa-
tives of Ceraphronidea (Figure 4(c)) are either endoparasitoids of caterpillars,
beetles, dipteran pupae, Hemiptera, other apocritans or Thysanoptera or
hyperparasitoids on apocritans or aphidid hemipterans (McKellar and Engel,
2011).
Among Ichneumonoidea (Braconidae and Ichneumonidae) the species
of less-derived in-groups are ectoparasitoids of wood-boring beetles, cater-
pillars and wood wasps. Within Ichneumonoidea different lineages switched
multiple times from ecto- to endoparasitoidism (Grimaldi and Engel, 2005).
All ichneumonid endoparasitoids possess a special poison that manipulates
the host to construct a specific protection for the ichneumonid larva. Ecto-
parasitoids in contrast suppress the immune system of the host with a poly-
DNA-virus (Longdon and Jiggins, 2012). Although ichneumonids are well
reported since the Early Cretaceous, such a virus could not be directly
detected or visualized (Longdon and Jiggins, 2012).
In Aculeata (including ants, bees, stinging wasps), only a few lineages
retain a plesiomorphic parasitoid lifestyle. This is known in Chrysidoidea,
Vespoidea and Apoidea. Chrysidoidea appears to be the less-derived group
and is rather species poor. Representatives are ectoparasitoids of butterfly
larvae (Lepidoptera), of beetle larvae (Coleoptera), of web spinners
(Embioptera), of walking sticks eggs (Phasmatodea) and sometimes klepto-
parasitic on bees and vespids (Azar, 2007). The larvae of species of Dryinidae,
an in-group of Chrysidoidea, parasitize adult and nymphal cicadas (Auche-
norrhyncha; Olmi et al., 2014a; Guglielmino and Olmi, 2011). The oldest
fossil dryinids have been ascribed to a cosmopolitan extant genus. This has
been seen as an indication that the morphology of these forms was preserved
since the Cretaceous (Olmi et al., 2014b).
Within Vespoidea only few ‘primitive’ lineages are parasitoids, namely
Tiphiidae, Pompilidae, Sapygidae, Mutillidae, Rhopalosomatidae and
Scoliidae. Tiphiidae comprises species that are ectoparasitoids of ground-
dwelling larval beetles, such as species of Scarabeidae. Species of Pompilidae
are also ectoparasitoids of spiders, they paralyze their hosts. Sapygidae in-
cludes forms that are ectoparasitoids of larval wasps or kleptoparasites of
bees. Species of Mutillidae are ectoparasitoids of different holometablous
166 Christina Nagler and Joachim T. Haug

groups, such as bees, wasps, flies, moths, beetles, but also cockroaches. Spe-
cies of Rhopalosomatidae are ectoparasitoids of crickets. Lastly species of
Scoliidae are ectoparasitoids of ground-burrowing beetles (Grimaldi and
Engel, 2005).
Also in the lineages that diverged from the ancestral type of parasitoidism
and switched to other strategies, parasitoidism (or parasitism?) has (re-)
evolved. For example, some ants are secondarily ectoparasitoids or klepto-
parasitic, but no fossil forms of these have been found so far (Grimaldi
and Engel, 2005). Within eusocial bees, different types of kleptoparasitism
are known (Litmann et al., 2014). Also species of spheciform wasps, Ampu-
licidae, Sphecidae and Crabronidae are parasitoids. Their larvae feed on the
living host for several larval stages, entering the host’s body and live as endo-
parasitoid. After the host is dead, the larvae spin a cocoon and pupate within
the host (Grimaldi and Engel, 2005). Spheciform wasps are evolutionarily
important as they are the sister group to bees (e.g. Johnson et al., 2013).

5.1.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
It has been put forward that the key feature in hymenopteran evolution was
the development of parasitoidism, which triggered the largest radiation of
parasitic arthropods in the wide sense (including parasitoidism). Derivatives
of early branchings in Hymenoptera are phytophagous, xylophagous and
fungivorous, but with the origin of parasitoidism (estimated about
210 ma) hymenopterans diversified rapidly.
These forms should have evolved at least in the Early Triassic, as Hyme-
noptera is considered to be the sister group to the remaining holometabolans
(Wang et al., 2014a; Engel et al., 2013a; Klopfstein et al., 2013; Sharkey
et al., 2012; Heraty et al., 2011; McKenna and Farrell, 2010). Yet, it remains
unclear if these suggestions include the possibility that we might have holo-
metabolous forms already in the Carboniferous.
The fossil wood wasp Cratoenigma articulata from the Lower Cretaceous
of Brazil seems to be an ‘intermediate’ form between the endophytic (non-
parasitoid) forms of Xyledoidea and the parasitoid species of Orussidae
(Krogmann and Nel, 2012).
Based on phylogenetic and morphological studies, it is likely that the
radiation of Orussidae took place in the Middle Cretaceous (100 ma;
Vilhelmsen and Zimmermann, 2014). Fossils ascribed to Proctotrupomor-
pha from the Cretaceous support a possible Jurassic age of the group
(Grimaldi and Engel, 2005). Within the proctotrupomorph Platygastroidea
Insects as Parasites and Hosts 167

the in-group Chalcidoidea underwent a spectacular radiation during the


Eocene (55e35 ma; Heraty and Darling, 2009). Chalcidoidean species are
endoparasitoid or hyperparasitoid on a large range of hosts among insect par-
asitoids. The group has been estimated to comprise not less than 500,000
species (Heraty and Darling, 2009). For Ceraphronidea, a Cretaceous origin
has been proposed, based on their biogeographic distribution (Perrichot
et al., 2014).
In short, Hymenoptera evolved latest in the Middle Triassic (230 ma, but
probably even earlier), parasitoidism in this group originated near the
TriassiceJurassic boundary (210 ma). This lead to a rapid radiation of
Apocrita in the Jurassic (195 ma). Aculeatan species existed from the Jurassic
(155 ma) onward and radiated in the Early Cretaceous (140 ma). The
occurrence of ants and bees was coupled to the radiation of angiosperms
in the Middle Cretaceous (125e100 ma; Ward et al., 2014; McKellar and
Engel, 2012).

5.1.3 Fossil representatives


• A specimen of Xyleoidea (saw flies) represents the oldest hymenopteran
fossil. It comes from the Middle Triassic (220 ma) of Kyrgyzstan in Cen-
tral Asia (Rasnitsyn, 1969). Slightly younger finds are from the Upper
Triassic of Argentina (Lara et al., 2014) and from the Middle Jurassic
of China (Wang et al., 2014b). Among the saw flies is the largest
hymenopteran fossil with an estimated wing span of 92 mm (Gao
et al., 2013b).
• The fossil record of Orussidae is sparse with only two specimens from
Cretaceous amber of Siberia and New Jersey, and two specimens from
the Late Jurassic of Kazakhstan and from the Oligocene of Colorado
(Vilhelmsen and Zimmermann, 2014; Vilhelmsen et al., 2013; Delcl os
et al., 2007; Grimaldi and Engel, 2005; Vilhelmsen, 2003).
• There are only a few fossils of Stephanoidea known so far. They have
been found in the Late Cretaceous of New Jersey, the Middle Eocene
Baltic amber and the Middle Cretaceous Burmese amber. Additionally
there is a single compression fossil from the EoceneeOligocene bound-
ary of Florissant (Engel et al., 2013a; Grimaldi and Engel, 2005).
• The few fossils ascribed to Ephialtioidea stem from the Early Jurassic to
the Early Cretaceous (Li et al., 2013).
• Megalyroideans are known from Cretaceous Siberian amber, Burmese
amber and New Jersey amber and from Baltic amber (Delcl os et al.,
2007). An in-group, Stigmaphronidae, exclusively known from the
168 Christina Nagler and Joachim T. Haug

Cretaceous, includes an additional fossil that provides direct evidence for


their endoparasitoidism (Engel and Perrichot, 2014; Arillo, 2007): a fossil
stigmaphronid hymenopteran from Mesozoic Spanish amber shows ovi-
positing into a dipteran (Alonso et al., 2000).
• Trigonalyoideans are rare in amber, but abundant as compression fossils
(Early Cretaceous; 125 ma) (Engel and Perrichot, 2014; Delcl os et al.,
2007).
• Evanoid fossils have been described from Cretaceous Lebanese amber,
Myanmar amber and New Jersey amber, as well as based on compression
fossils from the Lower Cretaceous (Engel, 2013a; Jennings et al., 2013;
Delcl os et al., 2007; Grimaldi and Engel, 2005; Nel et al., 2004).
• Proctotrupomorphans have been reported from the Early Jurassic of Asia,
from the Cretaceous of China, from the Cretaceous of Spain, from the
Lower Cretaceous of Brazil and from the following amber deposits:
New Jersey Baltic, Burmese, Dominican, Canadian, Mexican, Late
Eocene, Eocene Rovno, Miocene amber from Peru (Perrichot et al.,
2014; Barling et al., 2013; Engel, 2013b; Krogmann, 2013; Shih et al.,
2013; Kolyada and Petrovsky, 2011; McKellar and Engel, 2011; Poinar
and Huber, 2011; Heraty and Darling, 2009; Delcl os et al, 2007;
Grimaldi and Engel, 2005).
• Representatives of platygastrid Serphitoidea as well as of Scelionidae and
Platygastridae are already known from the Cretaceous, but are not found
in younger deposits (Engel and Perrichot, 2014; Grimaldi and Engel,
2005; Grimaldi et al., 2002).
• The early record of modern in-groups of Ichneumonidea comes from
the Lower Cretaceous. Fossils are abundant as both, amber inclusions
and as compression fossils until the Palaeogene (McKellar et al., 2013;
Ortega-Blanco et al., 2012; Delcl os et al., 2007; Grimaldi and Engel,
2005). Notable specimens are fossils come from Baltic and Dominican
amber: (1) a braconid larva emerging from an ant (Poinar, 2002) (2) an
ichneumonid female ovipositing into a caterpillar (Poinar and Miller,
2002) (3) an ichneumonid larva attached to a spider (Poinar, 1992) and
(4) an ichneumonid larva that spins its cocoon over the eggs of a spider
(Poinar, 2004b).
• There are more fossil forms of Chrysidoidea, dating from the Early
Cretaceous to the Early Miocene, of (Figure 4(d)) than there are extant
species (Engel et al., 2013b; Grimaldi and Engel, 2005). Within Chrys-
idoidea, the group Dryinidae (Figure 4(e)) is the best recorded ectopar-
asitic group in amber (Arillo, 2007). Grimaldi and Engel (2005) reported
Insects as Parasites and Hosts 169

a fossilized leafhopper with an attached thylacium, a protective structure


formed by retained parts of the exuviae of the first instar larva of the
parasitoid.
• Most vespoid in-groups occur since the Early Cretaceous, but the oldest
known specimen is a pompilid spider wasp (hence likely a parasitoid)
from Middle Cretaceous amber of Myanmar (Grimaldi et al., 2002).
• There are several spheciform wasps from the Caenozoic preserved as
compression fossils and amber inclusions (Antropov, 2000; Antropov
and Pulawski, 1996). The Caenozoic fossils resemble the modern spheci-
form wasps (Grimaldi and Engel, 2005).
• The Cretaceous Pompilopterous (Rasnitsyn, 1975), as the name suggest,
was interpreted as a spider wasp, but has been redescribed as a spheciform
wasp (Rasnitsyn, 1998). In a comparable case, the Lower Cretaceous fos-
sil Cariridris bipetiolata (Brandao et al., 1990) has been originally described
as ant, but was redescribed as a spheciform wasp (Dlussky and Rasnitsyn,
2003).

5.2 Strepsiptera
5.2.1 General aspects
Strepsipterans are obligate endoparasitoids; Strepsiptera comprises ca. 600
extant species. These parasitize numerous different types of insects, mainly
representatives of Auchenorrhyncha (like Cicadellidae, Membracidae),
Hymenoptera, Hemiptera and Zygentoma (Rasnitsyn and Quicke, 2002).
Part of their success has been attributed to their pronounced sexual
dimorphism and life cycle. Strepsipteran males bare halteres and hind wings.
They are agile fliers, but live only a few hours, to find a female
(Kathithiramy, 2009; Grimaldi and Engel, 2005). Most females are wingless
and sack-like, lack functional eyes and antennae. In some groups which
retain some more ancestral features, females bear short antennae, reduced
(but still present) eyes and legs. They live their entire life within the host.
The hosts’ life span can last until all strepsipteran larvae emerge from the
host (Kathithiramby, 2009).
The larvae hatch immediately when the egg is laid (ovovivipary). The
first larva, is called a triungulin (cf. Meloidae). This larva actively searches
for a host. Most of them search for larvae or nymphs, some go for eggs
(Kathithiramby, 2009). After finding a potential host, the strepsipteran larva
attaches to the host; special enzymes allow the larva to enter the host.
To avoid the host’s immune defence, the larvae mimic the host at a mo-
lecular level. This mechanism is unique among insect parasitoids and
170 Christina Nagler and Joachim T. Haug

explains their success (Kathitiramby, 2009). The development of strepsipter-


ans is classified as hypermetamorphosis (again cf. Meloidae). After the triun-
gulin larva, there are four further larval stages. Finally the last larval stage
emerges from host to pupate externally in all males. In species in the less-
derived group Mengenillidae, the females also leave the host to pupate.
All other females remain endoparasitic in the host, only the anterior region
is visible from the outside (Kathithiramby, 2009).
One special case within Strepsiptera, Myrmecolacidae shows the so-
called heterotrophic heteronomy (see Kathirithamby et al., 2010). The hosts
for males are ants, those for females are grasshoppers, crickets and mantids.
Pohl and Kinzelbach (2001) suggested that the primary hosts of Myrmeco-
lacidae were ants. A related case is known in endoparasitoid Aphelinidae
(Kathithiramby, 2009). Representatives of the extinct strepsipteran in-group
Mengeidae were parasitoids of Zygentoma (silverfish; Grimaldi and Engel,
2005).

5.2.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
The systematic and phylogenetic position of Strepsiptera has long been
controversial. They have been thought to be the most problematic group
of insects in terms of its systematic placement (Pohl et al., 2010). Grimaldi
and Engel (2005) proposed that it would not be surprising, if Strepsiptera
is a highly modified lineage, derived from holometabolans, such as stem-
group neuropterids or mecopteroids. Probably during the Eocene (45
ma), the transition between less-derived and modern strepsipterans took
place (Kathithitramby, 2009).
Recent molecular- and morphologically-based phylogenetic studies
provide support that Strepsiptera is the sister group to Coleoptera (Boussau
et al., 2014; Pohl and Beutel, 2013; Niehuis et al., 2012). The extinct forms
Protoxenos, Cretostylops and Mengea are considered to be the derivatives of the
evolutionary lineage towards modern Strepsiptera. This is based on morpho-
logical data, especially CT analysis of Mengea tertiaria in Baltic amber (Beutel
et al., 2014; Pohl et al., 2010). The newly discovered Bahiaxenidae
(Figure 5(a)) from Brazil has been considered to be the sister group of the
remaining modern Strepsiptera. They have been interpreted as a relic group
having changed only little since the Permian (Bravo et al., 2009). Within the
remaining modern strepsipterans, excluding Bahiaxenidae, Eoxenos has been
resolved as the sister group to all remaining forms (McMahon et al., 2011;
Bravo et al., 2009).
Insects as Parasites and Hosts 171

Figure 5 Fossil parasitoid strepsipterans (Figure 5a and b); plant parasitic gall wasp
(Figure 5(c)). (a) Bahiaxenos relictus (Bahiaxenidae, Strepsiptera) (modified with permis-
sion from John Wiley & Sons from Bravo et al. (2009, Figure 2)). (b) Bohartilla kinzelbachi
(Bohartillidae, Strepsiptera) (modified from Grimaldi and Engel (2005, Figure 10.88) with
kind permission of Cambridge University Press and David Grimaldi). (c) Gall-forming cyn-
ipidan (Cynipidae, Hymenoptera) (modified from Grimaldi and Engel (2005, Figure 11.17)
with kind permission of Cambridge University Press and David Grimaldi).

5.2.3 Fossil representatives


• The majority of strepsipteran fossils have been found in various types of
amber: Burmese (100 ma), Canadian (75 ma), Eocene Baltic (42e45 ma)
and Miocene Dominican (25 ma; Cook, 2014; Pohl et al., 2012, 2010,
2005; Kogan and Poinar, 2010; Kathirithamby and Hendricks, 2008;
Poinar, 2002; Pohl and Kinzelbach, 2001, 1995; Kulika, 2001;
Kinzelbach and Pohl, 1994; Kathirithamby and Grimaldi, 1993; Lutz,
1990; Kinzelbach and Lutz, 1985; Kinzelbach, 1983, 1979; Kulika,
1979; Ulrich, 1927; Meng, 1866). Mainly forms ascribed to Mengeidae,
Mengenillidae and Myrmecolacidae have been found. A great diversity
of less-derived Strepsiptera has been found in Cretaceous Burmese
172 Christina Nagler and Joachim T. Haug

amber. In Baltic and Dominican amber, more modern (Figure 5(b)) and
less-derived strepsipterans occur together.
• Some fossils provide quite direct evidence for the extraordinary life cycle
of strepsipterans and direct evidence for their parasitoid lifestyle: (1) two
puparia on an ant specimen in Middle Eocene oil slate of Messel (Lutz,
1990), (2) an ant with an emerging parasite preserved in amber (Pohl and
Kinzelbach, 2001), (3) an empty male puparium associated with a hyme-
nopteran preserved in amber, (4) two parasitized planthoppers, from one
of these planthoppers emerge larvae (Poinar, 2004b) and preserved in
amber, (5) a female specimen ascribed to Myrmecolacidae parasitizing
an ant in Baltic amber (Pohl and Kinzelbach, 2001).

6. PLANT PARASITISM (VERSUS PHYTOPHAGY)


6.1 General aspects
Numerous insects consume plants. Yet some insects also make use of
plants in other ways. Some insects oviposit eggs into plants and induce the
growth of galls; this is a form of plant parasitism. In leaf miners, the larvae
feed on the plants until they emerge as pupa or adult, thus it is comparable
to larval parasitoids feeding within insects. However, here the line to phy-
tophagy is not easy to draw.
Galls are remarkable structures occurring on all organs of a plant. With
galls an insect is forcing the plant to provide nutrition and/or a habitat,
thus this should represent a case of true parasitism. Galls can be also induced
by bacteria, fungi, mites, viruses and nematodes (Knor et al., 2013). Gall-
inducing insects are representatives of Cynipidae (gall wasps), Cecidomyii-
dae (gall midges), Psyllidae (jumping plant lice), Thysanoptera (thrips),
Anisoptera (dragonflies), phytophagous hymenopterans (sawflies), Coleop-
tera (beetles), Aphididae (plant lice), Tingidae (lace bugs), Cicadellidae (leaf-
hoppers) and some heterocerans (moths) (Knorr et al., 2013; Grimaldi and
Engel, 2005). Moth larvae are important leaf miners, the egg is oviposited
into the leaf and the larvae feed on the mesophyll until they pupate and
fall on the ground (Labandeira et al., 1994).
Exceptionally preserved fossils from the US provided an uncompressed
planteinsect system from the Palaeocene (Donovan et al., 2014; Wappler
et al., 2009). In such systems, it is possible to identify the parasite that caused
fossil mines and galls (Carvalho et al., 2014; Donovan et al., 2014; Grimaldi
and Engel, 2005), based on their unique characteristics; e.g. position on the
plant, shape, host specificity, path of the mine and size.
Insects as Parasites and Hosts 173

6.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Inducing galls has evolved multiple times independently among the insects.
This ability is considered to have originated first about 300 ma in the
Carboniferous in a not further known holometabolous larva (Labandeira,
2011). But this attribution has been considered doubtful, as also mites
may produce comparable galls (Grimaldi and Engel, 2005; Rasnitsyn and
Quicke, 2002).
A significant radiation of gall-inducing and also leaf-mining insects
occurred in the Late Cretaceous, most likely coupled to the radiation of
the angiosperms (Knorr et al., 2013; Rasnitsyn and Quicke, 2002;
Labandeira et al., 1994). Several fossils indicate an origin of Cynipodea in
the Jurassic. The origin of gall making in Hymenoptera occurred first in
the Middle Cretaceous when first forms of Fagacea (beeches and oaks)
appeared in the fossil record (Grimaldi and Engel, 2005).
The gall midges (Cecidomyiidae) are known since the Late Jurassice
Early Cretaceous (Grimaldi and Engel, 2005). The main radiation of Ceci-
domyiidae took place during the Palaeogene coupled to the evolution of
flowers (Wappler et al., 2010). These dipterans have been considered to
have originated in the Late Jurassic, at first parasitizing gymnosperms and
switching later to angiosperms (Labandeira et al., 1994). Most fossil galls
resemble their modern counterparts, which emphasizes the long-term
evolutionary coexistence and coevolution of host plant and parasitic gall-
making insect (Knorr et al., 2013; Krassilov, 2007). For a more comprehen-
sive review, see Labandeira (2013) and Labandeira and Currano (2013).

6.3 Fossil representatives


There are numerous plant fossils with galls (Palaeogallidae) and mines
(Palaeominidae) on plants as well as adult free-living plant parasitic insects.
Hence we only list here some examples:
• The oldest representative of Cynipidae (Figure 5(c)) is from Baltic amber
and could induce galls on rosaceans, like modern relatives (Grimaldi and
Engel, 2005).
• There is one extraordinary fossil: galls on the seed of a redwood from the
Miocene of Germany containing larval and pupal gall midges (M€ ohn,
1960).
• Pott et al. (2008) reported eggs which were found inside the interior of a
fossil leaf from the Carnian of Austria (150 ma). They suggested
174 Christina Nagler and Joachim T. Haug

the possible parent could have been a dragonfly, due to the elongated
ovipositor, that is needed to pierce through the robust leaves of bennet-
titalean plants. Yet also other insect groups possess long ovipositors.
• There are some leaf-mining fossils that appear extremely similar to their
modern counterparts: moths on oaks (for 20 million years), moths on
poplars (for 20 million years), moths on mahogany (for 40 million years)
and beetles on Heliconia (for 70 million years) (Grimaldi and Engel,
2005).

7. INSECTS AS HOSTS
7.1 Nematoida
7.1.1 General aspects
The monophyletic group Nematoidea consists of the two sister groups
Nematomorpha (hairworms) and Nematoda (roundworms) (Bleidorn
et al., 2002; Schmidt-Rhaesa, 2013).
Within the hairworms, 350 extant species are known which fall into two
groups: Gordiida with freshwater and terrestrial forms and Nectoma with
marine forms (Schmidt-Rhaesa, 2013). Gordiids have a free-living aquatic
phase and an obligate parasitic phase on terrestrial invertebrates. The parasite
goes through two hosts; the final host (insect) has to eat the intermediate
host (freshwater snail, small aquatic insects; Poinar, 2011a; Labandeira,
2002; Schmidt-Rhaesa and Ehrmann, 2001). Extant species parasitize
carabid beetles, grasshoppers, cockroaches and praying mantids (Schmidt-
Rhaesa and Ehrmann, 2001).
Nematoda includes about 10,500 species that are endoparasites of arthro-
pods and vertebrates (Poinar and Thomas, 2014; Poulin, 2006). Although
there are about 20,000 described extant nematode species, several million
species have been estimated (Schierenberg and Sommer, 2011; Poinar,
2012a). Fossil nematode parasites are treated in detail elsewhere in this
volume (Poinar, 2015).

7.1.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Nematoidea has been considered to have originated in the Precambrian
(Poinar, 2011a, 2013a, 2015b; Rota-Stabelli et al., 2013). Cambrian fossils
of cycloneuralians are diverse and hence indicate an older divergence of
the subgroups (e.g. Maas, 2013 and references therein).
Insects as Parasites and Hosts 175

7.1.3 Fossil representatives


It has been considered that less-derived forms parasitize less-derived insect
groups (Blattaria and Mantoptera) and more derived nematomorph groups
parasitize more derived insects, like coleopterans (Schmidt-Rhaesa and
Ehrmann, 2001), due to coevolutionary processes. Hence although the
direct fossils are rare, we could assume a similar age to these insect groups.
The oldest nematomorph fossils were found in Cretaceous deposits (Poinar
and Buckley, 2006).
There are some fossils that need to be discussed in reference to Gordiida
recorded so far.
• Cretachordodes burmitis comes from the Early Cretaceous Burmese amber
(100e110 ma), yet it is an isolated specimen, without a host (Poinar and
Buckley, 2006).
• Other nematomorph fossils include two females in association with
their cockroach host from Dominican amber (15e45 ma) (Poinar,
1999). A fungus and its ecology indicate, that their habitat was
moist to wet, which is attractive to roaches and carabids and that
small forest streams seem to be their aquatic habitat (Schmidt-Rhaesa,
2013).
• Other fossil worms of Palaeoscolecida may resemble Nematomorpha
(Xianguang and Bergstr€ om, 1994), but have been interpreted to
be early offshoots of the lineage towards Priapulida (Harvey et al.,
2010).
• Shergoldana australiensis from the Middle Cambrian of Queensland,
Australia, resembles a nematomorph larva in many aspects (Maas et al.,
2007). Yet, this form may have lived in the meiofauna, similar to kino-
rhynch cycloneuralians.

7.2 Mites
7.2.1 General aspects
Acari comprises about 55,000 described extant species, known as mites.
They can be free-living herbivorous and scavenging, as well as predatory,
or parasites of insects and vertebrates (Dunlop and Penney, 2012; Pereira
et al., 2012). They are also important vectors of several viruses and bacteria,
for example Borrelia burgdorferi that causes Lyme disease and might have
already been so in the past (Poinar 2014a,b, 2015). Mites use insects, partic-
ularly termites, ants and other arthropods for transport in a phoretic way also
(Dunlop and Penney, 2012).
176 Christina Nagler and Joachim T. Haug

7.2.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
The origin of mites was estimated to be in the Late Silurian (Mans et al.,
2011). For the parasitic in-group Astigmata an origin in the Middle
Permian, 270 ma, has been proposed based on molecular data (Garwood
and Dunlop, 2014; Dabert et al., 2010). Phoretic behaviour in mites has
been proposed to have evolved at least in the Eocene (44e49 ma; Dunlop
et al., 2012).

7.2.3 Fossil representatives


Acari are comparably abundant in the fossil record with 309 fossil records
(Dunlop and Penney, 2012).
• A representative from the supposedly more ancestral Anactinotrichida is
known from Rhynie chert of Scotland (395 ma) and other Devonian
sites (Norton et al., 1988).
• Some parasitic mites in amber are associated with their host. Parasitic
mites (Figure 6(c)) of the species-rich group Leptus have been preserved
in Baltic amber. They occur together with their host, a wide range of
dipterans, like fungus gnat (Limoniidae), window gnats (Empididae)
and Dolichopodidae, but also with different types of hymenopterans
(Dunlop et al., 2014; Arillo, 2007).
• In Baltic amber, water mites have been found too. They parasitize rep-
resentatives of Chironomidae (nonbiting midges) or Trichoptera (caddis
flies; Arillo, 2007). Mites parasitizing representatives of Drosophilidae,
Chironomidae, Sciaridae and moths have been reported from Domin-
ican amber (Arillo, 2007).
• Cretaceous fossils (in amber, but also compression fossils) of mites
(Figure 6(d)) preserve them still attached to their insect hosts (Judson
and Maakol, 2009; Arillo, 2007; Dunlop, 2007; Klompen and Grimaldi,
2001; Poinar et al., 1997).
• Mites from Spanish amber and Lebanese amber, both from the Lower
Cretaceous, have been reported to parasitize several forms of Diptera
(Arillo, 2007; Labandeira, 2002).
• There are also mites parasitizing hemipterans from several different
amber deposits (Poinar et al., 2012; Azar, 2007; Koteja and Poinar,
2005).
• Phoretic behaviour has been reported from Cretaceous amber sites and
Eocene Baltic amber (Dunlop et al., 2012; Azar, 2007).
Insects as Parasites and Hosts 177

Figure 6 Vectors and hosts. (a) Pheidole dentata (Myrmicinae, Formicidae) parasitized by
nematodes (modified after Poinar (2012a, Figure 6(c))). (b) Lutzomyia adiketis
(Phlebotominae, Diptera) (modified after Poinar (2008b, Figure 1)). (c) Myrmozercon sp.
(Mesostigmata, Acari) attached to Ctenobethylus goepperti (Dolichoderinae, Formicidae)
(modified from Dunlop et al. (2014, Figure 1(b)) with kind permission of the Royal Society).
(d) Erythraeid mite (Erythraeidae, Acari) on a chironomid midge (Chironomidae, Diptera)
(modified from Poinar and Krantz 1997, Figure 2 with kind permission from Springer Science
and Business Media). (e) Cheliferidae indet (Cheliferoidea, Pseudoscorpionida) (modified
from Judson (2009, Figure 5(b)) with kind permission of Geodiversitas and Mark Judson).

7.3 Pseudoscorpions
7.3.1 General aspects
There are currently 3533 described extant species and 41 fossil species of
pseudoscorpions (Harvey, 2013). Pseudoscorpions distantly resemble
scorpions, yet their posterior segments are not differentiated as a tail. Also
pseudoscorpions possess a characteristic cheliceral spinning apparatus
178 Christina Nagler and Joachim T. Haug

(Dunlop, 2010). The minute pseudoscorpions live in cryptic environments


and are predators of other nonvertebrates (Del-Claro and Tizo-Pedrosos,
2009; Grimaldi and Engel, 2005). Representatives of a few pseudoscorpion
groups show highly complex behaviour and are subsocial in the form of
brood care (Del-Claro and Tizo-Pedroso, 2009). All pseudoscorpions
have chelate pedipalps (a major reason why they distantly resemble scor-
pions) and many groups have a venom gland inside the pedipalp.
With their chelate pedipalps some pseudoscorpions temporarily attach
themselves to larger nonvertebrates. This is done in a way that does not
affect the locomotion capacities of the host, and allows dispersal via phoresy
(Poinar, 2013b; Boucot and Poinar, 2010; Judson, 2009; Szymkowiak et al.,
2007). Some pseudoscorpions are host-specific and some are also site-
specific on the host (Poinar, 2013b; Boucot and Poinar, 2010). Species of
Tridenchthoniidae, Lechytiidae, Syarinidae, Geogarypidae, Larcidae,
Sternophoridae, Cheiridiidae, Chernetidae and Cheliferidae are known to
be phoretic (Szymkowiak et al., 2007). Host species are known from
Coleoptera, Lepidoptera, Homoptera, Diptera (Aguiar and B€ uhrnheim,
1998) and other arthropods (Grimaldi and Engel, 2005).

7.3.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
The origin of arachnids has been considered to be in the Silurian. The fossil
record of pseudoscorpions reaches back to the Devonian (Dunlop, 2010).
Due to some complex morphological characters, traditionally pseudoscor-
pions are interpreted as the sister group to camel spiders (Solifugae; Dunlop,
2010). Recent morphological and molecular studies resolved a sister group
relationship between pseudoscorpions and scorpions (Garwood and
Dunlop, 2014; Pepato et al., 2010), yet this would mean that the long
time stable group Lipoctena would represent an artefact.
Cheliferoidea (which has representatives showing phoretic behaviour)
has been considered to be the most derived group of pseudoscorpions due
to their morphology and behaviour. The origin has been suggested to be
older than the Cretaceous, because representatives were diverse and wide-
spread in the Cretaceous already (Judson, 2009).
It has been proposed that phoresy originated at least five times within
pseudoscorpions (Judson, 2004). The phoretic relationship between pseudo-
scorpions and invertebrates provides evidence for a long-standing and stable
coevolution (Boucot and Poinar, 2010; Judson, 2004).
Insects as Parasites and Hosts 179

7.3.3 Fossil representatives


• The only non-amber and oldest pseudoscorpion fossil is Dracochela depre-
hendor from the Middle Devonian (374e92 ma), based on isolated cuticle
fragments (Schawaller et al., 1991). It remains unclear whether this spe-
cies was phoretic.
• The next younger fossil is known from approximately 250 million years
later in Cretaceous amber. It is a nymphal specimen attributed to Che-
liferoidea from Upper Cretaceous Canadian amber (Schawaller, 1991).
About the same age are three pseudoscorpions (two specimens of Heur-
taultia rossiorum and one undescribed specimen) (Figure 6(e)) from the
Lower Cretaceous Albian amber from France (Judson, 2009). Also there
are two specimens from Burmese amber (Judson, 2000; Cockerell,
1920), several specimens from Dominican amber (Judson, 1998) and
also from Baltic amber (Hendrickx et al., 2006). Another species has
been described from Miocene Chiapas amber (Riquelme et al., 2014).
Undescribed pseudoscorpions have been reported in Upper Cretaceous
New Jersey amber (Grimaldi et al., 2002), Lower Cretaceous Lebanese
amber (Grimaldi, 1996) and Lower Cretaceous Albian amber (Delcl os
et al., 2007).
• Five subfossil specimens of unclear exact age have been reported from:
(1) Zanzibar copal, Chelifer eucarpus, (2) two pseudoscorpions from
Madagascan copal and (3) two specimens from Colombian copal
(Judson, 2010).
• There are several fossils with direct very evidence for phoresy: pseudo-
scorpions on beetles (Poinar, 2013b, 1992; Grimaldi, 1996), on flies
(Judson, 2004; Poinar et al., 1998), on a snipe fly (Ross, 1998), on a nem-
atoceran fly (Kosmowska-Ceranowicz, 2001), on a braconid wasp
(Weitschart and Wichard 1998), on a moth (Boucot and Poinar,
2010), all preserved in various types of amber.

8. INSECTS AS VECTORS
8.1 General aspects
Extant insects are important vectors for numerous viruses, bacteria and
protist, but it is difficult to give fossil evidence for it (Poinar, 2014b). The
most important arthropods that serve as vectors are haematophagous insects
and arachnids (ticks).
180 Christina Nagler and Joachim T. Haug

Extant lice transmit significant human diseases. Examples are trench


fever, typhus, relapsing fever, but also wildlife diseases, like swine pox,
anaplasmosis, dermatomycosis, Lebombo virus or heartworms (Mitchell,
2015; Light et al., 2010).
Fleas via their mouthparts can transmit one of the most devastating
diseases in human history: the plague that is caused by the bacteria Yersinia
pestis. Also they are vectors for several worm parasites and protists (Grimaldi
and Engel 2006).
Additionally cockroaches, coprophagic beetles and synanthropic flies,
whiteflies, leafhoppers and treehoppers are important vectors of protozoan
parasites and viral elements, called geminiviruses (Graczyk et al., 2005;
Czosnek et al., 2001).
One of the most important groups in transmitting pathogens is Diptera,
especially in-groups such as sand flies, mosquitoes, blackflies and tsetse flies
(Poinar, 2014b). Almost 10% of all described extant sand flies transmit bac-
teria, viruses and protozoans, with trypanosomatids as the most widely
distributed pathogen. All forms of Trypanosomatida are parasitic (Lopes
et al., 2010). The intracellular protozoan Leishmania, an in-group of Trypa-
nosomatida is spread through bites by sand flies.
Phlebotomines transmit several animal trypanosomes, malaria parasites,
bacteria and rhabdovirus (Azar and Nel, 2003). It has been proposed that
free-living trypanosomatids associated with food sources for dipteran larvae
were eaten by these, and then were still present in the adult stage of a
dipteran, and then transmitted to vertebrates (Poinar, 2008b, 2007).
Ceratopogonids and mosquitoes are vectors for more than 300 viruses
that cause, for example, breakbone fever, river blindness or elephantiasis
(Grimaldi and Engel, 2005). The most ‘famous’ disease transmitted by culi-
comorphans is malaria that is also caused by a protozoan parasite.
An important endosymbiotic bacterium is Wolbachia that has been
reported to be transmitted by almost all terrestrial arthropods as well as nem-
atodes (Koutsovoulos et al., 2014).
Aphids and some beetle species mechanically transport fungus spores
and infect plants, due to their phytophagy, with phytopathogen fungi
(Kluth et al., 2002). Aphids, leafhoppers, spittlebugs, flea beetles, cucumber
beetles, psyllids and fruit flies serve as vectors for several bacterial phyto-
pathogens that damage plants (Nadarasah and Stavrinides, 2011). Endosym-
biotic bacteria, viruses and protozoans can influence the biology of
their insect vectors. They can be mutualistic, parasitic and commensalic
(Longdon and Jiggins, 2012). Also some viruses, cytoplasmic polyhedrosis
Insects as Parasites and Hosts 181

virus (CPV) and nuclear polyhedrosis virus (NPV), infect only insects and
are distributed among their populations, whereas some attack mammals
and birds also. These viruses are always fatal to the population of the
infected insects.

8.2 Phylogenetic inference of appearance and molecular


estimations of early evolution
Throughout their evolution sand flies and parasitic organisms likely have
coevolved. Due to their haematophagy parasitic diseases are likely as old
as the sand flies themselves (Azar and Nel, 2003). Thus the protozoan
parasites should have originated at least in the Early Cretaceous, 140 ma,
together with species of Ceratopogonidae as their ancient vectors (Pérez-
de la Fuente et al., 2011; Poinar, 2008a).
The association of hemipterans and trypanosomas originated at a similar
time, in the Middle Caenozoic, supposedly with bats as primary hosts
(Otalora-Luna et al., 2015; Poinar, 2005c). Similar parasites that were
vectored by sand flies have been proposed to have contributed to the extinc-
tion of the dinosaurs (Azar and Nel, 2003). The similarity of fossil and extant
and fossil trypanosomes has been seen as a strong indication for haematoph-
agy in the past (Poinar, 2004a; Greenwalt et al., 2013).
The interaction between wolbachians and insects is stable at least since
100 million years according to some molecular studies (Cerveau et al.,
2011).
The recent discovery of Arsenophonus, a c-proteobacterium, probably
played an important role in the evolution of bat flies and louse flies for
the development of their true ectoparasitism, at least 20 ma (Duron et al.,
2014).
Current research on phytopathogens transmitted via insects sheds new
light on their evolution. The traditional role of insects might have evolved
over a transitional stage, where the insect served as vector and host, to the
final form with the insect as host. Thus phytopathogenic bacteria evolved
to become insectopathogenic bacteria (Nadarasah and Stavrinides, 2011).
Different insect-borne viruses, which are viruses transmitted by insects,
have been considered to exist for at least 100 million years. These may
have coevolved with certain species of Diptera (Psychodidae and Culico-
morpha) since the Early Cretaceous, thus providing an evolutionary path
for vertebrate infections (Poinar and Poinar, 2005). The long relationship
of parasites and insects indicates a coevolutionary mechanism (Czosnek
et al., 2001).
182 Christina Nagler and Joachim T. Haug

8.3 Fossil representatives


Despite the small size of many of the organisms, there have indeed been
some fossil finds in amber that were interpreted as including:
• numerous flagellates (similar to Trypanosoma) on a faecal pellet of a hae-
matophagic hemipteran from Dominican amber (Poinar, 2005c)
• two trypanosome parasites (Paleoleishmania proterus and Paleoleishmania
neotropicum) from the gut and proboscis of a specimen of Phlebotominae
(Figure 6(b)) from Dominican amber and Burmese amber (Poinar,
2008a; Poinar and Poinar, 2004b)
• a supposed CPV and Trypanosoma infection in a midge from Burmese
amber (Poinar and Poinar, 2005); the provided images cannot easily be
used to judge this ascription.
• a supposed NPV in a sand fly from Burmese amber (Poinar and Poinar,
2005). Also here the images provided in the description hardly allow a
confident judgement or interpretation.
• three fungal pathogens from insects in Dominican amber (Poinar and
Poinar, 2005).
• different plasmodia on dipteran specimens from Cretaceous amber
(Poinar, 2011b; Poinar and Telford, 2005)

9. CONCLUSION
Given the shear vastness of examples of parasitism in and by insects, it
is hard to draw any major conclusions. However, a few overriding patterns
can be mentioned.
Two of the four hyperdiverse lineages appear to have their early radia-
tions coupled to parasitism sensu lato, namely Hymenoptera and Diptera.
So we can indeed ask what made these two lineages so successful. In both
cases, the parasitic act is directly linked to the provision of nutrition to
offspring, in early hymenopterans by injecting the egg into a host, and in
dipterans by haematophagy of the female to have enough nutrients to be
stored in the eggs. Compared to other lineages the parasitic act is therefore
less of an ‘everyday business’, which mainly means that in both lineages the
mobility of the adult is retained. This is quite different, for example, in fleas
or lice, where developing successful dispersal mechanisms are likely one of
the important factors in the early diversification of these two lineages.
Another similarity is that in both Hymenoptera and Diptera, there were in-
dependent switches to feeding on flower products, which likely triggered a
Insects as Parasites and Hosts 183

Figure 7 Timescale of the discussed groups. Dashed lines indicate the ghost range of
the respective groups. Solid lines are based on actual findings of the respective
groups.

second diversification. Also we find lineages within Hymenoptera and


Diptera that switched back again to parasitic lifestyles. This parallel type of
pattern in Hymenoptera and Diptera might be superficial, yet it would be
an interesting candidate pair for a closer comparison, including fossil data.
This must remain a task for future approaches. An overview of the origin
of the discussed species is given in Figure 7.

10. OUTLOOK
With the given examples, we provide direct and indirect evidence for
palaeoparasitology concerning insects. Regrettably most fossil examples for
palaeoparasitism only provide indirect evidence for parasitism (pathological
changes of the host, functional morphology of the isolated parasite, phylo-
genetic inference) and these are more difficult to compare to their extant
counterparts. Nevertheless, these fossils have the potential to reveal impor-
tant aspects of the origins and diversification of parasitic lifestyles in various
evolutionary lineages. But even with all these fossil examples, for which
there is a much better record than in many other metazoan groups, we
184 Christina Nagler and Joachim T. Haug

only get a phenomenological insight. To understand the evolution, biology,


ecology and morphology of these fossil parasitic insects, we have to combine
several methods, e.g. phylogenetics, biogeographics, taphonomy, geochem-
ical studies, molecular studies, fossil record and extant species.
Fossils provide a snapshot of the historical biogeography, evolutionary
history and morphological diversification. They also provide their minimum
age. While molecular data are generally thought to help to develop an
evolutionary scenario only fossils allow a true look into former times. Yet,
this is not a one-way road, also knowledge on extant forms is necessary to
better understand fossil specimens. Only holistic approaches with a mature
concept of reciprocal illumination between fossil and extant knowledge
will provide a more detailed understanding of the evolutionary processes
leading to the modern forms.
To tell their ‘true’ story (or at least to come closer to it) more and
more fossils are needed, especially in groups that are considered to have
a sparse fossil record. Often parasitic insects would not be recognized
due to their size or lifestyle, but it is important to find these fossils.
Much of the direct evidence for parasitism is found in amber, because
this resin preserves specimens with microscopic details. Thus, the smallest
mite can be visible on an insect. But also fine-grained sedimentary rocks
preserve fossil specimens very well. Unfortunately ectoparasites may well
be removed during the preparation process, often because they were not
recognized as parasites.
With the aid of additional fossils, we can hope to identify more interme-
diate forms between fossils and extant species or between different lifestyles,
we are able to answer questions of the functional morphology, their ecol-
ogy, their environmental impact, their evolution with their minimum
age, radiations and extinctions and the changing of their lifestyle from free
living to parasites to vectors.

ACKNOWLEDGEMENTS
Foremost we thank the editors for the invitation to contribute to this volume. CN is funded
by the Studienstiftung des deutschen Volkes with a PhD fellowship. JTH is kindly funded by
the German Research Foundation (DFG) under Ha 6300/3-1. Both authors thank J. M.
Starck, Munich, for his support. We thank all people involved in providing free and low-
cost software, such as OpenOffice, CombineZM, CombineZP, Image Analyzer.

REFERENCES
Aguiar, N.O., B€uhrnheim, P.F., 1998. Phoretic pseudoscorpions associated with flying in-
sects in Brazilian Amazonia. J. Arachnol. 26, 451e459.
Insects as Parasites and Hosts 185

Alonso, J., Arillo, A., Barron, E., Corral, J.C., Grimalt, J., L
opez, J.F., L
opez, R., Martínez-
Delclos, X., Ortu~ no, V., Pe~
nalver, E., Trincao, P.R., 2000. A new fossil resin with bio-
logical inclusions in Aptian deposits from the Sierra de Cantabria (Alava, Northern Spain,
Basque-Cantabrian Basin). J. Paleontol. 74, 158e178.
Andrade Filho, J.D., Brazil, R.P., 2003. Relationship of new world phlebotomine sand flies
(Diptera: Psychodidae) based on fossil evidence. Mem. Inst. Oswaldo Cruz 98, 145e149.
Andrade Filho, J.D., Brazil, R.P., Falcao, A.L., Bianchi Galati, E.A., 2007. Description of
Pintomyia (Pifanomyia) paleotrichia, a miocene period new species from the Dominican
Republic (Diptera: Psychodidae; Phlebotominae). Mem. Inst. Oswaldo Cruz 102,
901e903.
Andrade Filho, J.D., Bianchi Galati, E.A., Falcao, A.L., Brazil, R.P., 2008. Description of
Micropygomyia brandaoi sp. n. (Diptera: Psychodidae: Phlebotominae), a fossil phleboto-
mine from the Dominican Republic. Mem. Inst. Oswaldo Cruz 103, 344e346.
Andrade Filho, J.D., Bianchi Galati, E.A., Brazil, R.P., 2009. Review of American fossil
Phlebotominae (Diptera: Psychodidae) with a description of two new species. J. Med.
Entomol. 40, 969e979.
Ansorge, J., 1994. Tanyderidae and Psychodidae (Insecta: Diptera) from lower Jurassic of
northeastern Germany. Pal€aontol. Zeitschrift 68, 199e210.
Ansorge, J., 1996. Insekten aus dem oberen Lias von Grimmen (Vorpommern,
Norddeutschland). Neue Pal€aontol. Abhandl. 2, 1e132.
Ansorge, J., Schl€ uter, T., 1990. The earliest chrysopid: Liassochrysa stigmatica n.g., n.sp. from
the lower Jurassic of Dobbertin, Germany. Neuroptera International 6, 87e93.
Antropov, A.V., 2000. Digger wasps (Hymenoptera, Sphecidae) in Burmese amber. Bull.
Nat. Hist. Mus. 56, 59e77.
Antropov, A.V., Pulawski, W.J., 1996. Pison antiquum, a new species from Dominican amber
(Hymenoptera: Sphecidae). J. Hymenopt. Res. 5, 16e21.
Arillo, A., 2007. Paleoethology: fossilized behaviours in amber. Geol. Acta 5, 159e166.
Asp€ock, U., Asp€ ock, H., 2008. Phylogenetic relevance of the genital sclerites of Neuropter-
ida (Insecta: Holometabola). Syst. Entomol. 33, 97e127.
Azar, D., 2007. Preservation and accumulation of biological inclusions in Lebanese amber
and their significance. Comptes Rendus Palevol 6, 151e156.
Azar, D., Nel, A., 2003. Fossil psychodoid flies and their relation to parasitic diseases. Mem.
Inst. Oswaldo Cruz 98, 35e37.
Azar, D., Nel, A., 2010. Two new non-biting midges from the early Cretaceous Lebanese
amber (Diptera: Chironomidae). Ann. Soc. Entomo. Fr. 46, 198e203.
Azar, D., Nel, A., 2012. Evolution of hematophagy in “non-biting midges” (Diptera:
Chironomidae). Terr. Arthropod Rev. 5, 15e34.
Azar, D., Salame, Y., 2015. A new genus of sycoracinae (Diptera: Psychodidae) from upper
Cretaceous amber of New Jersey. Cretac. Res. 52, 539e647.
Azar, D., Veltz, I., Nel, A., 2008. Mandibulate chironomids: primitive or derived? (Diptera:
Chironomidae). Syst. Entomol. 33, 688e699.
Azar, D., Huang, D., Cai, C., Nel, A., 2015a. The earliest records of pachytroctid booklice
from Lebanese and Burmese Cretaceous ambers (Psocodea, Troctomorpha, Nanopsoce-
tae, Pachytroctidae). Cretac. Res. 52, 336e347.
Azar, D., Mouwad, R., Salame, Y., 2015b. A new genus of Trichomyiinae (Diptera: Psycho-
didae) from upper Cretaceous amber of New Jersey. Cretac. Res. 52, 531e538.
Barling, N., Heads, S.W., Martill, D.M., 2013. A new parasitoid wasp (Hymenoptera: Chal-
cidoidea) from the lower Cretaceous crato formation of Brazil: the first mesozoic
pteromalidae. Cretac. Res. 45, 258e264.
Bethoux, O., 2009. The oldest beetle identified. J. Paleontol. 83, 931e937.
Beutel, R.G., Friedrich, F., H€ ornschemeyer, T., Pohl, H., H€ unefeld, F., Beckmann, F.,
Meier, R., Misof, B., Michael, F., Whiting, M.F., Vilhelmsen, L., 2011. Morphological
186 Christina Nagler and Joachim T. Haug

and molecular evidence converge upon a robust phylogeny of the megadiverse


holometabola. Cladistics 27, 341e355.
Beutel, R.G., Friedrich, F., Ge, S.Q., Yang, X.K., 2014. Insect Morphology and Phylogeny.
Walter de Gruyter, Berlin.
Blagoderov, V., Grimaldi, D.A., Fraser, N.C., 2007. How time flies for flies: diverse Diptera
from the Triassic of Virginia and early radiation of the order. Am. Mus. Novitates 3572,
1e39.
Bleidorn, C., Schmidt-Rhaesa, A., Garey, J.R., 2002. Systematic relationship of Nemato-
morpha based on molecular and morphological data. Invertebr. Biol. 121, 357e364.
Bologna, M.A., Oliverio, M., Pitzalis, M., Mariottini, P., 2008. Phylogeny and evolutionary
history of the blister beetles (Coleoptera, Meloidae). Mol. Phylogenet. Evol. 48, 679e693.
Borkent, A., 2000. Biting midges (Ceratopogonidae: Diptera) from lower Cretaceous Leb-
anese amber with a discussion of the diversity and patterns found in other ambers. In:
Grimaldi, D.A. (Ed.), Studies on Fossils in Amber, with Particular Reference to the
Cretaceous of New Jersey. Backhuys Publishers, Leiden, pp. 355e451.
Borkent, A., 2001. Leptoconops (Diptera: Ceratopogonidae): the earliest extant lineage of
biting midge, discovered in 120e122 million-year-old Lebanese amber. Am. Mus. Nov-
itates 3328, 1e11.
Borkent, A., 2008. The frog-biting midges of the world (Corethrellidae: Diptera). Zootaxa
1804, 1e456.
Borkent, A., 2014. World catalog of extant and fossil Corethrellidae (Diptera). Zootaxa 3796,
435e468.
Borkent, A., Grimaldi, D.A., 2004. The earliest fossil mosquito (Diptera: Culicidae), in mid-
Cretaceous amber. Ann. Entomol. Soc. Am. 97, 882e888.
Borkent, A., Craig, D.A., 2004. Austroconops Wirth and Lee, a lower Cretaceous genus of
biting midges yet living in Western Australia: a new species, first description of the im-
matures and discussion of their biology and phylogeny (Diptera: Ceratopogonidae). Am.
Mus. Novitates 3449, 1e67.
Borkent, A., Coram, R.A., Jarzembowski, E.A., 2013. The oldest fossil biting midge
(Diptera: Ceratopogonidae) from the purbeck Limestone group (lower Cretaceous) of
Southern Great Britain. Pol. J. Entomol. 82, 273e279.
Boucot, A.J., Poinar Jr., G.O., 2010. Fossil Behavior Compendium. CRC Press, New York.
Boussau, B., Walton, Z., Delgado, J.A., Collantes, F., Beani, L., Stewart, I.J., Cameron, S.A.,
Whitfield, J.B., Johnston, J.S., Holland, P.W.H., Bachtrog, D., Kathirithamby, J.,
Huelsenbeck, J.P., 2014. Strepsiptera, phylogenomics and the long branch attraction
problem. PLoS One 9, e1077079.
Brand~ao, C.R.F., Martins-Neto, R.G., Vulcano, M.A., 1990. The earliest known fossil ant
(first southern hemisphere Mesozoic record) (Hymenoptera: Formicidae: Myrmeciinae).
Psyche 96, 195e208.
Bravo, F., Pohl, H., Silva-Neto, A., Beutel, R.G., 2009. Bahiaxenidae, a “living fossil” and a
new family of Strepsiptera (Hexapoda) discovered in Brazil. Cladistics 25, 614e623.
Brazil, R.P., Andrade Filho, J.D., 2002. Description of Pintomyia (Pifanomyia) falcaorum sp.n.
(Diptera: Psychodidae: Phlebotominae), a fossil sand fly from Dominican amber. Memo-
rias do Inst. Oswaldo Cruz 97, 501e503.
Briggs, D.E., 2013. A mosquito’s last supper reminds us not to underestimate the fossil record.
Proc. Natl. Acad. Sci. 110, 18353e18354.
Capinera, J.L., 2003. Striped blister beetle, Epicauta vittata (Fabricius) (Coleoptera: Meloidae).
In: Featured Creatures. University of Florida.
Carvalho, M.R., Wilf, P., Barrios, H., Windsor, D.M., Currano, E.D., Labandeira, C.C.,
Jaramillo, C.A., 2014. Insect leaf-chewing damage tracks herbivore richness in modern
and ancient forests. PLoS One 9, e94950.
Insects as Parasites and Hosts 187

Castellani, C., Maas, A., Waloszek, D., Haug, J.T., 2011. New pentastomids from the late
Cambrian of Sweden e deeper insight of the ontogeny of fossil tongue worms. Palae-
ontogr. Abt. A 293, 95e145.
Cerveau, N., Leclercq, S., Leroy, E., Bouchon, D., Cordaux, R., 2011. Short- and long-term
evolutionary dynamics of bacterial insertion sequence: insight from Wolbachia
endosymbionts. Genome Biol. Evol. 3, 1175e1186.
Chen, J., Wang, B., Engel, M.S., Wappler, T., Jarzembowski, E.A., Zhang, H., Wang, X.,
Zheng, X., Rust, R., 2014. Extreme adaptations for aquatic ectoparasitism in a Jurassic
fly larva. eLife 3, e02844.
Chmielewski, W., 1977. Wyniki obserwacji powi1za~ n roztoczy z owadami [Results of
observations on associations of mites with insects (Acari-Insecta)] Pol. Pismo Entomol.
47, 59e78.
Choufani, J., Perrichot, V., Azar, D., Nel, A., 2014. New biting midges (Diptera: Ceratopo-
gonidae) in late Cretaceous Vendean amber. Paleontol. Contrib. 10H, 34e40.
Choufani, J., Perrichot, V., Girard, V., Garrouste, R., Azar, D., Nel, A., 2013. Two new
biting midges of the modern type from Santonian amber of France (Diptera: Ceratopo-
gonidae). In: Azar, D., Engel, M., Jarzembowski, E., Krogmann, L., Nel, A., Santiago-
Blay, J. (Eds.), Insect Evolution in an Amberiferous and Stone Alphabet: Proceedings of
the 6th International Congress on Fossil Insects, Arthropods and Amber. Brill, Leiden,
pp. 73e95.
Cockerell, T.D.A., 1920. Fossil arthropods in the British museum I. Ann. Mag. Nat. Hist. 9,
273e279.
Cockerell, T.D.A., 1921. Fossil arthropods in the British museum VI. Ann. Mag. Nat. Hist.
9, 453e480.
Cook, J.L., 2014. A new Species of Stichotrema Hofeneder (Strepsiptera: Myrmecolacidae)
from the Dominican Republic, with notes on new world member of the genus.
J. Kans. Entomol. Soc. 87, 66e73.
Craig, D.A., Craig, R.E.G., Crosby, T.K., 2012. Simuliidae (Insecta: Diptera). In:
Emberson, R.M., Fletcher, M.J., Hoare, J.B., Lariviere, M.C., Palma, R.L. (Eds.), Fauna
of New Zealand, vol. 68. Manaaki Whenua Press, Lincoln, pp. 1e336.
Currie, D.C., Grimaldi, D.A., 2000. A new black fly (Diptera: Simuliidae) genus from mid
Cretaceous (Turonian) amber of New Jersey. In: Grimaldi, D.A. (Ed.), Studies on Fossils
in Amber with Particular Reference to the Cretaceous of New Jersey. Backhuys Pub-
lishers, Leiden, pp. 473e485.
Czosnek, H., Ghanim, M., Morin, S., Rubinstein, G., Fridman, V., Zeidan, M., 2001.
Whiteflies: vectors and victims (?) of Geminiviruses. Adv. Virus Res. 57, 291e322.
Dabert, M., Witalinski, W., Kazmierski, A., Olszanowski, Z., Dabert, J., 2010. Molecular
phylogeny of acariform mites (Acari, Arachnida): strong conflict between phylogenetic
signal and long-branch artifacts. Mol. Phylogenet. Evol. 56, 222e241.
Dalgleish, R.C., Palma, R.L., Price, R.D., Smith, V.S., 2006. Fossil lice (Insecta: phthirap-
tera) reconsidered. Syst. Entomol. 31, 648e651.
Del-Claro, K., Tizo-Pedroso, E., 2009. Ecological and evolutionary pathways of social
behavior in Pseudoscorpions (Arachnida: Pseudoscorpiones). Acta Ethol. 12, 13e22.
Delclos, X., Arillo, A., Penalver, E., Barr
on, E., Soriano, C., del Valle, R.L., Bernardes, E.,
Corral, C., Ortuno, V.M., 2007. Fossiliferous amber deposits from the Cretaceous
(Albian) of Spain. Comptes Rendus Palevol 6, 135e149.
De Silva, P., Bernal, X.E., 2013. First report of mating behavior of a species of frog biting
midge (Diptera: Corethrellidae). Fla. Entomol. 96, 1522e1529.
Dlussky, G.M., Rasnitsyn, A.P., 2003. Ants (Hymenoptera: Formicidae) of formation Green
river and some other middle eocene deposits of North America. Russ. J. Entomol. 11,
411e436.
188 Christina Nagler and Joachim T. Haug

Donovan, M.P., Wilf, P., Labandeira, C.C., Johnson, K.R., Peppe, D.J., 2014. Novel insect
leaf-mining after the end-Cretaceous extinction and the demise of Cretaceous leaf
miners, Great Plains, USA. PLoS One 9, e103542.
Dunlop, J.A., 2007. A large parasitengonid mite (Acari, Erythraeoidea) from the early Creta-
ceous Crato formation of Brazil. Mitt. Mus. Naturkund. Berlin 10, 91e98.
Dunlop, J.A., 2010. Geological history and phylogeny of Chelicerata. Arthropod Struct. Dev.
39, 124e142.
Dunlop, J.A., Penney, D., 2012. Fossil Arachnids. Siri Scientific Press, Manchester.
Dunlop, J.A., Wirth, S., Penney, D., McNeil, A., Bradley, R.S., Withers, P.J., Preziosi, R.F.,
2012. A minute fossil phoretic mite recovered by phasecontrast X-ray computed
tomography. Biol. Lett. 8, 457e460.
Dunlop, J.A., Kontschan, J., Walter, D.R., Perrichot, V., 2014. An ant-associated mesostig-
matid mite in Baltic amber. Biol. Lett. 10, 20140531.
Duron, O., Schneppat, U.E., Berthomieu, A., Goodman, S.M., Droz, B., Paupy, E.,
Nkoghe, J.O., Rahola, N., Tortosa, P., 2014. Origin, acquisition and diversification
of heritable bacterial endosymbionts in louse flies and bat flies. Mol. Ecol. 23,
2105e2117.
Engel, M.S., 2004. Thorny lacewings (Neuroptera: Rachiberothidae) in Cretaceous amber
from Myanmar. J. Syst. Palaeontol. 2, 137e140.
Engel, M.S., 2005. An Eocene ectoparasite of bees: the oldest definitive record of phoretic
meloid triungulins (Coleoptera: Meloidae; Hymenoptera: Megachilidae). Acta Zool.
Cracov. 48, 43e48.
Engel, M.S., 2013a. A new genus and species of Baissidae in late Cretaceous amber from New
Jersey (Hymenoptera: Evanioidea). Novitates Paleoentomologicae 3, 1e8.
Engel, M.S., 2013b. A ceraphronid wasp in early Miocene amber from the Dominican
Republic (Hymenoptera: Ceraphronidae). Novitates Paleoentomol. 2, 1e6.
Engel, M.S., Grimaldi, D.A., 2004. New light shed on the oldest insect. Nature 427, 627e630.
Engel, M.S., Grimaldi, D.A., 2007. The neuropterid fauna of dominican and mexican amber
(Neuropterida: Megaloptera, neuroptera). Am. Mus. Novitates 3587, 1e58.
Engel, M.S., Perrichot, V., 2014. The extinct wasp family Serphitidae in late Cretaceous
Vendean amber (Hymenoptera). Paleontol. Contrib. 10, 46e51.
Engel, M.S., Grimaldi, D.A., Ortega-Blanco, J., 2013a. A stephanid wasp in mid-Cretaceous
Burmese amber (Hymenoptera: Stephanidae), with comments on the antiquity of the
Hymenopteran radiation. J. Kans. Entomol. Soc. 86, 244e252.
Engel, M.S., Ortega-Blanco, J., McKellar, R.C., 2013b. New scolebythid wasps in Creta-
ceous amber from Spain and Canada, with implications for the phylogeny of the family
(Hymenoptera: Scolebythidae). Cretac. Res. 46, 31e42.
Fraser, N.C., Grimaldi, D.A., Olsen, P.E., Axsmith, B., 1996. A triassic lagerst€atte from
Eastern North America. Nature 380, 615e619.
Friedrich, F., Beutel, R.G., 2010. The thoracic morphology of Nannochorista (Nannochoris-
tidae) and its implications for the phylogeny of Mecoptera and Antliophora. J. Zoolog.
Syst. Evol. Res. 48, 50e74.
Gall, J.C., Grauvogel, L., 1966. Faune d
u Buntsandstein 1. Pontes d’invertébres du Buntsand-
stein supérieur. Ann. Paléontol. Invertébr. 52, 155e161.
Gao, T., Shih, C., Xu, X., Wang, S., Ren, D., 2012. Mid-Mesozoic flea-like ectoparasites of
feathered or haired vertebrates. Curr. Biol. 22, 731e735.
Gao, T., Shih, C., Rasnitsyn, A.P., Xu, X., Wang, S., Ren, D., 2013a. New transitional fleas
from China highlighting diversity of early Cretaceous ectoparasitic Insects. Curr. Biol.
23, 1261e1266.
Gao, T., Shih, C., Rasnitsyn, A.P., Ren, D., 2013b. Hoplitolyda duolunica gen.et sp.nov.
(Insecta, Hymenoptera, Praesiricidae), the hitherto largest sawfly from the Mesozoic of
China. Plos One 8, e62420.
Insects as Parasites and Hosts 189

Garrouste, R., Clement, G., Nel, P., Engel, M.S., Grnadcolas, P., D’Haese, C., Lagebro, L.,
Denayer, J., Gueriau, P., Lafaite, P., Olive, S., Prestianni, C., Nel, A., 2012. A complete
insect from the late Devonian period. Nature 488, 82e85.
Garwood, R.J., Dunlop, J., 2014. Three-dimensional reconstruction and the phylogeny of
extinct chelicerate orders. PeerJ 2, e641.
Graczyk, T.K., Knight, R., Tamang, L., 2005. Mechanical transmission of human protozoan
parasites by insects. Clin. Microbiol. Rev. 18, 128e132.
Greenwalt, D.E., Goreva, Y.S., Siljestr€om, S.M., Rose, T., Harbach, R.E., 2013. Hemoglo-
bin-derived porphyrins preserved in a middle Eocene blood-engorged mosquito. Proc.
Natl. Acad. Sci. U.S.A. 110, 18496e18500.
Grimaldi, D.A., 1996. Amber: Window to the Past. Harry N. Abrams Incorporation,
New York.
Grimaldi, D.A., 2000. A diverse fauna of Neuropterodea in amber from the Cretaceous of
New Jersey. In: Grimaldi, D.A. (Ed.), Studies on Fossil in Amber, with Particular Refer-
ence to the Cretaceous of New Jersey. Backhuys Publishers, Leiden, pp. 259e303.
Grimaldi, D.A., Engel, M.S., 2005. Evolution of the Insects. Cambridge University Press,
Cambridge.
Grimaldi, D.A., Engel, M.S., 2006. Fossil Liposcelididae and the lice ages (Insecta: Psocodea).
Proc. R. Soc. B 273, 625e633.
Grimaldi, D.A., Engel, M.S., Nascimbene, P.C., 2002. Fossiliferous Cretaceous amber from
Myanmar (Burma): Its rediscovery, biotic diversity, and paleontological significance. Am.
Mus. Novitates 3361, 1e71.
Grimaldi, D.A., Arillo, A., Cumming, J., Hauser, M., 2011. Brachyceran Diptera (Insecta)
in Cretaceous ambers, part IV, significant new Orthorrhaphous taxa. ZooKeys 148,
293e332.
Guglielmino, A., Olmi, M., 2011. Revision of fossil species of Deinodryinus, with description
of a new species (Hymenoptera, Dryinidae). ZooKeys 130, 495e504.
Harbach, R.E., Greenwalt, D., 2012. Two Eocene species of Culiseta (Diptera: Culicidae)
from the Kishenehn formation in Montana. Zootaxa 3530, 25e34.
Haring, E., Asp€ock, U., 2004. Phylogeny of Neuropterida: a first molecular approach. Syst.
Entomol. 29, 415e430.
Harvey, M.S., 2013. Order Pseudoscorpiones. Zootaxa 3703, 34e35.
Harvey, T.H.P., Dong, X., Donoghue, P.C.J., 2010. Are palaeoscolecids ancestral
ecdysozoans? Evol. Dev. 12, 177e200.
Henderickx, H., Cnudde, V., Masschaele, B., Dierick, M., Vlassenbroeck, J.,
Hoorebekevan, L., 2006. Description of a new fossil Pseudogarypus (Pseudoscorpiones:
Pseudogarypidae) with the use of X-ray micro-CT to penetrate opaque amber. Zootaxa
1305, 41e50.
Heraty, J.M., Darling, C., 2009. Fossil Eucharitidae and Perilampidae (Hymenoptera: chal-
cidoidea) from Baltic amber. Zootaxa 2306, 1e16.
Heraty, J.M., Ronquist, F., Carpenter, J.M., Hawks, D., Schulmeister, S., Dowling, A.P.,
Murray, D., Munro, J., Wheeler, W.C., Schiff, N., Sharkey, M.J., 2011. Evolution of
the hymenopteran megaradiation. Mol. Phylogenet. Evol. 60, 73e88.
Huang, D., 2015. Tarwinia australis (Siphonaptera: Tarwiniidae) from the lower Cretaceous
Koonwarra fossil bed: morphological revision and analysis of its evolutionary
relationship. Cretac. Res. 52, 507e515.
Huang, D., Nel, A., Cai, C., Lin, Q., Engel, M.S., 2013. Amphibious flies and paedomor-
phism in the Jurassic period. Nature 495, 94e97.
Huang, D., Engel, M.S., Cai, C., Wu, H., Nel, A., 2012. Diverse transitional giant fleas from
the Mesozoic era of China. Nature 483, 201e204.
Hulden, L., Hulden, L., 2014. Checklist of the family Culicidae (Diptera) in Finland.
ZooKeys 441, 47e51.
190 Christina Nagler and Joachim T. Haug

H€ornschemeyer, T., Haug, J.T., Bethoux, O., Beutel, R.G., Charbonnier, S., Hegna, T.A.,
Koch, M., Rust, J., Wedmann, S., Bradler, S., Willmann, R., 2013. Is strudiella a Devo-
nian insect? Nature 494, 82e85.
Ibanez-Bernal, S., Kraemer, M.S., Stebner, F., Wagner, R., 2014. A new fossil species of
Phlebotominae sand fly from Miocene amber of Chiapas, Mexico (Diptera:
Psychodidae). Pal€aontol. Z. 88, 227e233.
Jarzembowski, E.A., Azar, D., Nel, A., 2008. A new chironomid (Insecta: Diptera)
from Wealden amber (lower Cretaceous) of the Isle of Wight (UK). Geol. Acta 6,
285e291.
Jell, P.A., Duncan, P.M., 1986. Invertebrates, mainly insects, from the freshwater, lower
Cretaceous, koonwarra fossil bed (Korumburra group), South Gippsland, Victoria.
Mem. Assoc. Australas. Palaeontol. 3, 111e205.
Jennings, J.T., Krogmann, L., Mew, S.L., 2013. Cretevania bechlyi sp. nov., from Cretaceous
Burmese amber (Hymenoptera: Evaniidae). Zootaxa 3609, 91e95.
Johnson, K.P., Clayton, D.H., 2003. The biology, ecology and evolution of chewing lice.
Syst. Biol. 53, 449e476.
Johnson, B.A., Borowiec, M.L., Chiu, J.C., Lee, E.K., Atallah, J., Ward, P.S., 2013. Phylo-
genomics resolves evolutionary relationships among ants, bees, and wasps. Curr. Biol. 23,
2058e2062.
Judson, M.L.I., 1998. A sternophorid Pseudoscorpion (Chelonethi) in Dominican Amber,
with remarks on the family. J. Arachnol. 26, 419e428.
Judson, M.L.I., 2000. Electrobisium acutum Cockerell, a cheiridiid pseudoscorpion from Bur-
mese amber, with remarks on the validity of the Cheiridioidea (Arachnida, Chelonethi).
Bull. Nat. Hist. Mus. Lond. 56, 79e83.
Judson, M.L.I., 2004. Baltic amber fossil of Garypinus electri Beier provides first evidence of
phoresy in the pseudoscorpion family Garypinidae (Arachnida: Chelonethi). In:
Logunov, D.V., Penney, D. (Eds.), European Arachnology 2003. KMK Scientific Press,
Moscow, pp. 127e131.
Judson, M.L.I., 2009. Cheliferoid pseudoscorpions (Arachnida, Chelonethi) from the lower
Cretaceous of France. Geodiversitas 31, 61e71.
Judson, M.L.I., 2010. Redescription of Chelifer eucarpus Dalman (Arachnida, Chelonethi,
Withiidae) and first records of pseudoscorpions in copal from Madagascar and
Colombia. Paleodiversity 3, 33e42.
Judson, M.L.I., Maakol, J., 2009. A mite of the family Tanaupodidae (Arachnida, Acari, Para-
sitengona) from the lower Cretaceous of France. Geodiversitas 31, 41e47.
Kalugina, N.S., 1974. Changes in the subfamily composition of chironomids (Diptera, Chi-
ronomidae) as indicator of a possible eutrophication of bodies of water during the late
Mesozoic. Bulleten’ Mosk. Obshch. Ispyt. Prir. Biol. 6, 45e56.
Kalugina, N.S., 1976. Non-biting midges of the subfamily Diamesinae (Diptera: Chironomi-
dae) from the upper Cretaceous of the Taimyr. Paleontol. Zhurnal 1, 87e93.
Kalugina, N.S., 1980. Cretaceous Aphroteniinae from North siberia (Diptera, chiro-
nomidae), Electrotenia brundini gen. nov., sp. nov. Acta Univ. Carol. Biol. 1978,
89e93.
Kalugina, N.S., 1993. Chaoborids and non-biting midges from the mesozoic of Eastern
Transbaikalia (Diptera: Chaoboridae and Chironomidae). Trudy Paleontol. Instituta
252, 117e139.
Kalugina, N.S., Kovalev, V.G., 1985. Jurassic Diptera of Siberia. Nauka, Moscow, pp. 198 (in
Russian).
Karolyi, F., Colville, J.F., Handschuh, S., Metscher, B.D., Krenn, H.W., 2014. One probos-
cis, two tasks: adaptations to blood-feeding and nectar-extracting in long-proboscid
horse flies (Tabanidae, Philoliche). Arthropod Struct. Dev. 43, 403e413.
Insects as Parasites and Hosts 191

Kathirithamby, J., 2009. Host-parasitoid associations in Strepsiptera. Annu. Rev. Entomol.


54, 227e249.
Kathirithamby, J., Grimaldi, D.A., 1993. Remarkable stasis in some lower tertiary parasitoids:
descriptions, new records and review of Strepsiptera in the Oligo-Miocene amber of the
Dominican Republic. Entomol. Scand. 24, 31e41.
Kathirithamby, J., Henderickx, H., 2008. First record of the Strepsiptera genus Caenocholax in
Baltic amber with the description of a new species. Phegea 36, 149e156.
Kathirithamby, J., Hayward, A., McMahon, D.P., Ferreira, R.S., Andreazze, R., Tadeu de
Almeida-Andrade, H., Fresneau, D., 2010. Conspecifics of a heterotrophic heterono-
mous species of Strepsiptera (Insecta) are matched by molecular characterization. Syst.
Entomol. 35, 234e242.
Kerr, P.H., 2010. Phylogeny and classification of Rhagionidae, with implications for
Tabanomorpha (Diptera: Brachycera). Zootaxa 2592, 1e133.
Kinzelbach, R.K., 1979. Das erste neotropische Fossil der F€acherfl€ ugler (Stuttgarter
Bernsteinsammlung: Insecta, Strepsiptera). Stuttg. Beitr. Naturkd. B 52, 1e14.
Kinzelbach, R.K., 1983. F€acherfl€ugler aus dem dominikanischen Bernstein (Insecta, Strepsiptera:
Myrmecolacidae). Verhandlungen Naturwissenschaftlichen Vereins Hambg. 26, 29e36.
Kinzelbach, R.K., Lutz, H., 1985. Stylopid larva from the Eocene: a spotlight on the
phylogeny of the stylopids (Strepsiptera). Ann. Entomol. Soc. Am. 78, 600e602.
Kinzelbach, R.K., Pohl, H., 1994. The fossil strepsiptera (Insecta: strepsiptera). Ann.
Entomol. Soc. Am. 87, 59e70.
Klompen, H., Grimaldi, D.A., 2001. First mesozoic record of a parasitiform mite: a larval
argasid tick in Cretaceous amber (Acari: Ixodida: Argasidae). Ann. Entomol. Soc. Am.
94, 10e15, 6.
Klopfstein, S., Vilhelmsen, L., Heraty, J.M., Sharkey, M., Ronquist, F., 2013. The hyme-
nopteran tree of life: evidence from protein-coding genes and objectively aligned
ribosomal data. PLoS One 8, e69344.
Kluth, S., Kruess, A., Tscharntke, T., 2002. Insects as vectors of plant pathogens: mutualistic
and antagonistic interactions. Oecologia 133, 193e199.
Knor, S., Skuhrava, M., Wappler, T., Prokop, J., 2013. Galls and gall makers on plant leaves
from the lower miocene (Burdigalian) of Czech Republic: systematic and palaeoecolog-
ical implications. Rev. Paleobotany Palynology 188, 38e51.
Kogan, M., Poinar Jr., G.O., 2010. A new fossil Stylops (Strepsiptera: Stylopidae) from
Dominican amber. Neotropical Entomol. 39, 227e234.
Kolyada, V., Perkovsky, E., 2011. A new species of the genus Disogmus F€ orster (Hymenop-
tera, Proctotrupoidea, Proctotrupidae) from the Eocene Rovno amber. ZooKeys 130,
455e459.
Kosmowska-Ceranowicz, B., 2001. The Amber Treasure Trove. Polish Academy of
Sciences, Warsaw.
Koteja, J., Poinar Jr., G.O., 2005. Scale insects (Coccinea) associated with mites in the fossil
record. In: X. International Symposium on Scale Insects Studies (April 2004), Adana,
Proceedings, pp. 281e294.
Koutsovoulos, G., Makepeace, B., Tanya, V.N., Blaxter, M., 2014. Palaeosymbiosis revealed
by genomic fossils of wolbachia in a strongyloidean nematode. PLos Genet. 10, e1004397.
Krassilov, V.A., 2007. Mines and galls on fossil leaves from the late Cretaceous of southern
Negev, Israel. Afr. Invertebr. 48, 13e22.
Krenn, H.W., Asp€ ock, H., 2012. Form, function and evolution of the mouthparts of blood-
feeding Arthropoda. Arthropod Struct. Dev. 41, 101e118.
Krogmann, L., 2013. First fossil record of cerocephaline wasps with a description of a new
genus and species from Dominican amber (Hymenoptera: Chalcidoidea: Pteromalidae:
Cerocephalinae). Hist. Biol. 25, 43e49.
192 Christina Nagler and Joachim T. Haug

Krogmann, L., Nel, A., 2012. On the edge of parasitoidism: a new lower Cretaceous wood-
wasp forming the putative sister group of Xiphydriidae þ Euhymenoptera. Syst. Ento-
mol. 37, 215e222.
Krzeminski, W., Krzeminska, E., 2003. Triassic Diptera: descriptions, revision and phyloge-
netic relations. Acta Zool. Cracov. 46, 153e184.
Kukalova-Peck, J., Beutel, R.G., 2012. Is the Carboniferous yAdiphlebia lacoana really the
“oldest beetle”? Critical reassessment and description of a new Permian beetle family.
Eur. J. Entomol. 109, 633e645.
Kulika, R., 1979. Mengea mengei sp. n. from the Baltic amber. Pr. Mus. Ziemi-Warszawa 32,
109e112.
Kulika, R., 2001. New genera and species of Strepsiptera from the Baltic amber. Pr. Mus.
Ziemi-Warszawa 46, 3e16.
Kumar, P., 2001. Antiquity of Phthiraptera: fossil evidence. J. Palaeontol. Soc. India 49,
159e168.
Kumar, P., Kumar, P., 2001. Phthirapteran insect and Acanthocephala from late Triassic sed-
iments of the Satpura Basin India. J. Palaeontol. Soc. India 46, 141e146.
Labandeira, C.C., 2002. Paleobiology of predators, parasitoids, and parasites: death and accom-
modation in the fossil record of continental invertebrates. Paleontol. Soc. Pap. 8, 211e250.
Labandeira, C., 2006. Silurian to Triassic plant and hexapod clades and their associations: new
data, a review, and interpretations. Arthropod Syst. Phylo. 64, 53e94.
Labandeira, C.C., 2010. The pollination of mid Mesozoic seed plants and the early history of
long-proboscid insects. Ann. Missouri Bot. Gard. 97, 469e513.
Labandeira, C.C., 2011. Evidence for an earliest late carboniferous divergence time and the
early larval ecology and diversification of major holometabola lineages. Entomol. Am.
117, 9e21.
Labandeira, C.C., 2013. A paleobiologic perspective on plant-insect interactions. Curr.
Opin. Plant Biol. 16, 414e421.
Labandeira, C.C., Currano, E.D., 2013. The fossil record of plant-insect dynamics. Annu.
Rev. Earth Planet. Sci. 41, 287e311.
Labandeira, C.C., Dilcher, D.L., Davis, D.R., Wagner, D.L., 1994. Ninety-seven million
years of angiosperm-insect association: paleobiological insights into the meaning of
coevolution. Proc. Natl. Acad. Sci. U.S.A. 91, 12278e12282.
Lambkin, C., Sinclair, B.J., Pape, T., Courtney, G., Skevington, J.H., Meier, R.,
Yeates, D.K., Blagoderov, V., Wiegmann, B., 2013. The phylogenetic relationships
among infraorders and superfamilies of Diptera based on morphological evidence.
Syst. Entomol. 38, 164e179.
Lara, M.B., Rasnitsyn, A.P., Zavattieric, A.M., 2014. Potrerilloxyela menendezi gen. et sp. nov.
from the late Triassic of Argentina: the oldest representative of Xylelidae (Hymenoptera:
Symphyta) for Americas. Paleontol. J. 42, 182e190.
Larsson, S.G., 1978. Baltic Amber e a paleobiological study. Entomonograph 1, 1e192.
Lewis, R.E., Grimaldi, D.A., 1997. A pulicid flea in miocene amber from the Dominican
Republic (Insecta: Siphonaptera: Pulicidae). Am. Mus. Novitates 3205, 1e9.
Li, L., Shih, C., Ren, D., 2013. Two new wasps (Hymenoptera: Stephanoidea: Ephialtitidae)
from the middle Jurassic of China. Acta Geol. Sin. 87, 1486e1494.
Li, L., Shih, C., Ren, D., 2014. New fossil Praeaulacinae wasps (Insect: Hymenoptera: Evan-
ioidea: Praeaulacidae) from the middle Jurassic of China. Zootaxa 3814, 432e442.
Li, M., Tian, Y., Zhao, Y., Bu, W., 2012. Higher level phylogeny and the first divergence
time estimation of Heteroptera (Insecta: Hemiptera) based on multiple genes. PLoS One
7, e32152.
Light, J.E., Smith, V.S., Allen, J.M., Durden, L.A., Reed, D.L., 2010. Evolutionary history of
mammalian sucking lice (Phthiraptera: Anoplura). BMC Evol. Biol. 10, 292.
Insects as Parasites and Hosts 193

Litman, J.R., Praz, C.J., Danforth, B.N., Griswold, T.L., Cardinal, S., 2014. Origins, evolu-
tion and diversification of cleptoparasitic lineages in long-tongued bees. Evolution 67,
2982e2998.
Longdon, B., Jiggins, F.M., 2012. Vertically transmitted viral endosymbionts of insects: do
sigma viruses walk alone? Proc. R. Soc. B 279, 3889e3898.
Lopes, A.H., Souto-Padron, T., Dias, F.A., Gomes, M.T., Rodrigues, G.C.,
Zimmermann, L.T., Silva, T.L., Vermelho, A.B., 2010. Trypanosomatids: odd organ-
isms, devastating diseases. Open Parasitol. J. 4, 30e59.
Lukashevich, E.D., Mostovski, M.B., 2003. Hematophagous insects in the fossil record. Pale-
ontol. J. 37, 153e161.
Lukashevich, E.D., Przhiboro, A.A., 2015. A new tribe of Diamesinae (Diptera: Chironomi-
dae) from the Lower Cretaceous of Mongolia. Cretac. Res. 52, 562e569.
Lutz, H., 1990. Systematische und pal€aokologische Untersuchungen an Inseken aus dem
Mittel-Eoz€an der Grube Messel bei Darmstadt. Cour. Forschungsinst. Senckenb. 124,
1e165.
Maas, A., Waloszek, D., Haug, J.T., M€ uller, K.J., 2007. A possible larval roundworm from
the Cambrian “Orsten” and its bearing on the phylogeny of Cycloneuralia. Mem. Assoc.
Australas. Palaeontol. 34, 499e519.
Maas, A., 2013. Gastrotricha, Cycloneuralia and Gnathifera: the fossil record. In: Schmidt-
Rhaesa, A. (Ed.), Handbook of Zoology. Gastrotricha, Cycloneuralia and Gnathifera,
1, pp. 11e28.
Makarkin, V.N., 1990. New neuroptera from the upper Cretaceous of asia. Novosti Faunis-
tiki i Sistematiki 63e68.
Makarkin, V.N., 1996. Fossil neuroptera of the lower Cretaceous of Baissa, east siberia. Part
5. Mantispidae. Russ. Entomol. J. 5, 91e93.
Mans, B.J., de Klerk, D., Pienaar, R., Latif, A.A., 2011. Nuttalliella namaqua: a living fossil and
closest relative to the ancestral tick lineage: implications for the evolution of blood-
feeding in ticks. PLoS One 6, e23675.
Martill, D.M., Davis, P.G., 1998. Did dinosaurs come up to scratch? Nature 396, 528e529.
Mayhew, P.J., 2007. Why are there so many insect species? Perspectives from fossils and
phylogenies. Biol. Rev. 82, 425e454.
McKellar, R.C., Engel, M.S., 2011. New stigmaphronidae and Megaspilidae (Hymenoptera:
Ceraphronoidea) from Canadian Cretaceous amber. Cretac. Res. 32, 794e805.
McKellar, R.C., Engel, M.S., 2012. Hymenoptera in Canadian Cretaceous amber (Insecta).
Cretac. Res. 35, 258e279.
McKellar, R.C., Kopylov, D.S., Engel, M.S., 2013. Ichneumonidae (Insecta: Hymenoptera)
in Canadian late Cretaceous amber. Foss. Rec. 16, 217e227.
McKenna, D.D., Farrell, B.D., 2010. 9-genes reinforce the phylogeny of Holometabola and
yield alternate views on the phylogenetic placement of Strepsiptera. PLoS One 5,
e11887.
McMahon, D.P., Hayward, A., Kathirithamy, J., 2011. The first molecular phylogeny of
Strepsiptera (Insecta) reveals an early burst of molecular evolution correlated with the
transition to endoparasitism. PLoS One 6, e21206.
McNaughton, S.J., 1983. Compensatory plant growth as a response to herbivory. Oikos 40,
329e336.

Meng, A., 1866. Uber ein Rhipidopteron und einige Helminthen im Bernstein. Schr. Nat.
Ges. Danz. 2, 1e8.
Mey, E., 2005. Psittacobrosus bechsteini: ein neuer Federling (Insecta, Phthiraptera, Ambly-

cera) vom Dreifarbenara Ara tricolor (Psittaciformes), nebst einer annotierten Ubersicht
€ber fossile und rezent ausgestorbene Tierl€ause. Anz. Des. Ver. Th€
u uring. Ornithol. 5,
202e217.
194 Christina Nagler and Joachim T. Haug

Misof, B., Liu, S., Meusemann, K., Peters, R.S., Donath, A., Mayer, C., Frandsen, P.B.,
Ware, J., Flouri, T., Beutel, R.G., Neihuis, O., Petersen, M., Izquierdo-Carrasco, F.,
Wappler, T., Rust, J., Aberer, A.J., Asp€ ock, U., Asp€
ock, H., Bartel, D., Blanke, A.,
Berger, S., B€ ohm, A., Buckley, T.R., Calcott, B., Chen, J., Fiedrich, F., Fukui, M.,
Fujita, M., Greve, C., Grobe, P., Gu, S., Huang, Y., Jermiin, L.S., Kawahara, A.Y.,
Krogmann, L., Kubiak, M., Lanfear, R., Letsch, H., Li, Y., Li, Z., Li, Y., Lu, H.,
Machida, R., Mashimo, Y., Kapli, R., McKenna, D.D., Meng, G., Nakagaki, Y.,
Navarrete-Heredia, J.L., Ott, M., Ou, Y., Pass, G., Podsiadlowski, L., Pohl, H., von
Reumont, B.M., Sch€ utte, K., Sekiya, K., Shimizu, S., Slipinski, A., Stamatakis, A.,
Song, W., Su, X., Szucsich, N.U., Tan, M., Tang, X., Tang, M., Tang, J.,
Timelthaler, G., Tomizuka, S., Trautwein, M., Tong, X., Uchifune, T., Walzl, M.G.,
Wiegmann, B.M., Wilbrandt, J., Wipfler, B., Wong, T.K.F., Wu, Q., Wu, G.,
Xie, Y., Yang, S., Yang, Q., Yeates, D.K., Yoshizawa, K., Zhang, Q., Zhang, R.,
Zhang, W., Zahng, Y., Zhang, J., Zhao, J., Zhou, C., Zhou, L., Ziesmann, T.,
Zou, S., Li, Y., Xu, X., Zhang, Y., Yang, H., Wang, J., Wang, U., Kjer, K.M.,
Zhou, X., 2014. Phylogenomics resolves the timing and pattern of insect evolution.
Science 346, 763e767.
Mitchell, P.D., 2015. Human parasites in medieval Europe: lifestyle, sanitation and medical
treatment. Adv. Parasitol. 90, 389e420.
M€
ohn, E., 1960. Eine neue Gallm€ ucke aus der niederrheinischen Braunkohle, Sequoiomyia
krauseli n. g., n. sp. (Diptera, Itonididae). Senckenberg. Lethaea 41, 513e522.
Mostovski, M.B., 2008. Contributions to the study of fossil snipe-flies (Diptera: Rhagioni-
dae). The genus Protorhagio. Palaeontol. J. 42, 75e80.
Mostovski, M.B., Jarzembowski, E.A., Coram, R.A., 2003. Horseflies and anthericids
(Diptera: Tabanidae, Athericidae) from the lower Cretaceous of England and
Transbaikalia. Paleontol. J. 37, 162e169.
Nadarasah, G., Stavrinides, J., 2011. Insects as alternative hosts for phytopathogenic bacteria.
FEMS Microbiol. Rev. 35, 555e575.
Nel, A., 1988. Deux nouveaux Mantispidae (Planipennia) fossiles de l’Oligocene du sudest de
la France. Neuroptera Int. 5, 103e109.
Nel, A., de Pl€ oeg, G., 2004. New fossil bee flies (Diptera: Bombyloidea) in the Lowermost
Eocene amber of the Paris Basin. Geologica Acta 2, 57e66.
Nel, A., Waller, A., de Pl€ oeg, G., 2004. An aulacid wasp in the lowermost Eocene amber
from the Paris Basin (Hymenoptera: Aulacidae). Geol. Acta 2, 67e74.
Niehuis, O., Hartig, G., Grath, S., Pohl, H., Lehmann, J., Tafer, H., Donath, A., Krass, V.,
Eisenhardt, C., Hertel, J., Petersen, M., Mayer, C., Meusemann, K., Peters, R.S.,
Stadler, P.F., Beutel, R.G., Bornberg-Bauer, D., McKenna, D.D., Misof, B., 2012.
Genomic and morphological evidence converge to resolve the enigma of Strepsiptera.
Curr. Biol. 22, 1309e1313.
Norton, R.A., Bonamo, P.M., Grierson, J.D., Shear, W.A., 1988. Oribatid mite fossils from a
terrestrial Devonian deposit near Gilboa, New York. J. Paleontol. 62, 259e269.
Ohl, M., 2011. Aboard a spider-a complex developmental strategy fossilized in amber.
Naturwissenschaften 98, 453e456.
Olmi, M., Mita, T., Guglielmino, A., 2014a. Revision of the Embolemidae of Japan (Hyme-
noptera: Chrysidoidea), with description of a new genus and two new species. Zootaxa
3793, 423e440.
Olmi, M., Xu, Z., Guglielmino, A., 2014b. Descriptions of new fossil taxa of Dryinidae (Hy-
menoptera: Chrysidoidea) from Burmese amber (Myanmar). Acta Entomol. Musei Natl.
Pragae 54, 703e714.
Ortega Blanco, J., Singh, H., Engel, M.S., 2012. First amber fossil Rhysipolini (Hymenop-
tera: Braconidae): a new genus and species in Early Eocene Cambay amber. Acta Ento-
mol. Musei Natl. Pragae 52, 585e594.
Insects as Parasites and Hosts 195

Otalora-Lina, F., Pérez-Sanchez, A., Sandoval, C., Aldana, E., 2015. Evolution of hematoph-
agous habit in Triatominae (Heteroptera: Reduviidae). Rev. Chil. Hist. Nat. 88, 1e13.
Panfilov, D.V., 1980. New representatives of lacewings (Neuroptera) from the Jurassic of
Karatau. In: Dolin, V.G., Panfilov, D.V., Ponomarenko, A.G., Pritykina, L.N. (Eds.),
Fossil insects of the Mesozoic. Naukova Dumka, Kiev, pp. 82e111.
Penalver, E., Grimaldi, D.A., 2005. Assemblages of mammalian hair and blood-feeding
midges (Insecta: Diptera: Psychodidae: Phlebotominae) in Miocene amber. Trans. R.
Soc. Edinb. Earth Sci. 96, 177e195.
Pe~
nalver, E., Pérez-de la Fuente, R., 2014. Unearthing the secrets of ancient immature
insects. eLife 3, e03443. http://dx.doi.org/10.7554/eLife.03443.
Perreira, A.I.A., Fadini, M.A.M., Pikart, T.G., Zanuncio, J.C., Serrao, J.E., 2012. New hosts
and parasitism notes for the mite Leptus (Acari: Erythraeidae) in fragments of the Atlantic
Forest, Brazil. Braz. J. Biol. 72, 611e616.
Pepato, A.R., da Rocha, C.E., Dunlop, J.A., 2010. Phylogenetic position of the acariform
mites: sensitivity to homology assessment under total evidence. BMC Evol. Biol. 10,
235.
Pérez-de la Fuente, R., Delclos, X., Penalver, E., Arillo, A., 2011. Biting midges (Diptera:
Ceratopogonidae) from the early Cretaceous El Soplao amber (N Spain). Cretac. Res.
32, 750e761.
Perrichot, V., 2013. New maimetshid wasps in Cretaceous amber from Myanmar (Insecta:
Hymenoptera). Ann. Paléontol. 99, 67e77.
Perrichot, V., Beaucounru, J.C., Velten, J., 2012. First extinct genus of flea (Siphonap-
tera: Pulicidae) in Miocene amber from the Dominican Republic. Zootaxa 3438,
54e61.
Perrichot, V., Antoine, P.O., Salas Gismondi, R., Flynn, J.J., Engel, M.S., 2014. The genus
Macroteleia Westwood in middle Miocene amber from Peru (Hymenoptera, Platygastri-
dae s.l., Scelioninae). ZooKeys 426, 119e127.
Peters, R.S., Meusemann, K., Petersen, M., Mayer, C., Wilbrandt, J., Ziesmann, T.,
Donath, A., Kjer, K.M., Asp€ ock, U., Asp€ ock, H., Aberer, A., Stamatakis, A.,
Friedrich, F., H€ unefeld, F., Niehuis, O., Beutel, R.G., Misof, B., 2014. The evolu-
tionary history of holometabolous insects inferred from transcriptome-based phylogeny
and comprehensive morphological data. BMC Evol. Biol. 14, 52.
Petrulevicius, J.F., Nel, A., de Francesci, D., Goillot, C., Antoine, P.O., Salsa-Gismondi, R.,
Flynn, J.J., 2011. First fossil blood sucking Psychodidae in South America: a sycoracine
moth fly (Insecta: Diptera) in the middle miocene Amazonian amber. Insect Syst. Evol.
42, 87e96.
Pohl, H., Beutel, R.G., 2013. The Strepsiptera-Odyssey: the history of the systematic place-
ment of an enigmatic parasitic insect order. Entomologia 4, 17e26.
Pohl, H., Kinzelbach, R.K., 1995. Neufunde von F€acherfl€ uglern aus dem Baltischen und
Dominikanischen Bernstein (Strepsiptera: Bohartillidae & Myrmecolacidae). Mitt.
Geol. Pal€aont. Inst. Univ. Hambg. 78, 197e209.
Pohl, H., Kinzelbach, R.K., 2001. First record of a female stylopid (Strepsiptera: Myrmeco-
lacidae) parasite of a prionomyrmecine ant (Hymenoptera: Formicidae) in Baltic amber.
Insect Syst. Evol. 32, 143e146.
Pohl, H., Beutel, R.G., Kinzelbach, R.K., 2005. Protoxenidae fam. nov. (Insecta, Strepsip-
tera) from Baltic amber: a ‘missing link’ in strepsipteran phylogeny. Zool. Scr. 34, 57e69.
Pohl, H., Wipfler, B., Grimaldi, D.A., Beckmann, F., Beutel, R.G., 2010. Reconstructing
the anatomy of the 42-million years old fossil Mengea tertiaria (Insecta, Strepsiptera).
Naturwissenschaften 97, 855e859.
Pohl, H., Niehuis, O., Gloyna, K., Misof, B., Beutel, R.G., 2012. A new species of Menge-
nilla (Insecta, Strepsiptera) from Tunisia. Zookeys 198, 79e101.
Poinar Jr., G.O., 1992. Life in Amber. Stanford University Press, Stanford.
196 Christina Nagler and Joachim T. Haug

Poinar Jr., G.O., 1999. Paleochordodes protus n.g., n.sp. (Nematomorpha: Chordodidae),
parasites of a fossil cockroach, with a critical examination of other fossil hairworms
and helminthes of extant cockroaches (Insects: Blattaria). Invertebr. Biol. 118, 109e115.
Poinar Jr., G.O., 2002. First fossil record of nematode parasitism of ants: a 40-million-year
tale. Parasitology 125, 1e3.
Poinar Jr., G.O., 2004a. Palaeomyia burmitis (Diptera: Phlebotomidae), a new genus and
species of Cretaceous sand flies with evidence of blood-sucking habits. Proc. Entomol.
Soc. Wash. 106, 598e605.
Poinar Jr., G.O., 2004b. Fossil evidence of spider egg parasitism by ichneumonid wasps. Beitr.
Araneol. 3, 1874e1877.
Poinar Jr., G.O., 2005a. Culex malariager, n. sp. (Diptera: Culicidae) from Dominican amber:
the first fossil mosquito vector of Plasmodium. Proc. Entomol. Soc. Wash. 107, 548e553.
Poinar Jr., G.O., 2005b. Plasmodium dominicana n. sp. (Plasmodiidae: Haemospororida) from
Tertiary Dominican amber. Syst. Parasitol. 61, 47e52.
Poinar Jr., G.O., 2005c. Triatoma dominicana sp. n. (Hemiptera: Reduviidae: Triatominae),
and Trypanosoma antiquus sp. n. (Stercoraria: Trypanosomatidae), the first fossil
evidence of a triatomine-trypanosomatid vector association. Vector Borne Zoonotic
Dis. 5, 72e81.
Poinar Jr., G.O., 2006. Feroseta priscus (Neuroptera: Mantispidae), a new genus and species of
mantidflies in Dominican amber. Proc. Entomol. Soc. Wash. 108, 411e417.
Poinar Jr., G.O., 2007. Early Cretaceous trypanosomatids associated with fossil sand fly larvae
in Burmese amber. Mem. Inst. Oswaldo Cruz 102, 635e637.
Poinar Jr., G.O., 2008a. Leptoconops nosopheris sp. n. (Diptera: Ceratopogonidae) and Paleotry-
panosoma burmanicus gen. n., sp. n. (Kinetoplastida: Trypanosomatidae), a biting midge -
trypanosome vector association from the Early Cretaceous. Mem. Inst. Oswaldo Cruz
103, 468e471.
Poinar Jr., G.O., 2008b. Lutzomyia adiketis sp. n. (Diptera: Phlebotomidae), a vector of Pale-
oleishmania neotropicum sp. n. (Kinetoplastida: Trypanosomatidae) in Dominican amber.
Parasites & Vectors 1, 22.
Poinar Jr., G.O., 2011a. The Evolutionary History of Nematodes. Brill, Leiden.
Poinar Jr., G.O., 2011b. Vetufebrus ovatus n. gen., n. sp. (Haemospororida: Plasmodiidae)
vectored by a streblid bat fly (Diptera: Streblidae) in Dominican amber. Parasites &
Vectors 4, 229.
Poinar Jr., G.O., 2012a. Nematode parasites and associates of ants: past and present. Psyche
192017, 1e13.
Poinar Jr., G.O., 2012b. The 165-million-years itch. Curr. Biol. 22, 278e280.
Poinar Jr., G.O., 2013a. Palaeontology of nematodes. In: Schmidt-Rhaesa, A. (Ed.), Hand-
book of Zoology e a Natural History of the Phyla of the Animal Kingdom-Nematoda.
Walter de Gruyter, Berlin, pp. 173e179.
Poinar Jr., G.O., 2013b. Palaeoecological perspectives in Dominican amber. Int. J. Entomol.
46, 23e52.
Poinar Jr., G.O., 2014a. Spirochete-like cells in a Dominican amber Ambylomma tick
(Arachnida: Ixodidae). Hist. Biol. 27, 565e570. 1e6.
Poinar Jr., G.O., 2014b. Evolutionary history of terrestrial pathogens and endoparasites as
revealed in fossils and subfossils. Adv. Biol. 2014, 1e29.
Poinar Jr., G.O., 2015. The geological record of parasitic nematode evolution. Adv. Parasitol.
90, 53e92. http://dx.doi.org/10.1016/bs.apar.2015.03.002.
Poinar Jr., G.O., Buckley, R., 2006. Nematode (Nematoda: Mermithidae) and hairworm
(Nematomorpha: Chordodidae) parasites in early Cretaceous amber. J. Invertebr. Pathol.
93, 36e41.
Poinar Jr., G.O., Buckley, R., 2011. Doratomantispa burmanica n. gen., n. sp. (Neuroptera:
Mantispidae), a new genus of mantidflies in Burmese amber. Hist. Biol. 23, 169e176.
Insects as Parasites and Hosts 197

Poinar Jr., G.O., Huber, J.T., 2011. A new genus of fossil Mymaridae (Hymenoptera)
from Cretaceous amber and key to Cretaceous mymarid genera. ZooKeys 130,
461e472.
Poinar Jr., G.O., Krantz, W., 1997. A unique Mesozoic parasitic association. Naturwissen-
schaften 84, 321e322.
Poinar Jr., G.O., Miller, J.C., 2002. First fossil record of endoparasitism of adult ants (For-
micidae: Hymenoptera) by Braconidae (Hymenoptera). Annu. Entomol. Soc. Am. 95,
41e43.
Poinar Jr., G.O., Poinar, R., 2004a. Evidence of vector-borne disease of early Cretaceous
reptiles. Vector Borne Zoonotic Dis. 4, 281e284.
Poinar Jr., G.O., Poinar, R., 2004b. Paleoleishmania proterus n. gen., n. sp., (Trypanosoma-
tidae: Kinetoplastida) from Cretaceous Burmese amber. Protist 155, 305e310.
Poinar Jr., G.O., Poinar, R., 2005. Fossil evidence of insect pathogens. J. Invertebr. Pathol.
89, 243e250.
Poinar Jr., G.O., Telford, S.R., 2005. Paleohaemoproteus burmacis gen. n., sp. n. (Haemospor-
orida: Plasmodiidae) from an early Cretaceous biting midge (Diptera: Ceratopogonidae).
Parasitology 131, 79e84.
Poinar Jr., G.O., Thomas, D.B., 2014. Tripius gyraloura n. sp. (Aphelenchoidea: Sphaerular-
iidae) parasitic in the gall midge Lasioptera donacis Coutin (Diptera: Cecidomyiidae). Syst.
Parasitol. 89, 247e252.
Poinar Jr., G.O., Krantz, G.W., Boucot, A.J., Pike, T.M., 1997. A unique mesozoic parasitic
association. Naturwissenschaften 84, 321e322.
Poinar Jr., G.O., Curcic, B.P.M., Cokendolpher, J.C., 1998. Arthropod phoresy involving
pseudoscorpions in past and present. Acta Arachnol. 47, 79e96.
Poinar Jr., G.O., Zavortink, T.J., Pike, T., Johnston, P.A., 2000. Paleoculicis minutus (Diptera:
Culcidae) n. gen., n. sp., from Cretaceous Canadian amber with a summary of described
fossil mosquitoes. Acta Geol. Hisp. 35, 119e128.
Poinar Jr., G.O., Kritsky, G., Brown, A., 2012. Minyscapheus dominicanus n. gen., n. sp. (Hem-
iptera: Cicadidae), a fossil cicada in Dominican amber. Hist. Biol. 24, 329e333.
Ponomarenko, A.G., 1976. The new insect from the Cretaceous of Transbaikalia a probable
parasite of pterosaurs. Paleontol. Zhurnal 3, 102e106.
Pott, C., Labandeira, C.C., Krings, M., Kerp, H., 2008. Fossil insect eggs and ovipositional
damage on bennettitalean leaf cuticles from the Carnian (Upper Triassic) of Austria.
J. Paleontol. 82, 778e789.
Poulin, R., 2006. Evolutionary Ecology of Parasites, second ed. Princeton University Press,
Princeton.
Poulin, R., Morand, S., 2000. The diversity of parasites. Q. Rev. Biol. 75, 277e293.
Price, P.W., 1980. Evolutionary biology of parasites. vol. 15. Princeton University Press,
NY, 237 pp.
Rasnitsyn, A.P., 1969. Origin and evolution of lower Hymenoptera. Trans. Paleontol. Inst.
123, 1e196.
Rasnitsyn, A.P., 1975. Hymenoptera-apocrita of the mesozoic. Trudy Paleontol. Instituta
147, 1e132.
Rasnitsyn, A.P., 1992. Strashila incredibilis, a new enigmatic mecopteroid insect with possible
siphonapteran affinities from the upper Jurassic of Siberia. Psyche 99, 323e333.
Rasnitsyn, A.P., 1998. On the taxonomic position of the insect order Zorotypida Zoraptera.
Zool. Anz. 237, 185e194.
Rasnitsyn, A.P., Zherikhin, V.V., 1999. First fossil chewing louse from the lower Cretaceous
of Baissa, Transbaikalia (Insecta, Pediculida ¼ Phthiraptera, Saurodectidae fam. n.). Russ.
Entomol. J. 8, 253e255.
Rasnitsyn, A.P., Quicke, D.L., 2002. History of Insects. Kluwer Academic Publishers,
Dordrecht.
198 Christina Nagler and Joachim T. Haug

Reidenbach, K.R., Cook, S., Bertone, M.A., Harbach, R.E., Wiegmann, B.M.,
Besansky, N.J., 2009. Phylogenetic analysis and temporal diversification of mosquitoes
(Diptera: Culicidae) based on nuclear genes and morphology. BMC Evol. Biol. 9, 298.
Ribeiro, J.M.C., Assumpcao, T.C., Francischetti, I.M.B., 2012. An insight into the sialomes
of blood-sucking Heteroptera. Psyche 2012, e470436.
Riquelme, F., Pieder-Jimenez, D.F., Cordova-Tabares, V., Luna-Castro, B., 2014. A new
chernetid pseudoscorpion from the miocene chiapas- Amber Lagerst€atte, Mexico.
Can. J. Earth Sci. 51, 902e908.
Ross, A., 1998. Amber. Harvard University Press, Harvard.
Rota-Stabelli, O., Daley, A.C., Pisani, D., 2013. Molecular timetrees reveal a Cambrian colo-
nization of land and a new scenario for ecdysozoan evolution. Curr. Biol. 23, 392e398.
Rozen Jr., J.G., 2003. Eggs, ovariole numbers, and modes of parasitism of cleptoparasitic bees,
with emphasis on Neotropical species (Hymenoptera: Apoidea). Am. Mus. Novitates
3413, 1e36.
Schawaller, W., 1991. The first Mesozoic pseudoscorpion, from Cretaceous Canadian amber.
Palaeontology 34, 971e976.
Schawaller, W., Shear, W.A., Bonamo, P.M., 1991. The first Paleozoic pseudoscorpions
(Arachnida, Pseudoscorpionida). Am. Mus. Novitates 3009, 1e17.
Schierenberg, E., Sommer, R.J., 2011. Reproduction and development in Nematodes. In:
Schmidt-Rhaesa, A. (Ed.), Handbook of Zoology e a Natural History of the Phyla of
the Animal Kingdom e Gastrotricha, Cycloneuralia and Gnathifera. Walter de Gruyter,
Berlin, pp. 61e103.
Schmidt, A.R., Jancke, S., Lindquist, E.E., Ragazzi, E., Roghi, G., Nascimbene, P.C.,
Schmidt, K., Wappler, T., Grimaldi, D.A., 2012. Arthropods in amber from the Triassic
period. Proc. Natl. Acad. Sci. U.S.A. 109, 14796e14801.
Schmidt-Rhaesa, A., 2013. Nematomorpha. In: Schmidt-Rhaesa, A. (Ed.), Handbook of
Zoology e a Natural History of the Phyla of the Animal Kingdom-Nematomorpha,
Priapulida, Kinorhyncha, Loricifera. Walter de Gruyter, Berlin, pp. 29e123.
Schmidt-Rhaesa, A., Ehrmann, R., 2001. Horsehair worms (Nematomorpha) as parasites of
praying mantids with a discussion of their life cycle. Zool. Anz. 240, 167e179.
Sharkey, M.J., Carpenter, J.M., Vilhelmsen, L., Heraty, J.M., Liljeblad, J., Dowling, A.,
Schulmeister, S., Murray, D., Deans, A.R., Ronquist, F., Krogmann, L., Wheeler, W.C.,
2012. Phylogenetic relationships among superfamilies of Hymenoptera. Cladistics 28,
80e112.
Shear, W.A., Kukalova-Peck, J.K., 1990. The ecology of Paleozoic terrestrial arthropods: the
fossil evidence. Can. J. Zool. 68, 1807e1834.
Shih, X., Zhao, Y., Shih, C., Ren, D., 2013. New fossil helorid wasps (Insecta, Hymenop-
tera, Proctotrupoidea) from the Jehol Biota, China. Cretac. Res. 41, 136e142.
Shih, C., Ohl, M., Wunderlich, J., Ren, D., 2015. A remarkable new genus of Mantispidae
(Insecta, Neuroptera) from Cretaceous amber of Myanmar and its implications on rapto-
rial foreleg evolution in Mantispidae. Cretac. Res. 52, 416e422.
Sivinski, J., Marshall, S., Petersson, E., 1999. Kleptoparasitism and phoresy in the Diptera.
Fla. Entomol. 82, 179e197.
Smith, V.S., Ford, T., Johnson, K.P., Johnson, P.C.D., Yoshizawa, K., Light, J.E., 2011.
Multiple lineages of lice pass through the KePg boundary. Biol. Lett. 7, 782e785.
Solorzano-Kramer, M.M., Nel, A., 2009. First evidence in the fossil record of snipe flies
(Diptera: Rhagionidae) in Cretaceous amber, France. Cretac. Res. 30, 1367e1375.
Sontag, E., Szadziewski, R., 2011. Biting midges (Diptera: Ceratopogonidae) in Eocene
Baltic amber from the Rovno region (Ukraine). Pol. J. Entomol. 80, 779e800.
Strelow, J., Kraemer, M.M.S., Ibanez-Bernal, S., Rust, J., 2013. First fossil horsefly (Diptera:
Tabanidae) in Miocene mexican amber. Pal€aontologische Ztg. 87, 437e444.
Insects as Parasites and Hosts 199

Szadziewski, R., 1995. The oldest fossil corethrellidae (Diptera) from lower Cretaceous Leb-
anese amber. Acta Zool. Cracov. 38, 177e181.
Szadziewski, R., Ross, A., Gi, W., 2015. Further records of biting midges (Diptera: Ceratopo-
gonidae) from upper Cretaceous Burmese amber (Myanmar). Cretac. Res. 52, 556e561.
Szymkowiak, P., Gorski, G., Bajerlein, D., 2007. Passive dispersal in arachnids. Biol. Lett. 44,
75e101.
Trojan, P., 2002. First discovery of Bouvieromyiini (Diptera: Tabanidae: Chrysopsinae) in
Baltic amber. Ann. Zool. 52, 257e270.

Ulrich, W., 1927. Uber das bisher einzige Strepsipteron aus dem baltischen Bernstein und
€ber eine Theorie der Mengeidenbiologie. Z. f€
u €
ur Morphol. Okologie Tiere 8, 45e62.
Veltz, I., Azar, D., Nel, A., 2007. New chironomid flies in early Cretaceous Lebanese amber
(Diptera: Chironomidae). Afr. Invertebr. 48, 169e191.
Vilhelmsen, L., 2003. The old wasp and the tree: fossils, phylogeny and biogeography in the
Orussidae (Insecta, Hymenoptera). Biol. J. Linn. Soc. 82, 139e160.
Vilhelmsen, L., Zimmermann, D., 2014. Baltorussus total makeover: rejuvenation and sex
change in an ancient parasitoid wasp lineage. PLoS One 96, e98412.
Vilhelmsen, L., Stephan, A.F., Blank, M., Costac, V.A., Alvarengad, T.M., Smith, D.R.,
2013. Phylogeny of the ophrynopine clade revisited: review of the parasitoid sawfly
genera Ophrella Middlekauff, Ophrynopus Konow and Stirocorsia Konow (Hymenoptera:
Orussidae). Invertebr. Syst. 27, 450e483.
Vransky, P., Ren, D., Shih, C., 2010. Nakridletia ord.n. - enigmatic insect parasites support
sociality and endothermy of pterosaurs. Amba Proj. 8, 1e16.
Voigt, E., 1952. Ein Haareinschluss mit Phthirapteren Eiern in Bernstein. Mitt. Geol. Staat-
sinst. Hambg. 21, 59e74.
Vullo, R., Vincent, G., Azar, D., Neraudeau, D., 2010. Mammalian hairs in early Cretaceous
amber. Naturwissenschaften 97, 683e687.
Wang, M., Rasnitsyn, A.P., Shih, C., Ren, D., 2014a. A new Cretaceous genus of xyelydid
sawfly illuminating nygmata evolution in Hymenoptera. BMC Evol. Biol. 14, 131.
Wang, M., Rasnitsyn, A.P., Ren, D., 2014b. Two new fossil sawflies (Hymenoptera, Xye-
lidae, Xyelinae) from the middle Jurassic of China. Acta Geol. Sin. 88, 1027e1033.
Wappler, T., Smith, V.S., Dalgeish, R.C., 2004. Scratching an ancient itch: an Eocene bird
louse fossil. Proc. R. Soc. Lond. B 271, 255e258.
Wappler, T., Currano, E.D., Wilf, P., Rust, J., Labandeira, C.C., 2009. No post-Cretaceous
ecosystem depression in European forests? Rich insect-feeding damage on diverse middle
Palaeocene plants, Menat, France. Proc. R. Soc. B 276, 4271e4277.
Wappler, T., Tokuda, M., Yukawa, J., Wilde, V., 2010. Insect herbivores on Laurophyllum
lanigeroides (Engelhardt 1992) Wilde: a role of a distinct planteinsect associational suite
in host taxonomic assignment. Palaeontogr. B 283, 137e155.
Ward, P., Brady, S., Fisher, B., Schultz, T., 2014. The evolution of myrmicine ants: phylog-
eny and biogeography of a hyperdiverse ant clade (Hymenoptera: Formicidae). Syst.
Entomol. 40, 61e86.
Wedmann, S., Makarkin, V.N., 2007. A new genus of Mantispidae (Insecta: Neuroptera)
from the Eocene of Germany, with a review of the fossil record and palaeobiogeography
of the family. Zool. J. Linn. Soc. 149, 701e716.
Weitschart, W., Wichard, W., 1998. Atlas der Pflanzen und Tiere im Baltischen Bernstein.
Pfeil Verlag, M€ unchen. 256 pp.
Whiting, M.F., 2002. Mecoptera is paraphyletic: multiple genes and phylogeny of Mecoptera
and Siphonaptera. Zool. Scr. 31, 93e104.
Whiting, M.F., Carpenter, J.M., Wheeler, Q.D., Wheeler, W.C., 1997. The Strepsiptera
problem: phylogeny of the holometabolous insects orders inferred from 18S and 28S
ribosomal sequences and morphology. Syst. Biol. 46, 1e68.
200 Christina Nagler and Joachim T. Haug

Whiting, M.F., Whiting, A.S., Hastriter, M.W., Dittmar, K., 2008. A molecular phylogeny
of fleas (Insecta: Siphonaptera): origins and host association. Cladistics 24, 1e31.
Wiegmann, B.M., Trautwein, M.D., Kim, J.-W., Cassel, B.K., Bertone, M.A.,
Winterton, S.L., Yeates, D.K., 2009. Single-copy nuclear genes resolve the phylogeny
of the holometabolous insects. BMC Evol. Biol. 7, 34e50.
Wiegmann, B.M., Trautwein, M.D., Winkler, I.S., Barr, N.B., Kim, J.-K., Lambkin, C.,
Bertone, M.A., Cassel, B.K., Bayless, K.M., Heimberg, A.M., Wheeler, B.M.,
Peterson, K.J., Pape, T., Sinclair, B.J., Skevington, J.H., Blagoderov, V., Caravas, J.,
Kutty, S.N., Schmidt-Ott, U., Kampmeier, G.E., Thompson, F.C., Grimaldi, D.A.,
Beckenbach, A.T., Courtney, G.W., Friedrich, M., Meier, R., Yeates, D.K., 2011.
Episodic radiations in the fly tree of life. Proc. Natl. Acad. Sci. U.S.A. 108, 5690e5695.
Xianguang, H., Bergstr€ om, J., 1994. Palaeoscolecid worms may be nematomorphs rather
than annelids. Lethaia 27, 11e17.
Yao, Y., Cai, W., Xu, J.X., Shih, C., Engel, M.S., Zheng, X., Zhao, Y., Ren, D., 2014.
Blood-feeding true bugs in the early Cretaceous. Curr. Biol. 24, 1786e1792.
Yoshizawa, K., Johnson, K.P., 2003. Phylogenetic position of Phthiraptera (Insecta:
Paraneoptera) and elevated rate of evolution of mitochondrial 12S and 16S rDNA.
Mol. Phylogenet. Evol. 29, 102e114.
Yoshizawa, K., Johnson, K.P., 2010. How stable is the “Polyphyly of Lice” hypothesis
(Insecta: Psocodea)?: A comparison of phylogenetic signal in multiple genes. Molecular
Phylogenetics and Evolution 55, 939e951.
Yoshizawa, K., Lienhard, C., 2010. In search of the sister group of true lice: a systematic
review of booklice and their relatives, with an updated checklist of Liposcelididae
(Insecta: Psocodea). Athropod Syst. Phylog. 68, 181e195.
Zhang, J., 2012. New horseflies and water snipe-flies (Diptera: Tabanidae and Athericidae)
from the lower Cretaceous in China. Cretac. Res. 36, 1e5.
CHAPTER FIVE

Trace Fossil Evidence of


TrematodeeBivalve
ParasiteeHost Interactions
in Deep Time
John Warren Huntley*, 1, Kenneth De Baetsx
*Department of Geological Sciences, University of Missouri, Columbia, MO, USA
x
Fachgruppe Pal€aoUmwelt, GeoZentrum Nordbayern, Friedrich-Alexander-Universit€at
Erlangen-N€urnberg, Erlangen, Germany
1
Corresponding author: E-mail: huntleyj@missouri.edu

Contents
1. Introduction 202
2. Trematode-Induced Shell Malformations in Living Bivalve Molluscs 206
3. Occurrences of Trematode-Induced Pits in Fossil and Subfossil Bivalves 211
3.1 Taxonomic, temporal and ecological occurrences 211
3.2 Taphonomy and the origin of the trematodeebivalve parasiteehost 217
interaction
3.3 Trematode-induced malformations as palaeoenvironmental indicators 220
4. Detrimental Effects of Trematodes on Living Bivalves and Their Potential 222
Evolutionary Implications
5. Concluding Remarks 225
Acknowledgements 226
References 226

Abstract
Parasitism is one of the most pervasive phenomena amongst modern eukaryotic life
and yet, relative to other biotic interactions, almost nothing is known about its history
in deep time. Digenean trematodes (Platyhelminthes) are complex life cycle parasites,
which have practically no body fossil record, but induce the growth of characteristic
malformations in the shells of their bivalve hosts. These malformations are readily pre-
served in the fossil record, but, until recently, have largely been overlooked by students
of the fossil record. In this review, we present the various malformations induced by
trematodes in bivalves, evaluate their distribution through deep time in the phyloge-
netic and ecological contexts of their bivalve hosts and explore how various tapho-
nomic processes have likely biased our understanding of trematodes in deep time.
Trematodes are known to negatively affect their bivalve hosts in a number of ways
including castration, modifying growth rates, causing immobilization and, in some
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.05.004 All rights reserved. 201
202 John Warren Huntley and Kenneth De Baets

cases, altering host behaviour making the host more susceptible to their own
predators. Digeneans are expected to be significant agents of natural selection. To
that end, we discuss how bivalves may have adapted to their parasites via heterochrony
and suggest a practical methodology for testing such hypotheses in deep time.

1. INTRODUCTION
Interpreting the history and evolutionary implications of antagonistic bi-
otic interactions in the marine invertebrate fossil record has been a primary
theme in palaeobiology during the last few decades. The primary focus of
this pursuit has been the examination of predatoreprey interactions, which
has highlighted the likely role of antagonistic interactions in shaping macro-
evolutionary trends (Vermeij, 1977; Signor and Brett, 1984; Kowalewski
et al., 1998; Kelley and Hansen, 2003; Madin et al., 2006; Huntley and
Kowalewski, 2007). Competitive interactions have received some attention
in the literature (Hermoyian et al., 2002; Huntley et al., 2008; Tyler and
Leighton, 2011), but may not be an important agent of natural selection in
benthic marine environments (Stanley, 2008). Parasitism, however, is one
of the most pervasive phenomena amongst modern eukaryotic life (Poulin
and Morand, 2000) and, relative to predation, almost nothing is known about
it in deep time (Littlewood and Donovan, 2003). Parasites and the traces they
leave behind have been preserved as fossils (Moodie, 1923; Cameron, 1967;
Fry and Moore, 1969; Conway Morris, 1981; Ruiz and Lindberg, 1989;
Boucot, 1990; Savazzi, 1995; Feldmann, 1998; Feldman and Brett, 1998;
Littlewood and Donovan, 2003; Boucot and Poinar, 2010), but the docu-
mentation of the intensity of specific parasiteehost interactions through
deep time is rare indeed (for some notable exceptions see Brett, 1978;
Baumiller and Gahn, 2002; Gahn and Baumiller, 2003; Klompmaker et al.,
2014; Wilson et al., 2014; De Baets et al., 2011, 2015a). Based upon their
relative positions within trophic webs, parasites are likely to be more abun-
dant and diverse than predators in modern ecosystems. There is an increasing
appreciation for the roles of parasites in shaping communities (Hudson et al.,
2006) and promoting evolvability (Zaman et al., 2014), but what role have
they played in shaping ecological and macroevolutionary trends of their hosts
in deep time? In order to discern the importance of parasitism in deep time,
one must first explore its origin in a phylogenetically informed context, sys-
tematically analyze the occurrence and intensity of parasiteehost interactions
through geologic time, consider the potential biases introduced by the process
of fossilization and try to disentangle these biases from the biotic record so that
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 203

we can understand how the biases might distort our view of the history of
parasiteehost interactions.
Bivalve molluscs serve as hosts to a diverse group of parasites and disease-
causing agents including viruses, prokaryotes (e.g. Chlamydia and Rickettsia),
fungi, protistans, parazoans (e.g. boring sponges like Cliona) and metazoans
(e.g. platyhelminths, annelids, molluscs, bryozoans and arthropods); see
Lauckner (1983) for an excellent account of the myriad parasites and diseases
of bivalves. Given that many parasites are typically small-bodied and often
lack biomineralized skeletons, we should not expect an extensive record
of body fossils (although, of course there are exceptions: Fry and Moore,
1969; Baumiller and Gahn, 2002; Gahn and Baumiller, 2003; Poinar,
2003; amongst others); therefore, we must seek evidence for parasitism in
the form of traces and malformations on the mineralized hosts. Unfortu-
nately, not all interactions result in the production of a trace or shell
malformation.
A variety of taxa can produce or induce the formation of a trace or shell
malformation when interacting with the bivalve. Indeed, one of the primary
challenges when interpreting these traces is determining if the interaction
was ante-mortem or postmortem, relative to the bivalve. From this perspec-
tive, we can categorize such interactions and their resulting traces into two
groups: (1) interactions that are the result of active boring or shell destruction
by the putative parasite, and, thereby, we cannot distinguish between a
liveedead and a liveelive interaction, and (2) interactions that induce a
growth reaction by the bivalve, and, thereby, we know that the trace or mal-
formation was the result of a liveelive interaction.
Many taxa produce this first type of trace in bivalves and they may not
be, in some cases, strictly parasitic in origin. For instance, spionid polychaete
worms will frequently bore into calcium carbonate substrates (including
bivalves) and produce characteristic U-shaped (and sometimes more com-
plex shaped) borings with paired openings adjacent to one another on the
valve surface (Blake and Evans, 1973; Thayer, 1974; Zottoli and Carriker,
1974; Huntley, 2007; Rodrigues, 2007; Huntley and Scarponi, 2015). Spio-
nids are generally capable of suspension- and deposit-feeding (Dix et al.,
2005). Their boring activity likely weakens their host’s shell and valve and
site preference of spionids suggests they kleptoparasitize the feeding currents
of their hosts (Huntley, 2007; Rodrigues, 2007). The work of Rodrigues
et al. (2008) revealed that spionid borings in live brachiopods of coastal Brazil
were nearly always occupied by live spionids, whereas live spionids were not
identified in borings found in dead brachiopods from the same location.
204 John Warren Huntley and Kenneth De Baets

Despite evidence for liveelive interactions in modern settings, we cannot


always say with certainty that such traces were produced on live bivalves
in the past. There are many other taxa that bore into bivalves and other
carbonate substrates, which may or may not be parasitic, including sponges
(e.g. Cliona), foraminifera, bryozoans, barnacles, bivalves and predatory gas-
tropods (Boucot, 1990; Boucot and Poinar, 2010). Given that these traces
can be made following the death of the putative host they will not be
considered further here.
In order to maximize the likelihood of studying liveelive interactions in
the fossil record, it is imperative to investigate interactions where the actions
of the parasite induced a growth reaction by the host. Such traces are also
made by the aforementioned, potentially kleptoparasitic and shell weak-
ening, spionid polychaetes in the form of mudblisters. Mudblisters form
when the spionid begins boring at the growing margin of the bivalve, result-
ing in a raised blister within the valve, which is filled with mud and worm
faeces (Blake and Evans, 1973; Huntley, 2007; Huntley and Scarponi, 2015).
Similar blister and tube-like traces of unknown origin have been docu-
mented in the fossil record (Boucot, 1990; Boucot and Poinar, 2010).
Ozanne and Harries (2002) document the occurrence of ‘bubbly nacre’
and ‘Hohlkehle’ (a rib-like malformation on the valve interior) in Cretaceous
inoceramids of the Western Interior Seaway. Though the taxonomic iden-
tities of the culprits are uncertain, Ozanne and Harries (2002) suggested that
the traces were the result of pathogens or parasites between the mantle and
shell wall and parasitic polychaetes, respectively. Savazzi (1995) reported
shell malformation of diverse morphologies, including cavities, pits, ridges
and crests, near the ligament in Pliocene Isognomon (Hippochaeta) maxillatus
from Northwestern Italy. Savazzi (1995) hypothesized that these traces
were the result of bivalve growth (rather than bioerosion) responding to
what was likely a polychaete endoparasite. In both examples, we are left
hypothesizing the taxonomic nature of the putative parasite.
There is a parasiteehost interaction that has been comparatively well-
studied in modern ecosystems, which results in the induction of characteristic
host shell malformations, and is, therefore, amenable to study in deep time via
the fossil record of bivalve hosts. Digenean trematodes (Platyhelminthes) are
endoparasitic flatworms with complex life cycles that can include free-
swimming stages, one or more intermediate stages within an invertebrate
host (bivalves, gastropods, arthropods, annelids), occasionally and subsequently
a vertebrate intermediate and/or a final stage within a vertebrate definitive
host (Cribb et al., 2001, 2003; Littlewood, 2006). Trematodes can cause all
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 205

manner of trouble in their bivalve hosts including castration, stunted growth,


gigantism, immobilization and induced risky behaviour (Swennen, 1969; Lim
and Green, 1991; Ballabeni, 1995; Taskinen, 1998; Hechinger et al., 2009).
The soft-bodied trematodes themselves have very low body fossilization
potential (Poinar, 2003; Littlewood and Donovan, 2003; De Baets et al.,
2015b); however, gymnophallid trematodes are known to induce the growth
of an array of shell malformations on the interior of the bivalve, when it is
serving as a second intermediate host, within the pallial line (the best known
being oval pits with raised rims), which have a greater, more continuous
fossilization potential.
Ruiz and colleagues (Ruiz and Lindberg, 1989; Ruiz, 1991) docu-
mented a 5-million-year history of parasitism amongst bivalves of North
America and identified the oldest known occurrences of gymnophallid
trematode-induced pits in three genera from the Eocene (Lutetian) strata
of the Paris Basin (a record that was recently extended to an even earlier
Age of the Eocene (Ypresian) by Todd and Harper (2011) from the London
Clay of southern England). For nearly two decades this initial study stood
alone in its attempt to document trematodeebivalve interactions in deep
time. Recently, the fossil record of trematode-induced malformations has
garnered more attention due to the increasing recognition of its potential
for revealing the ecological and evolutionary implications of parasiteehost
interactions in deep time and amongst modern death assemblages (Ituarte
et al., 2001, 2005; Huntley, 2007; Todd and Harper, 2011; Huntley and
Scarponi, 2012, 2015; De Baets et al., 2015b). Indeed, high temporal reso-
lution records of Holocene trematode malformations provide unique in-
sights into the influence of climate change on parasitic interactions
through geologic time and enable us to make predictions for how such in-
teractions will potentially respond to anthropogenic climate change in the
future (Huntley et al., 2014). Ruiz and Lindberg (1989) hypothesized that
the gymnophallid-bivalve parasiteehost interaction originated in Europe
during the Eocene, expanded to the North American Atlantic coast by
the Miocene, and again to the North American Pacific coast by the Plio-
cene. These results are, of course, highly preliminary and warrant further
study to determine the timing of the origin and dispersal of this interaction.
The purpose of this paper is to review our current state of knowledge of
the fossil record of trematodeebivalve interactions and to outline prospects
for future inquiry. This includes discussions of the morphologies of gymno-
phallid trematode-induced malformations in bivalves, examples of living par-
asites in close proximity to malformations in the shells of their hosts, the
206 John Warren Huntley and Kenneth De Baets

temporal distribution of reliable trematode traces within the phylogenetic


context of their bivalve intermediate hosts, the life and feeding modes of
infested bivalve taxa through geologic time, problems of taphonomy and
how they influence our understanding of the origin and history of trema-
todeebivalve interactions, the utility of trematode traces as palaeoenviron-
mental indicators and the negative effects of trematodes on bivalves and
their likely evolutionary implications. Despite all that we have learnt about
this interaction in deep time, many more questions about the origin and
history of this interaction present themselves. Indeed, there is room for addi-
tional palaeontologists and neontologists to address these important (palaeo-)
ecological questions.

2. TREMATODE-INDUCED SHELL MALFORMATIONS IN


LIVING BIVALVE MOLLUSCS
Live digenean (mostly gymnophallid) trematodes have been found
living in close spatial proximity to shell malformations on the interior of
their bivalve hosts. These malformations have taken a variety of forms,
and these forms span a range of utility for being characteristic traces of trem-
atodes (Figure 1). Ranging from the most to the least reliable, indicators of
trematode parasitism are pits, igloos, blister pearls, free pearls, irregular
calcareous deposits (ICD) and valve discolouration (Table 1).
The best known characteristic shell malformations induced by trematode
parasites are oval-shaped pits with raised rims (Figure 1(a) and (b)). Ruiz and
Lindberg (1989) suggested that these pits are not features of shell erosion,
rather, the pits are the result of shell growth intended to encapsulate the
parasite. This interpretation is evidenced by growth lines wrapping around
the pit structure when viewed in thin section rather than the pit cutting
across growth lines (as would be expected with shell erosion). Most trema-
tode taxa found living in association with pits are members of the Gymno-
phallidae and were parasitizing taxa from the bivalve families Veneridae,
Semelidae, Cardiidae and Psammobiidae (Table 1; Johannessen, 1973;
Campbell, 1985; Ruiz and Lindberg, 1989; Ituarte et al., 2001, 2008).
Though gymnophallids are typically considered to be the digeneans respon-
sible for inducing pit growth in their hosts, an unknown species from the
digenean family Lepocreadiidae has been found living in association with
a pit on the interior of a cyamiid bivalve, Gaimardia trapesina from southern
Argentina (Table 1; Ituarte et al., 2001). Pits are the traces most frequently
cited as fossil indicators of trematode parasitism (Ruiz and Lindberg, 1989;
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 207

Figure 1 Images of trematode-induced shell malformations. Photomicrographs of


trematode-induced pits in (a) Holocene Potamocorbula amurensis from the Pearl River
Delta, China (Core PRD-10, sample 157) and (b) Holocene Chamalea gallina from the Po
Plain, Italy (Core 240S8, 13.10 m). (c) Flatbed scanner image of blister pearls in Eocene
(Lutetian) Sunetta semisulcata from the Paris Basin (Grignon, Faluni ere), France
(Gg2003/517, Museum National d’Histoire Naturelle). (d) Scanning electron micrograph
of igloo-shaped malformation in Recent Neolepton bennetti from the Falkland Islands
(Adapted from Ituarte et al. (2005).). (e) Photomicrograph of discolouration that may
be related to trematode infestation in Holocene Cyrenodonax formosana from the Pearl
River Delta, China (Core PRD-10, sample 120). (f) Photomicrograph of irregular calcar-
eous deposits that may be related to trematode infestation in Holocene C. formosana
from the Pearl River Delta, China (Core PRD-10, sample 124).

Ruiz, 1991; Huntley, 2007; Todd and Harper, 2011; Huntley and Scarponi,
2012, 2015; Huntley et al., 2014; De Baets et al., 2015b), and there is no
other biotic interaction (known so far) that induces the growth of an
oval-shaped pit with a raised rim.
Ituarte and colleagues (Ituarte et al., 2001, 2005; Presta et al., 2014) have
described unusual igloo-shaped morphologies of trematode-induced traces
from high-latitude shallow marine deposits in Argentina (Figure 1(d)).
Ituarte et al. (2001) interpret the igloo structures to represent a response
of the host meant to isolate the parasite. In this case, however, the parasite
is able to maintain an opening between the interior of the would-be blister
pearl (see below) and the space between the mantle and shell, thus
Table 1 Instances of living trematodes identified in contact with shell malformations in living bivalve hosts

208
Trematode parasite Bivalve host
Feeding
Family Species Trace type Family Genus species Life mode mode Reference
Gymnophallidae Gymnophallus sp. Pits, discolouration Veneridae Venerupis pullastra Infaunal Suspension Johannessen (1973)
Parvatrema rebecqui Pits Semelidae Abra tenuis Deposit Campbell (1985)
P. rebecqui Cardiidae Cerastoderma glaucum Suspension
Gymnophallus somateriae Pearls Mytilidae Mytilus sp. Epifaunal Jameson (1903)
e ‘Unknown distome’ Pearls Mytilidae Mytilus edulis Epifaunal Suspension Stunkard and Uzmann
Brachycoelium luteum ICD Donacidae Donax trunculus Infaunal (1958)
B. luteum Tellinidae Tellina fabula Deposit
Tellina tenuis
Tellina solidula
Distomum margaritarum Pearls Mytilidae M. edulis Epifaunal Suspension
D. margaritarum Mytilus galloprovincialis
‘Unknown trematode’ Margaritiferidae Margaritifera margaritifera
Proctoeces milfordensis Mytilidae M. edulis

John Warren Huntley and Kenneth De Baets


Gymnophallidae Bartolius pierrei ICD Mactridae Darina solenoides Infaunal Suspension Cremonte and Ituarte
(2003)
Unknown species Pits, pearls, discolouration Psammobiidae Tagelus plebius Ituarte et al. (2008)
ICD, discolouration Lomovasky et al. (2005)
ICD, discolouration, pearls Vazquez et al. (2006)
Bartolius sp. Igloo Neoleptonidae Neolepton cobbi e e Presta et al. (2014)
Unknown species Cyamiidae Gaimardia trapesina Epifaunal Suspension Ituarte et al. (2001)
Lepocreadiidae Unknown species Pits Cyamiidae G. trapesina Epifaunal Suspension Ituarte et al. (2001)
Gymnophallidae Parvatrema borealis Pits Veneridae Gemma gemma e e Ruiz and Lindberg
Transennella confusa Infaunal Suspension (1989)
Transennella tantilla

ICD ¼ irregular calcareous deposits.


Blank cells indicate the same content as the above cell and the dashed lines indicate that the value is unknown.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 209

producing an igloo-morphology. Bartolius sp. and two other gymnophallid


species whose genus- and species-level taxonomy were not resolved have
been identified living in igloo-shaped traces on the interior of cyamiid
and neoleptonid bivalves (Ituarte et al., 2001; Presta et al., 2014). Ituarte
et al. (2005) extended identified igloo-shaped traces into the subfossil record
in Holocene cyamiid, neoleptonid, and nuculanid Holocene bivalves from
Falkland Islands and Tierra del Fuego. Interestingly, Liljedahl (1985, 1994)
noted similar traces on the interior of praenuculid bivalves from Silurian
strata in Sweden, although these are not consistent with known extant
host-relationships between shore birds and gymnophallid trematodes
(compare De Baets et al., 2015b). Fossils give a minimum constraint to
the origin of modern shore birds or charadriiform-like forms to the Eocene
(Mayr, 2014), although various molecular clocks place the origin of Chara-
driiformes into the Cretaceous, although no unequivocal fossils have been
found from this interval (Smith, 2015).
Blister pearls are small, discrete, round to elongate, convex features on
the interior surface of bivalves which are sometimes associated with trem-
atode parasites (Figure 1(c)). When co-occurring with pits, blister pearls are
commonly of comparable size to the oval-shaped pits. The earliest known
blister pearls date from the late Silurian to early Devonian (Kríz, 1979;
Liljedahl, 1985, 1994; De Baets et al., 2011) and are often thought to repre-
sent overgrowths of dead parasites as in recent bivalves (Lauckner, 1983).
Campbell (1985) described shell malformations in the bivalve Abra tenuis
reacting against the trematode Gymnophallus rebecqui (termed ‘blisters’ by
the author) that are intermediate in morphology between oval pits and blis-
ter pearls. In no instance was the parasite completely isolated by the shell
malformation, but this morphology serves as further support for the
inducing of pearl growth by trematodes. Gymnophallid trematodes have
been found in association with pearl-bearing bivalve individuals from the
Mytilidae, Margaritiferidae and the Psammobiidae families in Europe,
North America and South America (Jameson, 1903; Stunkard and
Uzmann, 1958; Vazquez et al., 2006; Ituarte et al., 2008). Free pearls are
similar to blister pearls in that they are the result of a bivalve’s defensive
overgrowth of foreign bodies, however, they are not attached to the shell
wall. Free pearls date as far back as the Triassic (Kutassy, 1937; Conway
Morris, 1981; Littlewood and Donovan, 2003). Although various parasitic
flatworms have been associated with pearl formation (both blister and free),
it is important to note that other pathogens and inorganic particles can also
210 John Warren Huntley and Kenneth De Baets

result in pearl formation (Newton, 1908; G€ otting, 1974, 1979; Lauckner,


1983). Therefore one should use caution when attempting to positively
identify trematode parasitism via pearls.
ICD are low-relief malformations occurring in the interior of the shell.
ICDs are similar in morphology to spionid mudblisters except that there is
no void space in the valve underlying them. ICDs have been identified in
association with Brachycoelium luteum in donacid and tellinid bivalves from
France (Stunkard and Uzmann, 1958), and digenean trematodes (Bartolius
pierrei and two unidentified species) in mactrid and psammobiid bivalves
from Argentina (Cremonte and Ituarte, 2003; Lomovasky et al., 2005;
Vazquez et al., 2006; Table 1). Additionally, valve discolouration has
been identified in association with digenean trematodes in venerid and
psammobiid bivalves in Norway and South America (Johannessen, 1973;
Cremonte and Ituarte, 2003; Lomovasky et al., 2005; Vazquez et al.,
2006; Table 1). Features similar to ICDs and shell discolouration have
been identified in Cyrenodonax formosana from core-sampled Holocene de-
posits of the Pearl River Delta, China (Huntley et al., 2014). No association
with trematodes could be confirmed, of course, as these are subfossil
deposits. The probability of preserving discolouration likely decreases
dramatically with geologic age.
Blisters, ICD, and shell discolouration are the poorest diagnostic traces
of trematode parasites. Pits and igloos are highly diagnostic, in that they are
morphologically complex and no other known organisms induce the
growth of such structures today. Blister pearls and ICDs, by comparison,
are of a more simple morphology and could potentially be induced by
a diverse group of irritants (G€ otting, 1974, 1979; Lauckner, 1983).
Caceres-Martínez and Vasquez-Yeomans (1999) documented the presence
of pearls in Mytilus galloprovincialis and Mytilus californianus with copepods
and platyhelminths that also lacked trematodes (though other individuals
in the samples possessed various combinations of the three types of parasites
and pearls). Shell discolouration and ICDs can be linked to a number of
parasites including chromalvelolates (Haplosporidium nelsoni) and fungi
(Ostracoblabe implexa; Elston, 1990). One should only use blisters, ICDs
or shell discolouration as a direct evidence for trematode infestation in
the instance that the interaction is well documented amongst living repre-
sentatives or, at worst, with closely related taxa. Pits and igloos are the most
reliable proxies for gymnophallid trematode infestation in the fossil and
subfossil record.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 211

3. OCCURRENCES OF TREMATODE-INDUCED PITS IN


FOSSIL AND SUBFOSSIL BIVALVES
3.1 Taxonomic, temporal and ecological occurrences
It is instructive to explore the distribution of gymnophallid trematode
infestation across the phylogeny of their bivalve intermediate hosts in deep
time. Such a distribution can aid in the development of hypotheses regarding
the origin of this interaction and the potential nonbiological (taphonomic)
biases upon this record. In the interest of taking the most conservative
approach, only pits, igloos and igloo-like structures from fossil and subfossil
(Holocene samples from sub-surface deposits and surficial death assemblages)
records will be considered in this analysis. Data are compiled from occurrence
data from the North Sea coast of the Netherlands (Huntley, unpublished
data), museum survey of Eocene (Lutetian) bivalves in the collections of
the Museum National d’Histoire Naturelle (MNdHN; Paris; Huntley, un-
published data), and from the peer-reviewed literature. Bulk samples of death
assemblages from the swash zone were collected by the senior author at 10
locations in the Netherlands (Callantsoog, Egmond Aan Zee, Katwijk Aan
Zee, Hoek van Holland, Zandvoort, IJmuiden, Harlingen, Makkum,
Workum and Wierum) consisting of 7423 valves from 16 taxa. The senior
author conducted a cursory survey of the extensive Eocene (Lutetian) bi-
valves from the Paris Basin in the collections of the MNdHN on lots with
multiple specimens (1505 valves from 147 species occurrences) from
Grignon, Faluniere. The occurrence of trematode-induced traces was
plotted on the evolutionary time tree of the Bivalvia modified from Bieler
et al. (2014) (Figure 2, Table 2).
Trematode-induced traces have been identified in four of the six major
bivalve lineages (Bieler et al., 2014): Protobranchia, Pteriomorphia, Archihe-
terodonta and Imparidentia, including 46 genera from 19 families (Figure 2,
Table 2). This record is most complete in Eocene and younger sediments and
is characterized by significant temporal gaps. The two lineages in which trem-
atode traces have not yet been identified are Palaeoheterodonta and Anom-
alodesmata, but this might represent a sampling artefact given the dearth of
attention this interaction has historically received from palaeontologists.
The largest temporal gap in evidence for trematode infestation currently
exists in the Protobranchia lineage of bivalves. Liljedahl (1985) reported an
igloo-like trace in the praenuculid Nuculodonta gotlandica from the Silurian
of Sweden (compare Liljedahl, 1994). This trace is remarkably similar to
212 John Warren Huntley and Kenneth De Baets

Figure 2 Distribution of trematode-induced pits (P), trematode-induced igloos (I) and


igloo-like (IL) traces on bivalves found in fossil and subfossil death assemblages map-
ped onto the evolutionary time tree of Bivalvia (Adapted from Bieler et al. (2014).). Note
that geologic time is not to scale; the Cenozoic is exaggerated to show more details.
The Holocene time bin includes both sub-surface Holocene deposits and modern/Ho-
locene death assemblages. *The family Arcidae is not monophyletic in the analysis of
Bieler et al. (2014), but all members of this family are included within Pteriomorphia.
Temporal units from the oldest to the youngest are Cambrian, Ordovician, Silurian,
Devonian, Carboniferous, Permian, Triassic, Jurassic, Cretaceous, Paleogene (Paleocene,
Eocene, Oligocene), Neogene (Miocene, Pliocene), Quaternary (Pleistocene, Holocene).

the igloo structures described by Ituarte et al. (2005) in the nuculanid Yoldia
woodwardi from modern death assemblages in Argentina (Ituarte et al., 2005).
If Liljedahl’s igloo-like structures are indeed the result of trematode infesta-
tion, then these would be, by far, the oldest fossil evidence for trematodes and
parasitic flatworms in the fossil record (De Baets et al., 2015b). Currently,
more comparative work is required to determine the relatedness of the Silu-
rian igloo-like traces and trematode parasites. It cannot therefore be ruled
out, that they are related with other organisms with similar behaviour (De
Baets et al., 2015b). Significant temporal gaps in the trematode record also
Table 2 Occurrences of trematode-induced traces on fossil and subfossil bivalve death assemblages
Bivalve host

Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time


Trace type Family Species Life mode Feeding mode Environment Country Age Reference
Pits Cardiidae Cerastoderma edule Infaunal Suspension Marine Netherlands Modern Huntley NL survey
Brackish
Cerastoderma glaucum Marine
Brackish
Donacidae Donax vittatus Marine
Mactridae Spisula sp.
Mactra stultorum
Myidae Mya arenaria Brackish
Pharidae Ensis sp. Marine
Tellinidae Macoma balthica Deposit/suspension Marine
M. balthica Brackish
Angulus tenuis Suspension Marine
Cardiidae C. glaucum Italy Huntley and Scarponi
Donacidae Donax semistriatus (2015)
Lucinidae Loripes lucinalis Chemosymbiotic
Mactridae Spisula subtruncata Suspension
Mytilidae Mytilus galloprovincialis Epifaunal
Veneridae Chamalea gallina Infaunal
Igloo Neoleptonidae Neolepton bennetti e e Marine Argentina Modern Ituarte et al. (2005)
Neolepton concentricum
Nuculanidae Yoldia woodwardi
Pits Corbulidae Potamocorbula amurensis Semi-infaunal Suspension Brackish China Holocene Huntley et al. (2014)
Donacidae D. semistriatus Infaunal Suspension Marine Italy Huntley and Scarponi
(2012)
Donax sp. USA Ruiz and Lindberg
Mactridae Mulinia sp. (1989)
Tresus sp.
Myidae Mya sp.

213
(Continued)
Table 2 Occurrences of trematode-induced traces on fossil and subfossil bivalve death assemblagesdcont'd

214
Bivalve host
Trace type Family Species Life mode Feeding mode Environment Country Age Reference
Semelidae Abra sp. Deposit/suspension England
Tellinidae Macoma sp. Canada
Sweden
Veneridae Ch. gallina Suspension Italy Huntley and Scarponi
(2012)
Gemma sp. USA Ruiz and Lindberg (1989)
Canada
Gouldia sp. USA
Psephidia sp.
Transennella sp.
Venerupis sp. Sweden
Igloo Cyamiidae Cyamiomactra sp. e e Marine Argentina Holocene Ituarte et al. (2005)
Pits Mactridae Mactra sp. Infaunal Suspension Marine Sweden Pleistocene Ruiz and Lindberg

John Warren Huntley and Kenneth De Baets


Mulinia sp. USA (1989)
Rangia sp. Brackish
Psammobiidae Sanguinolaria sp. e Marine Yemen
Tellinidae Tellina sp. Deposit USA
Veneridae Anomalocardia sp. Suspension
Gemma sp.
Parastarte sp.
Protothaca sp.
Tivela sp.
Transennella sp.
Venus sp.
Donacidae Donax sp. Sweden Pliocene
Tellinidae Tellina sp. Deposit USA
Veneridae Cyclinella sp. Suspension
Gemma sp.
Parastarte sp.

Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time


Transennella sp.
Venericardia sp.
Tellinidae Arcopagia sp. Deposit/suspension France Miocene
Macoma sp. USA
Veneridae Chione sp. Suspension
Gemma sp.
Arcidae Arca (Barbatia) filigrana Epifaunal France Eocene Huntley MNdHN
Carditidae Cardita (Venericardia) planicosta Infaunal survey
Venericor clarendonensis England Todd and Harper (2011)
Glycymeridae Pectunculus (Glycymeris) pulvinatus France Huntley MNdHN survey
Lucinidae Corbis lamellosa Chemosymbiotic
Lucina caillati
Phacoides (Pseudomiltha) caillati
Phacoides (Pseudomiltha) giganteus
Phacoides concentricus
Mactridae Mactra semisulcata Suspension
Tellinidae Tellina patellaris Deposit
Tellina sp. Ruiz and Lindberg
(1989)
Veneridae Sunetta semisulcata Suspension Huntley MNdHN survey
Venus sp. Ruiz and Lindberg (1989)
Blisters Carditidae C. (Venericardia) planicosta Infaunal Suspension Marine France Eocene Huntley MNdHN survey
Tellinidae Tellina sinuata Deposit
Veneridae Cytheraea (Callista) laevigata Suspension
Meretrix (Callista) laevigata
Igloo-like Praenuculidae Nuculodonta gotlandica Infaunal Deposit Marine Sweden Silurian Liljedahl (1994)

Blank cells indicate the same content as the above cell and the dashed lines indicate that the value is unknown.

215
216 John Warren Huntley and Kenneth De Baets

occur in the oldest of bivalve lineages, Pteriomorphia. Trematode-induced


pits have been identified amongst three pteriomorphan families Arcidae
(Eocene), Glycymeridae (Eocene), and Mytilidae (Holocene). Fifty million
years separate these occurrences, and the seeming lack of trematode traces
from the Cambrian to Paleogene Periods might therefore also be partially
the result of insufficient sampling to date.
The oldest noncontroversial evidence for trematode infestation of bi-
valves is found in the Archiheterodonta lineage. Todd and Harper (2011)
presented clear evidence of trematode-induced pits in the carditid Venericor
clarendonensis from the Eocene (Ypresian) London Clay in the UK (compare
De Baets et al., 2015b). This is the only known instance of trematode infes-
tation amongst the archiheterodontids.
The majority of the instances of trematode traces have been identified
amongst 13 families in the Imparidentia. Pits have been identified amongst
12 of these families and igloos have been documented in the Cyamiidae and
Neoleptonidae families. The earliest occurrences of trematode traces in the
Imparidentia are pits in lucinid, mactrid, tellinid and venerid bivalves from
the Eocene-aged deposits (Lutetian) of the Paris Basin (Table 2). Amongst
these families, the Tellinidae and Veneridae display the most temporally
complete fossil record of trematode traces. Despite the large number of fam-
ilies displaying trematode-induced traces in the Imparidentia, the majority of
parasite occurrences are from the Pleistocene and Holocene.
The ecological distribution (i.e. life mode and feeding mode of hosts) of
trematode-induced malformations in bivalve intermediate hosts is roughly
comparable amongst living assemblages and death/fossil assemblages. As sug-
gested by Tables 1 and 2, appreciably more data are available from death and
fossil assemblages than from living assemblages. Trematode traces are most
commonly found amongst infaunal bivalve hosts, though this pattern is
more strongly expressed in death/fossil assemblages than in living ones
(Figure 3). Semi-infaunal and epifaunal taxa are less well-represented as
trematode trace-bearing hosts. Suspension feeding is the dominant feeding
mode of trace-bearing bivalve hosts (both live and dead/fossil) followed
by deposit-feeding and chemosymbiotic taxa (the latter only occurring in
the dead/fossil assemblages). These results, along with much of the other
data reported here, should be interpreted with the understanding that
much investigation remains to be done. To date, only a few research groups
have published one or more papers primarily on trematode malformations in
modern and fossil bivalves and a few more groups mention the interaction in
passing as a secondary topic.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 217

Figure 3 Life modes (a) and feeding modes (b) of bivalve hosts of trematodes (live in-
teractions and from traces in fossil and subfossil hosts).

3.2 Taphonomy and the origin of the trematodeebivalve


parasiteehost interaction
Patzkowsky and Holland (2012) argued that the primary problem in palae-
ontology is determining to what extent the fossil record can be literally inter-
preted as the history of life. Taphonomy is the study of all the processes
influencing the production of the fossil record, from the death of an organ-
ism until it is collected by a palaeontologist. Indeed, numerous taphonomic
filters introduce bias as organisms pass from the biosphere into the geosphere,
strongly influencing the record of biodiversity through deep time (Sepkoski
et al., 1981; Miller, 2000; Peters, 2005; Alroy et al., 2008; amongst many
218 John Warren Huntley and Kenneth De Baets

others); and this is no less true for the fossil record of trematodes. Trematodes
are amongst the least likely organisms to be preserved as fossils; indeed, they
have little to no body fossil record (Ruiz and Lindberg, 1989; Littlewood and
Donovan, 2003; Boucot and Poinar, 2010; De Baets et al., 2015b). Trema-
todes are small-bodied metazoans (typically submillimetre in length) that
produce neither biominerals nor recalcitrant organic material for skeletons.
The only means we have to study their occurrence in deep time are their
trace fossils (e.g. Littlewood and Donovan, 2003; Littlewood, 2006), the
characteristic pits and igloos whose growth they induce in their bivalve hosts.
Fortunately, the biomineralized skeletons of bivalves are much more robust
to the taphonomic vagaries of the fossil record than trematodes, and their
body fossil record extends to the Cambrian. Moreover, the morphology
of bivalves and the stratigraphic context of the sediments in which they
are preserved yield much information about the environment of deposition
and how the bivalves functioned in their environment (Stanley, 1970).
Despite the robust nature of the fossil record of bivalves (compared with
other taxa), it is still subject to significant taphonomic biases.
Bivalve skeletons are a composite material comprised of calcium carbon-
ate crystals (calcite and/or aragonite) and an organic matrix (Rhoads and
Lutz, 1980 and the chapters therein). The calcite polymorph is more stable
under temperature and pressure conditions at Earth’s surface than aragonite.
Indeed, aragonite often recrystallizes to calcite or preferentially dissolves dur-
ing the processes of sediment lithification and diagenesis. Anatomical detail is
typically lost during recrystallization as the primary aragonitic microstructure
recrystallizes into much larger crystals of blocky calcite. Specimens whose
aragonite dissolves can sometimes be preserved as steinkerns, or complete in-
ternal moulds formed by lithified infilling sediment. The preferential loss or
lowered resolution of aragonitic taxa has been cited as a significant factor
biasing our understanding of Phanerozoic diversity trends (Bush and
Bambach, 2004), and this factor could likely play a role in diminishing the
fossil record of trematode traces. Trematode traces are likely to be obliter-
ated by the coarse process of recrystallization, and internal moulds are not
as likely to display pits as the original shell material. Trematode pits have
been found preserved in positive relief on a Holocene steinkern (Huntley
et al., 2014) and blister pearls can potentially also be traced further back in
time (cf. De Baets et al., 2011); however, in many cases steinkerns are
comprised of sediment whose grain size is similar to or larger than the
millimetre-scale trematode pits, thereby reducing the likelihood of pre-
servation of parasite-induced malformations. Therefore, aragonitic or mixed
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 219

aragonite/calcite taxa likely have a lower preservation potential in deep


time, thereby reducing the preservation potential of the parasitic traces
they harboured in life.
The nature of the sediment in which bivalve taxa are preserved can poten-
tially bias our ability to properly interpret the fossil record as well (Hendy,
2009). Geologically younger sedimentary layers are often unlithified which
allows for the easy extraction of whole bivalve valves. In rocks older than the
Cretaceous, it is rare to find unlithified sediments. Unless one uses time-
intensive physical and, in some cases chemical (e.g. dissolution of carbonate
rock with acid to extract silicified bivalves), extraction techniques, then ob-
servations are restricted to specimens exposed on the surface of rock slabs.
Since trematode-induced malformations are restricted to the interior of
the valve, one could only collect data from specimens with exposed interiors
that are free from encrusting epibionts and lithified sediments.
The temporally disparate distribution of both aragonite preservation and
lithified sediment are likely to be contributing explanations for our current
understanding of the temporal distribution of trematode traces. The earliest
reliable indicator of trematode infestation of bivalves is in the Ypresian stage
of the Eocene (V. clarendonensis; Todd and Harper, 2011; De Baets et al.,
2015b), but it is not unreasonable to hypothesize that the interaction between
these two groups originated much earlier in time. Bivalve molluscs first
appear in the early Cambrian (Jell, 1980) and putative parasitic flatworm re-
mains are known from the late Paleozoic (Zangerl and Case, 1976; Upeniece,
1999, 2001, 2011; Dentzien-Dias et al., 2013; De Baets et al., 2015b).
Furthermore, molecular clocks place the origin of bivalves as early as the Pre-
cambrian and current parasiteehost extrapolations place the origin of parasitic
flatworms (Neodermata) in the CambrianeOrdovician (Littlewood, 2006;
De Baets et al., 2015b). The roots in deep time for both host and parasite
coupled with the diverse array of host taxa already parasitized in the Eocene
seem to suggest an earlier origin for the interaction. Perhaps the search for the
origination of trematodeebivalve parasiteehost interaction should proceed
from the perspective of the definitive host rather than the intermediate
host. Indeed, phylogenetic analyses suggest that the adoption of bivalves, gas-
tropods and polychaetes as either first or second intermediate hosts is a derived
state that has evolved multiple times. The ancestral digenean life cycle has
fewer stages and is strongly linked with marine teleost fishes as vertebrate
definitive hosts (Cribb et al., 2003). In nearly all documented cases of trem-
atodes found in association with traces in living bivalves, the traces were
induced by the Gymnophallidae (Table 1), and marine shore birds are the
220 John Warren Huntley and Kenneth De Baets

definitive hosts for gymnophallid trematodes (Ching, 1995; Galaktionov,


2006). This association would suggest that the origin of the modern
trematodeebivalve interactions might be linked to the origin and radiation
of marine shore birds somewhere between the Cretaceous and the middle
Eocene (Ruiz and Lindberg, 1989; Mayr, 2014; Smith, 2015), though this
does not preclude the possibility that trematodes infested closely related
and/or ecologically similar predecessors of modern shore birds (compare
De Baets et al., 2015b for a review).

3.3 Trematode-induced malformations as


palaeoenvironmental indicators
Ruiz and Lindberg (1989) suggested that traces of trematodeebivalve inter-
action should serve as precise palaeoenvironmental indicators of intertidal en-
vironments. Given that marine shore birds are the typical definitive host of
gymnophallid trematodes, the infested bivalve second intermediate hosts
should occur in depths shallow enough for the birds to prey upon the bi-
valves. The depth distribution of trematode malformations has not, to our
knowledge, been addressed amongst living molluscan communities, but has
received more attention in the fossil record. Huntley and Scarponi (2012)
were able to approach the question of trematode depth distribution quantita-
tively using PleistoceneeHolocene molluscan assemblages from the Po Plain
of Northern Italy. Scarponi and Kowalewski (2004) provided the environ-
mental context for these 89 samples from three cores comprised of 98 genera
and over 23,000 molluscan specimens by conducting a detrended correspon-
dence analysis (DCA), identifying a primary environmental gradient related to
depth, and calibrating this gradient using the modern depth preferences of 24
of the most common extant genera in the samples. DCA was particularly use-
ful in this case because it provides ordination scores for taxa and samples. Since
the ordination scores for taxa were strongly correlated with preferred depth, a
regression analysis allowed for the calculation of depth values for the samples.
This, in turn, allowed for a refined and independent interpretation of the sea-
level curve from these samples to complement the one derived from sedimen-
tary and stratigraphic analysis. Huntley and Scarponi (2012) examined the
same samples for trematode-induced malformations and were able to quantify
the distribution of parasitism along the depth gradient (Figure 4). Water depth
for the examined samples ranged from 0 to 14 m depth. Trematode-infested
species occurrences were restricted to between 7 and 9 m depth, contrary to
the prediction of Ruiz and Lindberg (1989) that pits would be a reliable in-
dicator of intertidal environments.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 221

Figure 4 Frequency of Holocene bivalve species occurrences along a water depth


gradient from the Po Plain of Italy. Grey columns are noninfested species occurrences
and black columns are trematode-infested species occurrences. Adapted from Huntley
and Scarponi (2012).

One of the more intriguing results from Huntley and Scarponi’s (2012)
study was that trematode prevalence was significantly higher in lower shore-
face environments during times of sea-level rise (transgression) than in com-
parable environments during sea-level fall (regression). Huntley and
Scarponi (2012) suggested that differences in sedimentation rate and salinity
stability in the lower shoreface during sea-level rise and fall could have influ-
enced trematode prevalence, but, of course, it is very difficult to disentangle
the numerous abiotic and biotic variables that change with sea-level cycles.
In a follow-up study, Huntley and Scarponi (2015) examined modern death
assemblages from 11 locations in 2 sectors separated by the Po River Delta
along the Northern Adriatic coast of Italy. North of the Po River Delta,
conditions are very much like that during Holocene transgression with bar-
rier island/lagoon/estuary complexes. Conversely, the prograding Po River
Delta and adjacent strand plains to the south are quite similar to the environ-
mental conditions during relative regression in the Holocene. The counter-
clockwise flow of surface currents in the Northern Adriatic Sea diverts the
influence of the Po Delta to the south. This influence wanes with distance
to the south and is virtually absent to the north of the delta. As predicted,
trematode prevalence values amongst modern death assemblages were
high at north of the delta, nearly absent in the vicinity of the delta and
returned to higher values much further south of the delta, beyond its
222 John Warren Huntley and Kenneth De Baets

influence (Huntley and Scarponi, 2015). Additionally, Huntley et al. (2014)


documented extensive evidence for trematode parasitism of the estuarine
bivalve Potamocorbula amurensis in the Pearl River estuarine/deltaic deposits
in China over the last 9600 years. Specifically, Huntley et al. (2014) demon-
strated that trematode prevalence was significantly higher during the first
300 years of sea-level rise than during any other phase of sea-level rise and
delta progradation (relative sea-level fall) during the Holocene in the Pearl
River Delta. They were able to rule out changing salinity and host availabil-
ity as driving factors of trematode prevalence, but were not yet able to
address other environmental factors like temperature and nutrient availabil-
ity (see Cheng and Combes, 1990 for a review on environmental factors
influencing the invasion of molluscs by parasites). In these three case studies,
ranging across environment (estuary vs marine), time (Pleistocene, Holo-
cene and Recent) and geography (northern Italy and southern China),
one finds hints of a previously unobserved modern macroecological pattern
that was predicted from palaeontological data. As sea level rises and falls on
the time scales of millennia, many biotic and abiotic factors change as well
(Patzkowsky and Holland, 2012). Responses of these environmental vari-
ables may be linear or nonlinear and their influence on the biota (e.g. the first
intermediate hosts and the definitive hosts) can seem to be unpredictable.
This especially seems to be the case when dealing with ecological time scales,
which are miniscule in comparison to geologic time, however, as illustrated
in these case studies, interpretable patterns can emerge. There is a difference,
of course, in documenting a predictable biotic response to sea-level change
and in understanding the underlying driving factors. Trematode-induced
malformations may not always indicate intertidal environments, but their
prevalence values do seem to be systematically elevated during times of
sea-level rise, a pattern which may have dire consequences for future gener-
ations in the context of anthropogenic climate change and sea-level rise
(Huntley et al., 2014). For this reason, it is important that we are able to un-
derstand the mechanics driving the increase in prevalence.

4. DETRIMENTAL EFFECTS OF TREMATODES ON


LIVING BIVALVES AND THEIR POTENTIAL
EVOLUTIONARY IMPLICATIONS
Trematodes are known to negatively affect their (mollusc) hosts in a
number of ways including castration, modified growth rates, immobilization
and, in some cases, altered behaviour making the host more susceptible to
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 223

predation (Swennen, 1969; Lim and Green, 1991; Ballabeni, 1995;


Taskinen, 1998; Hechinger et al., 2009); and are, therefore, expected to
be significant agents of natural selection. These negative impacts on individ-
uals scale up to influence population dynamics. Lauckner’s work with larval
digeneans on North Sea tidal flats (1984, 1987) demonstrated how parasitic
castration of first intermediate gastropod hosts was nearly always fatal (not so
in second intermediate hosts), targeted larger post-spawning individuals and
resulted in a seasonal reduction in mean body size giving the appearance of
negative growth. Trematode infestation also contributed to host mortality
by making individuals more susceptible to environmental stressors
(Lauckner, 1984, 1987). Trematodes affect the growth rates of their
molluscan intermediate hosts in various ways (Sousa, 1983; Sorensen and
Minchella, 2001 and references therein). Most studies have addressed the
effects of trematodes on the growth rates of gastropods as first intermediate
hosts; the stage during which trematodes are parasitic castrators. Gigantism
(Sorensen and Minchella, 1998) and stunted growth (Lafferty, 1993a,b)
have both been documented in numerous gastropod taxa, particularly
short-living freshwater species. The exact mechanism causing these changes
in growth rate is not always clear, but gigantism is likely either (1) a
nonadaptive response to parasitism wherein energy is diverted into growth
that would otherwise be spent on reproduction, (2) an adaptation of the host
to outlive the parasite or (3) an adaptation of the parasite to increase the
fitness of the host thereby increasing its own fitness (Taskinen, 1998).
Much less attention has been given to the influence of trematode parasites
on bivalve host life history. Bivalves typically serve as the second intermedi-
ate host for trematodes (the stage during which trematodes induce pits) and
occasionally as the first intermediate host (the castrating stage), so perhaps we
should not expect the same effects on bivalves as experienced by gastropods.
In a field-based experiment with freshwater Anodonta piscinalis bivalves from
Finland, Taskinen (1998) reported a density-dependent negative correlation
between parasite prevalence and bivalve growth rates. Similarly, Thieltges
(2006) demonstrated stunted growth rates in parasitized North Sea Mytilus
edulis relative to their nonparasitized counterparts. Thieltges (2006) attrib-
uted these differences to tissue disruptions, hampered ability of the bivalve
to ingest food particles and the growth of metacercarial cysts within the host.
It is clear that trematodes influence the growth rates of their hosts in
modern settings, but do these physiological responses scale up to evolu-
tionary changes in life history within lineages or are they merely expressed
at the scale of the individual? One would predict from life history theory that
224 John Warren Huntley and Kenneth De Baets

bivalves that reach first reproduction sooner (either via an increase in growth
rate or reproducing earlier at a smaller body size) would have an advantage
when dealing with trematode parasites. Ruiz and Lindberg (1989) and Ruiz
(1991) noted a statistically significant decrease in body size of the bivalve host
Transenella through the Pleistocene of California. They recognized that
trematode prevalence was positively correlated with host body size (also
recognized by Huntley (2007) and Huntley and Scarponi (2012)) and may
select for earlier first reproduction amongst bivalve hosts. Ruiz and Lindberg
(1989) interpreted the trend of decreasing body size through time as the
result of decreasing length of time to first reproduction in the ontogeny
of Transenella. However, this interpretation was based upon sizeeage rela-
tionships that need to be confirmed as consistent through geologic time.
An internal chronology can be established by investigating the presence of
seasonal variation and annual cycles in d18O (the ratio of 18O to 16O in a
sample relative to the same ratio in a standard, reported in per mil; often
an inverse proxy for temperature) values of bivalve carbonate along the
axis of maximum growth. Such seasonal variation could serve as an ‘ontoge-
netic clock’ for an individual that would enable one to calculate bivalve
longevity and growth rates through ontogeny (Jones et al., 1986; Romanek
and Grossman, 1989; Kirby et al., 1998; Goodwin et al., 2001; amongst
others). When these temperature proxy data are plotted relative to distance
from the umbo (the portion of the shell formed in the earliest stages of
ontogeny) growth rates can be quantified. Moreover, a significant decrease
in growth rate often indicates the onset of sexual maturity of marine inver-
tebrates (Romanek and Grossman, 1989). As an example, Figure 5 shows
the d18O values of shell carbonate samples (n ¼ 27) collected along the
growth axis from the umbo to the posteroventral margin of a Chamalea
gallina valve from the Northern Adriatic coast of Italy. The cycles are inter-
preted as three annual cycles and the sharp troughs likely represent growth
cessation in the winter. This specimen was live-collected and sacrificed in
February 2010; therefore, it is possible to determine the calendar years
that correspond to these annual cycles. Moreover, if we can estimate the
d18O value of the sea water in which this clam grew, it is possible to estimate
the actual temperature of the sea water (Grossman and Ku, 1986). In addi-
tion to providing an internal chronology for this individual clam, these data
suggest reduced growth in year two and an overall trend of increasing tem-
perature during the three years. Following this methodology, growth rates
and age at onset of sexual maturity of parasitized and nonparasitized individ-
uals could theoretically be approximated in order to test the influence of
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 225

Figure 5 Oxygen isotope (d18O V-PDB) profile for an individual specimen of Chamalea
gallina from the Northern Adriatic Sea (Primi Piatti sample). The cyclical nature of the
profile suggests 3 years of growth, and the sharp ‘bottoms’ of the winter portion sug-
gest a hiatus of growth during the coldest months.

trematode parasitism on the life history of hosts through geologic time and
should be a fruitful avenue for future research.

5. CONCLUDING REMARKS
There is an increasing interest in elucidating parasiteebivalve interac-
tions through deep time. Trematodes, though lacking a continuous body
fossil record, induce the growth of a number of characteristic malformations
on the interior shell walls of their bivalve hosts, which can be traced back
in the fossil record. Gymnophallid trematodes living in close association
with malformations on the interior of living bivalves have been well-docu-
mented. Trematode pits and igloos have been identified in four of the six
major lineages within the bivalve evolutionary tree spanning back as far as
the Eocene (when they appear rather suddenly, in terms of geologic time
that is, in seven families from three lineages). Igloo-like structures have
been found in Silurian Protobranchia, but further work is required to
confirm their trematode origin and to explain the subsequent 420-
million-year gap in the igloo fossil record (compare De Baets et al.,
2015b). Large gaps in the fossil record of such structures might potentially
226 John Warren Huntley and Kenneth De Baets

indicate that a different culprit (e.g. parasite) is involved in older occurrences


(cf. Boucot and Poinar, 2010, p. 28). The sudden appearance of trematode
traces in the Eocene probably does not reflect the origin of the interaction;
rather a combination of taphonomic biases and too few people looking for
these traces likely mask a much older origin. A pattern of significantly
increased prevalence values in transgressive settings (sea-level rise) seems
to be emerging from marine and estuarine settings spaced widely in space
and time; a relationship which suggests that trematode prevalence will in-
crease in estuarine settings in the coming decades and centuries in the
context of anthropogenic climate change and sea-level rise. Trematodes
are known to influence the growth rates of their intermediate hosts in
various ways, but the long-term evolutionary implications of this negative
effect can only be tested in the fossil record through the establishment of ‘in-
ternal chronologies’ within individuals from lineages through deep time.
Though the efforts of many palaeobiologists and neontologists have greatly
expanded our understanding of parasiteehost interactions amongst trema-
todes and bivalves, many questions remain unanswered and some have
not even yet been asked.

ACKNOWLEDGEMENTS
John Huntley is grateful to Kenneth De Baets and Tim Littlewood for the invitation to
contribute to this volume. We thank reviewers Michelle Casey (University of Kansas) and
David Thieltges (NIOZ, Royal Netherlands Institute for Sea Research) for their thoughtful
and constructive suggestions that improved this manuscript. Gabriel Carlier and Marie-
Madeleine Blanc-Valleron extended kind hospitality and assistance during Huntley’s visit
to the Museum National d’Histoire Naturelle in Paris. Daniele Scarponi (Universita di
Bologna) and Paola De Muro kindly prepared the Primi Piatti bivalve sample. Stable isotope
samples were collected and analyzed by Ken MacLeod’s Methods in Paleoclimatology class at the
University of Missouri (Claire Beaudoin, Jesse Broce, Shannon Haynes, Page Quinton and
Tara Selly). Funding to conduct these analyses was generously provided by the Keller
Opportunities for Excellence Fund of the Department of Geological Sciences, University
of Missouri. Franz T. F€ ursich kindly encouraged and enabled the exploratory fieldwork of
the senior author that led to this line of inquiry. Fieldwork was made possible by a generous
Ford Ka grant from the Corey and Heather Long Foundation (JWH). Funding for this proj-
ect was generously provided by the Alexander von Humboldt Stiftung (Renewed Research
Stay Grant; JWH) and the University of Missouri Research Council (SRF-14-019; JWH).

REFERENCES
Alroy, J., et al., (and 33 co-authors), 2008. Phanerozoic trends in the global diversity of ma-
rine invertebrates. Science 321, 97e100.
Ballabeni, P., 1995. Parasite-induced gigantism in a snail: a host adaptation? Funct. Ecol. 9,
887e893.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 227

Baumiller, T.K., Gahn, F.J., 2002. Fossil record of parasitism on marine invertebrates with spe-
cial emphasis on the platyceratid crinoid interaction. In: Kowalewski, M., Kelley, P.H.
(Eds.), The fossil record of predation. Yale University Press, New Haven, pp. 195e210.
Bieler, R., Mikkelsen, P.M., Collins, T.M., Glover, E.A., Gonzalez, V.L., Graf, D.L.,
Harper, E.M., Healy, J., Kawauchi, G.Y., Sharma, P.P., Staubach, S., Strong, E.A.,
Taylor, J.D., Temkin, I., Zardus, J.D., Clark, S., Guzman, A., McIntyre, E., Sharp, P.,
Giribet, G., 2014. Investigating the Bivalve Tree of Life e an exemplar-based approach
combining molecular and novel morphological characters. Invertebr. Syst. 28, 32e115.
Blake, J.A., Evans, J.W., 1973. Polydora and related genera as borers in mollusk shells and
other calcareous substrates. Veliger 15, 235e249.
Boucot, A.J., 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier,
Amsterdam.
Boucot, A.J., Poinar Jr., G.O., 2010. Fossil Behavior Compendium. CRC Press, Boca
Raton, FL.
Brett, C.E., 1978. Host-specific pit-forming epizoans on Silurian crinoids. Lethaia 11, 217e232.
Bush, A.M., Bambach, R.K., 2004. Did alpha diversity increase during the Phanerozoic? Lift-
ing the veils of taphonomic, latitudinal, and environmental biases. J. Geol. 112, 625e642.
Caceres-Martínez, J., Vasquez-Yeomans, R., 1999. Metazoan parasites and pearls in coexist-
ing mussel species: Mytilus californianus, Mytilus galloprovincialis, and Septifer bifurcatus, from
an exposed rocky shore in Baja California, Northwestern Mexico. Veliger 42, 10e16.
Campbell, D., 1985. The life cycle of Gymnophallus rebecqui (Digenea: Gymnophallidae) and
the response of the bivalve Abra tenuis to its metacercariae. J. Mar. Biol. Assoc. U.K. 65,
589e601.
Cameron, B., 1967. Fossilization of an ancient (Devonian) soft-bodied worm. Science 155,
1246e1248.
Cheng, T., Combes, C., 1990. Influence of environmental factors on the invasion of molluscs
by parasites: with special reference to Europe. In: Di Castri, F., Hansen, A.J.,
Debussche, M. (Eds.), Biological Invasions in Europe and the Mediterranean Basin.
Springer, Netherlands.
Ching, H.L., 1995. Evaluation of characters of the digenean family Gymnophallidae
Morozov, 1955. Can. J. Fish. Aquat. Sci. 52 (Suppl. 1), 78e83.
Conway Morris, S., 1981. Parasites and the fossil record. Parasitology 82, 489e509.
Cremonte, F., Ituarte, C., 2003. Pathologies elicited by the gymnophallid metacercariae of
Bartolius pierrei in the clam Darina solenoides. J. Mar. Biol. Assoc. U.K. 83, 311e318.
Cribb, T.H., Bray, R.A., Littlewood, D.T.J., 2001. The nature and evolution of the associ-
ation among digeneans, molluscs, and fishes. Int. J. Parasitol. 31, 997e1011.
Cribb, T.H., Bray, R.A., Olson, P.D., Littlewood, D.T.J., 2003. Life cycle evolution in the
Digenea: a new perspective from phylogeny. Adv. Parasitol. 54, 197e254.
De Baets, K., Keupp, H., Klug, C., 2015a. Parasites of ammonoids. In: Klug, C., Korn, D.,
De Baets, K., Kruta, I., Mapes, R.H. (Eds.), Ammonoid Paleobiology: From Anatomy to
Paleoecology. Springer, The Netherlands.
De Baets, K., Dentzian-Dias, P.C., Upeniece, I., Verneau, O., Donoghue, P.C.J., 2015b.
Constraining the deep origin of parasitic flatworms and host-interactions with fossil
evidence. Adv. Parasitol. 90, 93e135.
De Baets, K., Klug, C., Korn, D., 2011. Devonian pearls and ammonoid-endoparasite
co-evolution. Acta Palaeontol. Pol. 56, 159e180.
Dentzian-Dias, P.C., Poinar Jr., G., De Figuiredo, A.E.Q., Pacheco, A.C.L., Horn, B.L.D.,
Schultz, C.L., 2013. Tapeworm eggs in a 270 million-year-old shark coprolite. PLoS
One 8, e55007.
Dix, T.L., Karlen, D.J., Grabe, S.A., Goetting, B.K., Holden, C.M., Markham, S.E., 2005.
Spionid polychaetes as environmental indicators: an example from Tampa Bay, Florida.
In: Bortone, S.A. (Ed.), Estuarine Indicators. CRC Press, Boca Raton, FL, pp. 277e295.
228 John Warren Huntley and Kenneth De Baets

Elston, R.A., 1990. Mollusc Diseases: Guide for the Shellfish Farmer. Washington Sea Grant
Program.
Feldman, H.R., Brett, C.E., 1998. Epi- and endobiontic organisms on Late Jurassic crinoid
columns from the Negev Desert, Israel: implications for co-evolution. Lethaia 31, 57e71.
Feldmann, R.M., 1998. Parasitic castration of the crab, Tumidocarcinus giganteus Glaessner, from
the Miocene of New Zealand: coevolution with the Crustacea. J. Paleontol. 72, 493e498.
Fry, G.F., Moore, J.G., 1969. Enterobius vermicularis: 10,000-year-old human infection.
Science 166, 1620.
Gahn, F.J., Baumiller, T.K., 2003. Infestation of Middle Devonian (Givetian) camerate cri-
noids by platyceratid gastropods and its implications for the nature of their biotic
interaction. Lethaia 36, 71e82.
Galaktionov, K.V., 2006. Phenomenon of parthenogenetic metacercariae in gymnophallids
and aspects of trematode evolution. Proc. Zool. Inst. Russ. Acad. Sci. 310, 51e58.
Goodwin, D.H., Flessa, K.W., Schӧne, B.R., Dettman, D.L., 2001. Cross-calibration of
daily growth increments, stable isotope variation, and temperature in the Gulf of
California bivalve mollusk Chione cortezi: implications for paleoenvironmental analysis.
Palaios 16, 387e398.
G€otting, K.-J., 1974. Malakozoologie. Grundriss der Weichtierkunde. Fischer, Stuttgart.
G€otting, K.-J., 1979. Durch Parasiten induzierte Perlbildung bei Mytilus edulis L. (Bivalvia).
Malacologia 18, 563e567.
Grossman, E.L., Ku, T.L., 1986. Oxygen and carbon isotope fractionation in biogenic arago-
nite: temperature effects. Chem. Geol. 59, 59e74.
Hechinger, R.F., Lafferty, K.D., Mancini III, F.T., Warner, R.R., Kuris, A.M., 2009. How
large is the hand in the puppet? Ecological and evolutionary factors affecting body mass of
15 trematode parasitic castrators in their snail host. Evol. Ecol. 23, 651e667.
Hendy, A.J.W., 2009. The influence of lithification on Cenozoic marine biodiversity trends.
Paleobiology 35, 51e62.
Hermoyian, C.S., Leighton, L.R., Kaplan, P., 2002. Testing the role of competition in fossil
communities using limiting similarity. Geology 30, 15e18.
Hudson, P.J., Dobson, A.P., Lafferty, K.D., 2006. Is a healthy ecosystem one that is rich in
parasites? Trends Ecol. Evol. 21, 381e385.
Huntley, J.W., 2007. Towards establishing a modern baseline for paleopathology: trace-
producing parasites in a bivalve host. J. Shellfish Res. 26, 253e259.
Huntley, J.W., F€ ursich, F.T., Alberti, M., Hethke, M., Liu, C., 2014. A complete Holocene
record of trematode-bivalve infection and implications for the response of parasitism to
climate change. Proc. Natl. Acad. Sci. U.S.A. 111, 18150e18155.
Huntley, J.W., Kowalewski, M., 2007. Strong coupling of predation intensity and diversity
in the Phanerozoic fossil record. Proc. Natl. Acad. Sci. U.S.A. 104, 15006e15010.
Huntley, J.W., Scarponi, D., 2012. Evolutionary and ecological implications of trematode
parasitism of modern and fossil northern Adriatic bivalves. Paleobiology 38, 40e51.
Huntley, J.W., Scarponi, D., 2015. Geographic variation of parasitic and predatory traces on
mollusks in the northern Adriatic Sea, Italy: implications for the stratigraphic paleobi-
ology of biotic interactions. Paleobiology 41, 134e153.
Huntley, J.W., Yanes, Y., Kowalewski, M., Castillo, C., Delgado-Huertas, A., Iban ~ez, M.,
Alonso, M.R., Ortiz, J.E., de Torres, T., 2008. Testing limiting similarity in Quaternary
terrestrial gastropods. Paleobiology 34, 378e388.
Ituarte, C.F., Cremonte, F., Deferrari, G., 2001. Mantle-shell complex reactions elicited by
digenean metacercariae in Gaimardia trapesina (Bivalvia: Gaimardiidae) from the south-
western Atlantic Ocean and Magellan Strait. Dis. Aquat. Org. 48, 47e56.
Ituarte, C.F., Cremonte, F., Scarano, A., 2008. Tissue reaction of Tagelus plebius (Bivalvia:
Psammobiidae) against larval digeneans in mixohaline habitats connected to the south-
western Atlantic. J. Mar. Biol. Assoc. U.K. 89, 569e577.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 229

Ituarte, C., Cremonte, F., Zelaya, D.G., 2005. Parasite-mediated shell alterations in Recent
and Holocene sub-Antarctic bivalves: the parasite as modeler of host reaction. Invertebr.
Biol. 124, 220e229.
Jameson, H.L., 1903. The formation of pearls. Nature 67, 280e282.
Jell, P.A., 1980. Earliest known pelecypod on Earth e a new Early Cambrian genus from
South Australia. Alcheringa 4, 233e239.
Johannessen, O.H., 1973. Deformations of the inner shell surface of Venerupis pullastra
(Montagu) (Lamellibranchia) as a result of infection by a trematode metacercariae:
with a note of parasitism leading to parasitic castration. Sarsia 52, 117e122.
Jones, D.S., Williams, D.F., Romanek, C.S., 1986. Life history of symbiont-bearing giant
clams from stable isotope profiles. Science 231, 46e48.
Kelley, P.H., Hansen, T.A., 2003. The fossil record of drilling predation on bivalves and
gastropods. In: Kelley, P.H., Kowalewski, M., Hansen, T.A. (Eds.), PredatorePrey In-
teractions in the Fossil Record. Kluwer Academic/Plenum, New York, pp. 113e140.
Kirby, M.X., Soniat, T.M., Spero, H.J., 1998. Stable isotope sclerochronology of Pleistocene
and recent oyster shells (Crassostrea virginica). Palaios 13, 560e569.
Klompmaker, A., Artal, P., van Bakel, B.W.M., Fraaije, R.H.B., Jagt, J.W.M., 2014. Parasites
in the fossil record: a Cretaceous fauna with isopod-Infested decapod Crustaceans, infes-
tation patterns through time, and a new ichnotaxon. PLoS One 9 (3), e92551. http://
dx.doi.org/10.1371/journal.pone.0092551.
Kowalewski, M., Dulai, A., F€ ursich, F.T., 1998. A fossil record full of holes: the Phanerozoic
history of drilling predation. Geology 26, 1091e1094.
Kríz, J., 1979. Silurian Cardiolidae (Bivalvia). Sborník Geologických Ved. Palaeontologie 22,
1e160.
Kutassy, E., 1937. Die €alteste fossile Perle und Verletzungsspuren an einem triadischen
Megalodus. Math. Naturwiss. Anz. Ung. Akad. Wiss. 55, 1005e1023.
Lauckner, G., 1983. Diseases of Mollusca: Bivalvia. In: Kinne, O. (Ed.), Diseases of Marine
Animals, Introduction, Bivalvia to Scaphopoda, vol. II. Biologische Anstalt Helgoland,
Hamburg.
Lauckner, G., 1984. Impact of trematode parasitism on the fauna of a North Sea tidal flat.
Helgol€ander Meeres. 37, 185e199.
Lauckner, G., 1987. Ecological effects of larval trematode infestation on littoral marine inver-
tebrate populations. Int. J. Parasitol. 17, 391e398.
Lafferty, K.D., 1993a. Effects of parasitic castration on growth, reproduction and population
dynamics of the marine snail Cerithidea californica. Mar. Ecol. Prog. Ser. 96, 229e237.
Lafferty, K.D., 1993b. The marine snail, Cerithidea californica, matures at smaller sizes where
parasitism is high. Oikos 68, 3e11.
Liljedahl, L., 1985. Ecological aspects of a silicified bivalve fauna from the Silurian of Gotland.
Lethaia 18, 53e66.
Liljedahl, L., 1994. Silurian nuculoid and modiomorphid bivalves from Sweden. Fossil Strata
33, 1e89.
Lim, S.S.L., Green, R.H., 1991. The relationship between parasite load, crawling behaviour,
and growth rate of Macoma balthica (L.) (Mollusca, Pelecypoda) from Hudson Bay,
Canada. Can. J. Zool. 69, 2202e2208.
Littlewood, D.T.J., 2006. The evolution of parasitism in flatworms. In: Maule, A.G.,
Marks, N.J. (Eds.), Parasitic Flatworms: Molecular Biology, Biochemistry, Immunology
and Physiology.
Littlewood, D.T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
Lomovasky, B.J., Gutiérrez, J.L., Iribarne, O.O., 2005. Identifying repaired shell damage and
abnormal calcification in the stout razor clam Tagelus plebius as a tool to investigate its
ecological interactions. J. Sea Res. 54, 163e175.
230 John Warren Huntley and Kenneth De Baets

Madin, J.S., Alroy, J., Aberhan, M., F€


ursich, F.T., Kiessling, W., Kosnik, M.A., Wagner, P.J.,
2006. Statistical independence of escalatory ecological trends in Phanerozoic marine
invertebrates. Science 312, 897e900.
Mayr, G., 2014. The origins of crown group birds: molecules and fossils. Palaeontology 57,
231e242.
Miller, A.I., 2000. Conversations about Phanerozoic global diversity. Paleobiology 26
(Suppl.), 53e73.
Moodie, R.L., 1923. Paleopathology: An Introduction to the Study of Ancient Evidences of
Disease. University of Illinois Press, Urbana.
Newton, R.B., 1908. Fossil pearl-growths. J. Molluscan Stud. 8, 128e139.
Ozanne, C.R., Harries, P.J., 2002. Role of predation and parasitism in the extinction of the
inoceramid bivalves: an evaluation. Lethaia 35, 1e19.
Patzkowsky, M.E., Holland, S.M., 2012. Stratigraphic Paleobiology: Understanding the Dis-
tribution of Fossil Taxa in Time and Space. The University of Chicago Press, Chicago.
Peters, S.E., 2005. Geologic constraints on the macroevolutionary history of marine animals.
Proc. Natl. Acad. Sci. U.S.A. 102, 12326e12331.
Poinar, G., 2003. A rhabdocoel turbellarian (Platyhelminthes, Typhloplanoida) in Baltic
amber with a review of fossil and sub-fossil platyhelminths. Invertebr. Biol. 122,
308e312.
Poulin, R., Morand, S., 2000. The diversity of parasites. Q. Rev. Biol. 75, 277e293.
Presta, M.L., Cremonte, F., Ituarte, C., 2014. The fit between parasites and intermediate host
population dynamics: larval digeneans affecting the bivalve Neolepton cobbi (Galeomma-
toidea) from Patagonia. Mar. Biol. Res. 10, 494e503.
Rhoads, D.C., Lutz, R.A. (Eds.), 1980. Skeletal Growth of Aquatic Organisms: Biological
Records of Environmental Change. Topics in Geobiology, vol. 1.
Rodrigues, S.C., 2007. Biotic interactions recorded in shells of recent rhynchonelliform bra-
chiopods from San Juan Island, USA. J. Shellfish Res. 26, 241e252.
Rodrigues, S.C., Sim~ oes, M.G., Kowalewski, M., Petti, M.A.V., Nonato, E.F., Martinez, S.,
Del Rio, C.J., 2008. Biotic interaction between spionid polychaetes and bouchardiid
brachiopods: paleoecological, taphonomic and evolutionary implications. Acta Palaeon-
tol. Pol. 53, 657e668.
Romanek, C.S., Grossman, E.L., 1989. Stable isotope profiles of Tridacna maxima as environ-
mental indicators. Palaios 4, 402e413.
Ruiz, G.M., 1991. Consequences of parasitism to marine invertebrates: host evolution? Am.
Zool. 31, 831e839.
Ruiz, G.M., Lindberg, D.R., 1989. A fossil record for trematodes: extent and potential uses.
Lethaia 22, 431e438.
Savazzi, E., 1995. Parasite-induced teratologies in the Pliocene bivalve Isognomon maxillatus.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 116, 131e139.
Scarponi, D., Kowalewski, M., 2004. Stratigraphic paleoecology: Bathymetric signatures and
sequence overprint of mollusk associations from upper Quaternary sequences of the Po
Plain, Italy. Geology 32, 989e992.
Sepkoski Jr., J.J., Bambach, R.K., Raup, D.M., Valentine, J.W., 1981. Phanerozoic marine
diversity and the fossil record. Nature 293, 435e437.
Signor III, P.W., Brett, C.E., 1984. The mid-Paleozoic precursor to the Mesozoic marine
revolution. Paleobiology 10, 229e245.
Smith, N.A., 2015. Sixteen vetted fossil calibrations for divergence dating of Charadriiformes
(Aves, Neognathae). Palaeontol. Electron. 18.1.4FC, 1e18.
Sorensen, R.E., Minchella, D.J., 1998. Parasite influences on host life history: Echinostoma rev-
olutum parasitism of Lymnaea elodes snails. Oecologia 115, 188e195.
Sorensen, R.E., Minchella, D.J., 2001. Snail-trematode life history interactions: past trends
and future directions. Parasitology 123, S3eS18.
Trace Fossil Evidence of TrematodeeBivalve ParasiteeHost Interactions in Deep Time 231

Sousa, W.P., 1983. Host life history and the effect of parasitic castration on growth: a field
study of Cerithidea californica Haldeman (Gastropoda: Prosobranchia) and its trematode
parasites. J. Exp. Mar. Biol. Ecol. 73, 273e296.
Stanley, S.M., 1970. Relation of shell form to life habits of the Bivalvia. Geol. Soc. Am.
Mem. 125, 296.
Stanley, S.M., 2008. Predation defeats competition on the seafloor. Paleobiology 34, 1e21.
Stunkard, H.W., Uzmann, J.R., 1958. Studies on digenetic trematodes of the genera Gym-
nophallus and Parvatrema. Biol. Bull. Mar. Biol. Lab. Woods Hole 115, 276e302.
Swennen, C., 1969. Crawling-tracks of trematode infected Macoma balthica (L.). Neth. J. Sea
Res. 4, 376e379.
Taskinen, J., 1998. Influence of trematode parasitism on the growth of a bivalve host in the
field. Int. J. Parasitol. 28, 599e602.
Thayer, C.W., 1974. Substrate specificity of Devonian epizoa. J. Paleontol. 48, 881e894.
Thieltges, D.W., 2006. Effect of infection by the metacercarial trematode Renicola roscovita on
growth in intertidal blue mussel Mytilus edulis. Mar. Ecol. Prog. Ser. 319, 129e134.
Todd, J.A., Harper, E.M., 2011. Stereotypic boring behaviour inferred from the earliest
known octopod feeding traces: Early Eocene, southern England. Lethaia 44, 214e222.
Tyler, C.L., Leighton, L.R., 2011. Detecting competition in the fossil record: support for
character displacement among Ordovician brachiopods. Palaeogeogr. Palaeoclimatol.
Palaeoecol. 307, 205e217.
Upeniece, I., 1999. Fossil record of parasitic helminths in fishes. In: 5th International Sym-
posium on Fish Parasites: Abstracts, 154.
Upeniece, I., 2001. The unique fossil assemblage from the Lode Quarry (Upper Devonian,
Latvia). Fossil Rec. 4, 101e119.
Upeniece, I., 2011. Palaeoecology and Juvenile Individuals of the Devonian Placoderm and
Acanthodian Fishes from Lode Site, Latvia (unpublished doctoral thesis). University of
Latvia.
Vazquez, N.N., Ituarte, C., Navone, G.T., Cremonte, F., 2006. Parasites of the stout razor
clam Tagelus plebius (Psammobiidae) from the southwestern Atlantic Ocean. J. Shellfish
Res. 25, 877e886.
Vermeij, G.J., 1977. The Mesozoic marine revolution: evidence from snails, predators, and
grazers. Paleobiology 3, 245e258.
Wilson, M.A., Reinthal, E.A., Ausich, W.I., 2014. Parasitism of a new apiocrinitid crinoid spe-
cies from the Middle Jurassic (Callovian) of southern Israel. J. Paleontol. 88, 1212e1221.
Zaman, L., Meyer, J.R., Devangam, S., Bryson, D.M., Lenski, R.E., Ofria, C., 2014. Coevo-
lution drives the emergence of complex traits and promotes evolvability. PLoS Biol. 12,
e1002023.
Zangerl, R., Case, G.R., 1976. Cobelodus aculeatus (Cope), an anacanthous shark from Penn-
sylvanian black shales of North America. Palaeontogr. Abt. A 154, 107e157.
Zottoli, R.A., Carriker, M.R., 1974. Burrow morphology, tube formation, and microarch-
itecture of shell dissolution by the spionid polychaete Polydora websteri. Mar. Biol. 27,
307e316.
CHAPTER SIX

Fossil Crustaceans as Parasites


and Hosts
Adiël A. Klompmaker*, 1, Geoff A. Boxshallx
*Florida Museum of Natural History, University of Florida, Gainesville, FL, USA
x
Department of Life Sciences, Natural History Museum, London, UK
1
Corresponding author: E-mail: adielklompmaker@gmail.com

Contents
1. Introduction 234
2. Crustaceans as Hosts of Parasites 236
2.1 Fossil evidence 236
2.1.1 Isopod parasites in decapod crustaceans 236
2.1.2 Rhizocephalan barnacles in decapod crustaceans 244
2.1.3 Platyhelminthes in crustaceans 246
2.2 Equivocal fossil evidence 246
2.2.1 Ciliates on ostracods 246
2.3 Modern evidence only 246
2.3.1 Non-crustacean parasites 246
2.3.2 Crustacean parasites 247
3. Crustaceans as Parasites of Non-crustacean Hosts 249
3.1 Fossil evidence 249
3.1.1 Ascothoracidan barnacles in invertebrates 249
3.1.2 Copepods in echinoderms 252
3.1.3 Copepods in fish 253
3.1.4 Gall crabs (Cryptochiridae) in corals 259
3.1.5 Pentastomida 261
3.2 Equivocal fossil evidence 265
3.2.1 Barnacle borings attributed to Acrothoracica in marine invertebrates 265
3.2.2 Barnacles (Pyrgomatidae) in corals 266
3.2.3 Isopods (Cymothooidea) in fishes and squids 267
3.2.4 Crabs (Trapeziidae) and corals 269
3.3 Modern evidence only 269
3.3.1 Copepods 269
3.3.2 Tantulocarida 270
3.3.3 Branchiura 270
3.3.4 Ostracoda 271
3.3.5 Facetotecta 271
3.3.6 Thoracica 272
3.3.7 Malacostraca 272

Advances in Parasitology, Volume 90


© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.06.001 All rights reserved. 233
234 Adiël A. Klompmaker and Geoff A. Boxshall

4. Overview Fossil Evidence and Future Research 273


Acknowledgements 276
References 277

Abstract
Numerous crustacean lineages have independently moved into parasitism as a mode of
life. In modern marine ecosystems, parasitic crustaceans use representatives from many
metazoan phyla as hosts. Crustaceans also serve as hosts to a rich diversity of parasites,
including other crustaceans. Here, we show that the fossil record of such parasitic inter-
actions is sparse, with only 11 examples, one dating back to the Cambrian. This may be
due to the limited preservation potential and small size of parasites, as well as to pro-
blems with ascribing traces to parasitism with certainty, and to a lack of targeted
research. Although the confirmed stratigraphic ranges are limited for nearly every
example, evidence of parasitism related to crustaceans has become increasingly more
complete for isopod-induced swellings in decapods so that quantitative analyses can
be carried out. Little attention has yet been paid to the origin of parasitism in deep
time, but insight can be generated by integrating data on fossils with molecular studies
on modern parasites. In addition, there are other traces left by parasites that could
fossilize, but have not yet been recognized in the fossil record.

1. INTRODUCTION
Crustaceans can enter into a wide variety of interspecific associations
including mutualism, commensalism, phoresis, inquilinism and parasitism.
Discerning the precise nature of a close symbiotic association between two
species can be problematic, even when studying living organisms. It is espe-
cially difficult when dealing with fossils. Indeed, when considering fossil
symbiotic associations, Darrell and Taylor (1993) concluded that the terms
parasitism, mutualism and especially commensalism should be used with
caution or avoided altogether (compare Zapalski, 2011). In this chapter,
we define parasitism as a symbiotic relationship in which one organism
(the parasite) is nutritionally dependent upon another (the host) for at least
part of its life cycle and has a negative impact on the fitness of the host (cf.
Tapanila, 2008a, see also Kinne, 1980; Conway Morris, 1990; Rohde,
2005, for similar definitions). The parasitic nature of an association between
two fossil species can be inferred from (1) evidence of detrimental effects of
the inferred parasite on the host, such as reduced growth, by analogy with
related living taxa or (2) by the presence of cysts or swellings that were
inhabited by parasites. Parasitism can be further confirmed when the parasites
themselves are preserved. Surface scars caused by the attachment of sessile
Fossil Crustaceans as Parasites and Hosts 235

barnacles to the plastron and skull of turtles are known from the fossil record
(Hayashi et al., 2013), but modern turtle barnacles are not treated as parasites.
Such an association is a phoresis in which the barnacle is transported by the
host, but it is not nutritionally dependent on the host as it still uses its modi-
fied thoracic limbs for suspension feeding. Similarly, fossils of whale barnacles
such as Coronula are known (e.g. Hayashi et al., 2013), but again, these are
non-parasitic epibionts and are not the focus of this chapter.
The classification of the Crustacea is in a state of flux, although it is now
widely accepted that the Hexapoda emerged from within the Crustacea
and, therefore, that the traditional Crustacea is a paraphyletic taxon (see
Edgecombe, 2010; Giribet and Edgecombe, 2013). We follow a recent sys-
tematic review of the arthropods (Regier et al., 2010) that recognized four
main lineages, Oligostraca, Vericrustacea, Xenocarida and Hexapoda, within
a monophyletic Pancrustacea. The first three of these constitute the traditional
Crustacea and provide the focus for this chapter: only the first two are known
to contain parasites today. The Oligostraca contains two main parasitic linea-
ges, the Branchiura (e.g. Boxshall, 2005a) and Pentastomida (e.g. B€ ockeler,
2005; Christoffersen and De Assis, 2013). The Vericrustacea contains a far
greater diversity of parasitic forms within the sublineage Multicrustacea
including the wholly parasitic Tantulocarida (e.g. Boxshall, 2005b) as well
as numerous parasitic lineages within the subclasses Thecostraca (e.g. Asco-
thoracida, Rhizocephala, Thoracica) (e.g. Pérez-Losada et al., 2004; Glenner
and Hebsgaard, 2006; Rees et al., 2014), Copepoda (Boxshall, 2005c) and
Malacostraca. Within the Malacostraca, the majority of parasitic forms are pe-
racarids belonging to the Amphipoda (L€ utzen, 2005) or Isopoda (e.g. Lester,
2005; Williams and Boyko, 2012). The hosts used by parasitic crustaceans
include representatives of many aquatic metazoan phyla, from sponges and
cnidarians to chordates, including tunicates, fishes, reptiles and mammals.
Crustaceans also serve as intermediate or definitive hosts to an enormous
range of parasites, including protists (e.g. Levine, 1988), acanthocephalans
(e.g. Taraschewski, 2000), nematodes (e.g. McClelland, 2002), cestodes
(e.g. Dollfus, 1976), monogeneans (e.g. Okawachi et al., 2012) and dige-
nean trematodes (e.g. Cribb, 2005), and even other crustaceans. In the great
majority of platyhelminthecrustacean associations, the crustacean serves as
an intermediate host as, for example, in parasites of humans such as Guinea
worm (Dracunculus) and fish tapeworm (Diphyllobothrium latum), both of
which use freshwater copepods as intermediate hosts.
Crustaceans have an extensive fossil record stretching throughout the
Phanerozoic, but evidence of crustaceans serving as hosts for parasites, or
236 Adiël A. Klompmaker and Geoff A. Boxshall

exhibiting a parasitic mode of life is relatively rare. Body fossils of the para-
sites themselves are particularly rare, but evidence of palaeoparasitism may be
found more commonly in traces such as structures formed as a result of the
interaction between parasite and host. In this chapter, we aim to discuss the
fossil evidence of crustaceans as parasites and hosts, provide a brief overview
of exclusively modern examples and discuss the potential of finding their
traces in the fossil record. We also explore what molecular clock studies
reveal concerning the stratigraphic range over which such parasitic interac-
tions involving crustaceans may have occurred.
Institutional abbreviations: Geomuseum Faxe, Faxe, Denmark (OESM),
Naturhistorisches Museum Wien, Vienna, Austria (NHMW); Florida
Museum of Natural History, University of Florida, Gainesville, Florida,
USA (UF); Oertijdmuseum De Groene Poort, Boxtel, The Netherlands
(MAB k); Natural History Museum of Denmark, Geological Museum, Uni-
versity of Copenhagen, Denmark (MGUH).

2. CRUSTACEANS AS HOSTS OF PARASITES


2.1 Fossil evidence
2.1.1 Isopod parasites in decapod crustaceans
2.1.1.1 Modern evidence
Many modern decapod species exhibit a swelling in the branchial region
caused by a parasitic bopyroid isopod. Markham (1986, Table 3) found
that a total of 3.6% of species in various decapod clades were infested, but
Boyko and Williams (2009) raised this value to 4.9%, based on more com-
plete data (w12,200 instead of 7863 species). The Caridea and Anomura
both have relatively high infestation rates (12.7% and 12.0%, respectively)
compared to Brachyura (only 1.3%) (Boyko and Williams, 2009). Preva-
lence rates for bopyroids on decapod hosts are typically <30% (e.g. Rayner,
1935; O’Brien and Van Wyk, 1985; McDermott, 1991; Roccatagliata and
Lovrich, 1999; Gonzalez and Acu~ na, 2004; Dumbauld et al., 2011; Cericola
and Williams, 2015, Table 1), but rates may increase to 94% of specimens of
a population in the case of invasive bopyroids on naïve hosts (Smith et al.,
2008; Dumbauld et al., 2011).

2.1.1.2 Life cycle and parasitism


Decapod crustaceans are not the only hosts for bopyroids as life cycles typi-
cally involve two hosts. Adult females release epicaridium larvae that locate
Fossil Crustaceans as Parasites and Hosts 237

the intermediate host, typically a planktonic calanoid copepod (Baer, 1951;


Boyko and Williams, 2009, Figure 3), and moult into a microniscus larva.
After an extended period on the copepod host, the microniscus larva trans-
forms into a cryptoniscus larva that seeks out the definitive decapod host
(Dale and Anderson, 1982). This larva attaches to the gills in the branchial
chamber or to the abdomen of the decapod, develops into an adult female,
and subsequently mates with the smaller adult male. Bopyroid females feed
on haemolymph or ovarian fluids of the host (Williams and Boyko, 2012)
using their mandibles to pierce the inner cuticle (Bursey, 1978). Haemo-
lymph is also the source of nutrition for larvae on the copepod host (Boyko
and Williams, 2009). In addition to nutrients, these soft-bodied parasites also
benefit from the shelter that the decapod cuticle provides, in the case of
infestation in the branchial chamber. Negative impacts on the decapod
host include a reduced growth rate, lower fecundity leading to parasitic
castration in extreme cases, modification of secondary sex characters and
distortion of the epipodites (e.g. Tucker, 1930; Beck, 1980; Van Wyk,
1982; O’Brien and Van Wyk, 1985; McDermott, 1991; Gonzalez and
Acu~ na, 2004; Hernaez et al., 2010; Petric et al., 2010; Williams and Boyko,
2012). The effects of larval parasitism on the copepod hosts are not well
known, but may include reduced growth rate and a decrease in swimming
efficiency.

2.1.1.3 Fossil record


The fossil record of bopyroids is non-existent: no body fossils are known.
However, pronounced swellings of the branchial chamber of fossil decapods
are well known and widely accepted to represent swellings induced by para-
sitic isopods. These traces are, by far, the best known example of parasitism
in or by crustaceans from the fossil record, dating back to the Jurassic (e.g.
Hessler, 1969; Wienberg Rasmussen et al., 2008; Klompmaker et al.,
2014). Given the close morphological similarity to swellings induced by
bopyrids in modern decapods, Wienberg Rasmussen et al. (2008) referred
to them as bopyriform swellings. Boyko and Williams (2009) even specu-
lated that members of the Ioninae may have caused swellings in fossil crabs.
Klompmaker et al. (2014), however, expressed some doubt that these are all
caused by bopyrids, given the ancient age of such swellings, the lack of body
fossil evidence and the fact that other isopods may also cause swellings in the
branchial region. These authors referred to them as isopod-induced swel-
lings and formalized the name of this embedment structure by erecting
the ichnotaxon Kanthyloma crusta. They listed 88 host species, to which three
238 Adiël A. Klompmaker and Geoff A. Boxshall

further species are added here (see Figure 1) plus K. crusta in the crab Speo-
carcinus berglundi from the late Miocene (Tortonian) of California (Tucker
et al., 1994) and the crab Cristafrons praescientis from the Late Cretaceous
(SantonianeMaastrichtian) of Antarctica (Feldmann et al., 1993) based on
evidence from figures. Recently, Beschin et al. (2015) mentioned a swelling
in the branchial region of the Eocene (Ypresian) squat lobster Acanthogalathea
squamosa from Italy, but this record is difficult to confirm in the absence of a
figure. New fossil specimens exhibiting K. crusta have been reported recently

Figure 1 New examples of swellings (ichnotaxon Kanthyloma crusta) attributed to iso-


pods in the branchial region of fossil decapod crustaceans including squat lobsters (a,
d, e) and crabs (rest). (a) Galathea strigifera from the lower Palaeocene (Danian) Faxe For-
mation of the Faxe quarry, Denmark (e.g. Jakobsen and Collins, 1997; Lauridsen et al.,
2012; Lauridsen and Bjerager, 2014), OESM-10059-21721. (b) Goniodromites ?dentatus
from the Upper Jurassic (Tithonian) Ernstbrunn Limestone of the Ernstbrunn quarries,
Austria (e.g. Schweitzer and Feldmann, 2008; Robins et al., 2012, 2013; Schneider et al.,
2013), NHMW 2014/0194/0965. (c, g) Frontal and dorsal views of Cycloprosopon sp.
from the Upper Jurassic (Tithonian)eLower Cretaceous (Berriassian) part of the
Stramberk Formation of the Kotouc quarry, Czech Republic (e.g. Housa and Vasícek,
2004; Fraaije et al., 2013a), UF 252037. (d) Munida primaeva from the lower Palaeocene
(Danian) Faxe Formation of the Faxe quarry, Denmark, MGUH 31265. Specimen about
same size as (e). (e) Eomunidopsis navarrensis from the mid-Cretaceous (upper Albian)
Eguino Formation of the Koskobilo quarry, Spain (e.g. Fraaije et al., 2012, 2013b;
Klompmaker et al., 2011, 2012, 2013b; Klompmaker, 2013), MAB k2603. (f) Juvenile
Eodromites grandis from the Late Jurassic (Tithonian) Ernstbrunn Limestone of the
Ernstbrunn quarries, Austria, NHMW 1990/0041/4646. Kanthyloma crusta was previously
unknown from G. strigifera, M. primaeva and Goniodromites ?dentatus. Scale bars 5.0 mm.
Fossil Crustaceans as Parasites and Hosts 239

in taxa from which such swellings were already known (Fraaije, 2014, for
Gastrosacus wetzleri; Hyzný et al., 2014, for Galathea weinfurteri; Kornecki,
2014, for Cretocoranina testacea), highlighting the relatively common occur-
rence of such swellings. These swellings have been found mostly in fossil
true crabs (Brachyura) and squat lobsters (Galatheoidea); evidence of infes-
tation of lobsters and shrimps is relatively scarce (Wienberg Rasmussen
et al., 2008; Klompmaker et al., 2014). Only one swelling is known from
a shrimp (Franţescu, 2014), despite the common occurrence of isopods in
modern shrimps (Boyko and Williams, 2009). This can be explained by
the lower fossilization potential of shrimps due to their relatively soft
exoskeleton. Of note is the recent discovery of a swelling in the fixed finger
of a propodus of the Holo-Pleistocene ghost shrimp Glypturus panamacana-
lensis, but this swelling was probably not caused by a parasitic isopod
(Klompmaker et al., 2015).
Modification of secondary sex characters by epicaridean isopods, such as
feminization of the male chelae in shrimps (e.g. Tucker, 1930; Beck, 1980)
and widening of the abdomen in male crabs (Reinhard, 1956), has been
recorded in modern decapods, but this would be difficult to detect in fossils
unless specimens are exquisitely preserved.

2.1.1.4 Quantitative data per fauna


Quantitative data on infestation by isopods are sparse, although a few data
exist on prevalence rates by locality and/or by host taxon (Bachmayer,
1955; Housa, 1963; Radwa nski, 1972; Boucot, 1990; Wienberg Rasmussen
et al., 2008; Klompmaker et al., 2014). The latter authors collected data for
all species at a locality by systematically collecting all specimens they
encountered. They showed that more common species tend to have a
higher prevalence rate, potentially suggesting host specificity.

2.1.1.5 Infestation patterns through time


Inspired by the initial data set gathered by Wienberg Rasmussen et al.
(2008), Klompmaker et al. (2014) expanded this data set and figured infes-
tation patterns through geologic time (Figure 2). Both the number of species
exhibiting evidence of K. crusta and the percentage of species infested show a
peak during the Late Jurassic, after which values dropped to stabilize in the
Late Cretaceous and Cenozoic. Interestingly, the Late Cretaceous infesta-
tion of all species is <5%, which is comparable to today’s value (see Boyko
and Williams, 2009). This may suggest that infestation rates comparable to
today were already reached in the Late Cretaceous.
240 Adiël A. Klompmaker and Geoff A. Boxshall

Figure 2 The prevalence of parasitic, isopod-induced swellings in fossil decapod


crustaceans (ichnotaxon Kanthyloma crusta) throughout geologic time. Larger graph:
number of infested marine decapod, brachyuran, galatheoid anomuran and lobster
species standardized per 20 Myr per epoch. Inset shows the percentage of marine
decapod, brachyuran, galatheoid anomuran and lobster species infested per Jurassic
and Cretaceous epoch to circumvent the potential effect that the high species
diversity in the Late Jurassic (see Klompmaker et al., 2013a) may have on the pattern.
The peak infestation remains visible. (Modified from Klompmaker et al. (2014, Figures 5
and 6))

Klompmaker et al. (2014) suggested that the pattern through time is


likely to show one or more biological signals rather than being dominated
by collecting and reporting biases. Adaptations increasing resistance to infes-
tation, falls in diversity of infestation-prone groups (e.g. galatheoids),
copepod-related changes and/or changes in decapod size through time
may have contributed to the observed pattern. Most infested species from
the Late Jurassic originate from reef-associated settings, which is not surpri-
sing given that Klompmaker et al. (2013a, Figure 3) showed that most deca-
pods were found in reef-associated environments at that time. Given a
subsequent decline in the contribution of reef-associated decapods to overall
diversity, could the infestation peak in the Late Jurassic be solely explained
by reef abundance? This seems unlikely because coral-associated deposits
from the Late Cretaceous (Maastrichtian) of the Netherlands did not yield
a single specimen exhibiting K. crusta, and swellings in decapods in a reef
from the Palaeocene (Danian) of Denmark are rare (Klompmaker et al.,
2014; pers. obs. AAK).
Fossil Crustaceans as Parasites and Hosts 241

2.1.1.6 Age
This type of parasitism in decapods dates back to the Jurassic, possibly the
Early Jurassic (Toarcian) based on a lobster, although this is only inferred
from a drawing (Wienberg Rasmussen et al., 2008; Klompmaker et al.,
2014). Thus far, Middle Jurassic deposits have not yielded an example of
K. crusta, and this discontinuity adds to the uncertainty surrounding the Early
Jurassic record. Given the estimated occurrence of calanoid copepods as far
back as the Silurian, based on a phylogeny using morphological and mole-
cular evidence (Selden et al., 2010), and the presence of decapods since the
Devonian (e.g. Feldmann and Schweitzer, 2010; Gueriau et al., 2014; Jones
et al., 2014), Early Jurassic and older occurrences are theoretically possible as
the two host groups would have been available. Unfortunately, no rigorous
phylogenetic analyses of epicarideans including estimated divergence times
are known to us. The only study is based on limited evidence. Using three
species, Lins et al. (2012) showed that the last common ancestor of the
Bopyridae, Dajidae and Sphaeromatidae (non-parasitic) would root in the
Permian.
Some insight may come from related isopods such as the Cymothooidea
(e.g. Dreyer and W€agele, 2001; Boyko et al., 2013). Dreyer and W€agele
(2001, Figure 12) speculated that cirolanid-like cymothoids feeding on
fish evolved to become parasitic on/in fish (Cymothoidae), and also gave
rise to the Bopyridae infesting crustacean hosts, resulting in the radiation
of bopyrids. This radiation may be best expressed in the Late Jurassic
(Figure 2). The Cymothoidae have, however, no fossil record (Smit et al.,
2014), which may indicate a low preservation potential. The non-parasitic
Cirolanidae have an undisputed fossil record into the Middle Jurassic,
whereas the Cymothoida are known from the Early Triassic (Etter, 2014).
Thus, if Dreyer and W€agele’s (2001) hypothesis is correct and the fossil re-
cord is a reasonable representation of the occurrence of the Cymothoida,
then no isopod-induced swellings in decapod crustaceans are to be expected
from the Palaeozoic.

2.1.1.7 Biogeography
It is noteworthy that the oldest report of K. crusta originates from Indonesia
(although it is disputed), whereas all Late Jurassic and Early Cretaceous
occurrences are from Europe (Klompmaker et al., 2014, Table 3). A global
distribution of this type of parasitism does not occur until the Late Creta-
ceous with occurrences in Europe, North America, New Zealand and
Antarctica. Whether this European peak in the Late Jurassic and Early
242 Adiël A. Klompmaker and Geoff A. Boxshall

Cretaceous represents a sampling bias remains to be investigated. However,


given the abundant reefs in Europe during the Late Jurassic (Kiessling et al.,
1999, Figure 10) and the radiation of Brachyura and certain Anomura
(Klompmaker et al., 2013a), this peak may not be surprising.

2.1.1.8 On the erection of an ichnotaxon


In a note, Donovan (2015) cast doubt on the establishment of the ichno-
taxon K. crusta by Klompmaker et al. (2014) for the branchial swellings
exemplified in Figure 1. In his opinion, these represent pathological struc-
tures, and, as such, the ichnotaxon should not have been erected, following
Bertling et al. (2006, Table 1) who suggested that pathologies are in a ‘grey
zone’ and are ‘non-traces’. It was not addressed why these swelling do not
represent embedment structures. We welcome this opportunity to discuss
this issue further here.
Klompmaker et al. (2014, p. 12) did not specifically call these swellings
(classified as bioclaustrations) trace fossils in view of the ongoing discussion
as to whether such embedment structures should be treated as trace fossils.
Bertling et al. (2006) were of the opinion that bioclaustrations do not repre-
sent trace fossils, but many others disagree. Bioclaustrations were called trace
fossils by Palmer and Wilson (1988) and, more recently, by Tapanila and
Ekdale (2007), Bromley et al. (2008), Tapanila (2008b), McKinney (2009),
Knaust (2012), Santos et al. (2012) and C onsole-Gonella and Marquillas
(2014). The swellings in decapods were classified ichnotaxonomically
because as Bertling et al. (2006, p. 268) noted with regard to hybrid structures
resulting from borings and growth, ‘for convenience and in order to maintain
stability of names with a high ecological meaning, we advocate nomenclatu-
rally (not taxonomically) treating them as if they were ichnotaxa’. An ichno-
taxon name is convenient here and such swellings convey ecological insights
into the origin of parasitism in deep time. Boring and growth are both of note
in this case because the mandibles of isopods bore into the inner cuticle for
food extraction from the host resulting in the growth of the parasite today
(Bursey, 1978). Hybrid structures caused by parasites have been described
ichnotaxonomically (Vinn et al., 2014; Wilson et al., 2014a). Furthermore,
Wilson et al. (2014b, p. 1214) referred preserved embedment pits without
evidence of boring to an ichnotaxon (Oichnus). Similarly, Radwa nska and
Radwa nski (2005) and Boucot and Poinar (2010) called copepod-induced
swellings trace fossils.
Klompmaker et al. (2014) treated K. crusta as an embedment structure or
bioclaustration following Bertling et al. (2006, p. 267): ‘embedment
Fossil Crustaceans as Parasites and Hosts 243

structures in calcareous skeletons that are produced by an actively growing


organism around disturbing or irritating objects or living organisms’ because
the cuticle of the host continues to grow around the isopod. Such a swelling
represents a gradually formed indentation of the isopod embedding in the
substrate (¼ the cuticle of its decapod host), but in reverse so that the bulge
is formed from the inside of the branchial wall. The result is, however, the
same: the body is encased by or embedded within the host. K. crusta refers to
the (infilled) cavity that was created by the inhabitation of the isopod, not to
the cuticle surrounding it. Indeed, all the types represent internal moulds
without cuticle. Furthermore, Klompmaker et al. (2014, p. 13) referred to
the cuticle as the substrate (not the embedment structure), and they also
referred to swellings in the branchial region.
Donovan (2015) argued against erecting an ichnotaxon for a swelling
because, in his opinion, these swellings represent pathologies, and he gave
examples including the results of a mosquito bite and a gall-maker puncture,
citing Bertling et al. (2006). The latter authors noted that (2006, p. 268)
‘swollen tissue and plant reaction tissues may at best be considered ‘traces
of traces’ having an original tracemaker, whose trace (the puncture) is obli-
terated by the induced plant growth’, but also ‘if they do contain a recogni-
zable trace fossil, e.g. a boring, faecal pellets, a pupal chamber, or an exit hole
etc., they may be named as such’ in referring to an ichnotaxon. The swel-
lings described by Bertling et al. (2006) consist (nearly) entirely of animal
or plant tissue or carbonate created by the affected organism, which is the
direct result of the disease or irritation. This is very different in K. crusta.
The direct morphological result of infestation is a slower growth rate (as
recorded for modern isopod-infested decapods), a potentially thickened
cuticle and/or feminization of the claws (see above), but not the swelling it-
self, which represents the cavity of the isopod body, or the domicile of the
isopod. Unlike for the swellings described in Bertling et al. (2006) and
referred to in Donovan (2015), the isopod parasite creates space in the bran-
chial chamber in order to grow. This is not a direct expression or result of the
disease, but is the endpoint of a coevolutionary history of the embedment of
one organism within another. Therefore, the swellings in decapods do not fit
pathological structures as defined in Bertling et al. (2006) and cannot be
considered a ‘trace of a trace’: they constitute a recognizable trace of the
body of the isopod. Radwa nska and Radwa nski (2005, p. 114) made a similar
distinction between pathological and parasitic deformities in echinoid tests.
In conclusion, the swellings in the branchial region cannot be considered
pathological structures sensu Bertling et al. (2006). Rather, they represent
244 Adiël A. Klompmaker and Geoff A. Boxshall

bioclaustrations or embedment structures, and, as such, K. crusta is consi-


dered a valid ichnotaxon. Moreover, many other researchers also consider
bioclaustrations to be trace fossils.

2.1.2 Rhizocephalan barnacles in decapod crustaceans


Most modern rhizocephalan barnacles are parasitic on brachyurans and
anomurans, but some infest caridean and axiidean shrimps, stomatopods,
isopods, cumaceans and thoracican barnacles (e.g. Høeg, 1995; Høeg
et al., 2005). Infestation by rhizocephalans can cause castration of the
decapod host, feminization of males and a reduced host growth rate, to
the extent that moulting ceases and maximum size is reduced (e.g. O’Brien
and Van Wyk, 1985; Takahashi and Matsuura, 1994; Feldmann, 1998).
Infested females do not adopt male features, but males display female se-
condary sexual characters after infestation, best expressed in the form of the
claws and abdomen (Feldmann, 1998). The abdomen of a rhizocephalan-
parasitized male decapod may become broader, resembling that of a female
(O’Brien and Van Wyk, 1985) and the major claw of a male may not grow
as large, resembling a female claw (Feldmann, 1998). The parasite extracts
nutrients from the host via an embedded rootlet system that penetrates the
host’s internal tissues (see Feldmann, 1998, for additional details).
This type of parasitism is well known in modern decapods, but much less
is known from the fossil record. This is partly due to the incomplete preser-
vation of the majority of decapod specimens (i.e. disarticulation, complete
ventral sides are rare), to a lack of large samples of particular species and
perhaps to a lack of study of collections in which feminization might be
observed. Some evidence exists, however. Bishop (1974) was the first to
recognize a sexually aberrant fossil decapod, an intersex specimen of the
crab Dakoticancer overanus from the Late Cretaceous (Maastrichtian) of South
Dakota (USA). This specimen bears female gonopores, the abdomen appears
wider than in males (but narrower than in females), and the granular orna-
mentation on the sternum is the same as in males. The discovery of a second
conspecific specimen from the same locality (Bishop, 1983) confirmed that
the ornamentation on the sternum is the same as in males, the abdomen is of
intermediate width and the gonopores are most reminiscent of females (see
Bishop, 1983, Table 1). Feldmann (1998) considered that the latter would
qualify more likely as a parasitized specimen. Jones (2013) studied the Bishop
collection from the same area and found 11 intersex specimens of the same
fossil crab, and concluded that parasitism was the most likely explanation for
the occurrence of the intersex specimens. Feldmann (1998) described the
Fossil Crustaceans as Parasites and Hosts 245

best preserved examples of this type of parasitism, from specimens of the crab
Tumidocarcinus giganteus from the Miocene of New Zealand (Figure 3). A
feminized male showed a broader than usual abdomen and its major claw
was more like those of females in terms of size. He specifically attributed
this to parasitism by rhizocephalan barnacles, unlike previous accounts. It
should, however, be noted that parasitism by epicarideans (including non-
bopyroids such as entoniscids) can also result in the feminization of males,
including a widening of the abdomen and reduction in claws (Reinhard,
1956; Rasmussen, 1973), so that these infestations can ‘duplicate closely
the effects by Rhizocephala’ (Reinhard, 1956, p. 91). Careful examination
of the carapace may show slight deformations in entoniscid-infested crabs
(e.g. Miyashita, 1941; Shields and Kuris, 1985), but its expression may not

Figure 3 Normal female and male specimens of the Miocene crab (Tumidocarcinus
giganteus) from New Zealand, and a possible rhizocephalan barnacle infested conspe-
cific specimen showing feminization of the abdomen and claws. (a) Abdomen of a
mature female. (b) Ventral side of a mature male. (c) Ventral side of a juvenile female.
(d) Ventral side of a feminized male. Scale bars 10 mm. (From Feldmann (1998),
Figure 3)
246 Adiël A. Klompmaker and Geoff A. Boxshall

be obvious (e.g. Shields and Ward, 1998; Brockerhoff, 2004). Therefore,


ascribing feminization in fossil crabs to infestation by rhizocephalans is not
without uncertainty.

2.1.3 Platyhelminthes in crustaceans


Platyhelminth worms have been reported to infest Late Devonian (Frasnian)
fishes from Latvia based on the presence of hooks in 16 specimens of the pla-
coderm Asterolepis ornata and in 27 specimens of the acanthodian Lodeacan-
thus gaujicus (Upeniece, 1999, 2001, 2011; De Baets et al., 2015). Two
hooks were also found in a crustacean specimen ascribed to the Mysidacea,
but Upeniece (2001, 2011) considered the hooks too large relative to the
size of this crustacean for it to have been a host. Another hook was found
close to a Conchostraca (clam shrimp) specimen, but the inference that it
constitutes evidence of parasitism is not convincing. Modern crustaceans
can act as hosts of platyhelminthes (e.g. Fredensborg and Poulin, 2005;
Boyko and Williams, 2011), but the fact that only a single specimen out
of thousands of mysidaceans was reported to contain hooks (Upeniece,
2001) may indicate accidental parasitism of these crustaceans at best. For
comparison, over 300 A. ornata and more than 50 specimens of L. gaujicus
were found, of which 16 and 27 specimens yielded hooks.

2.2 Equivocal fossil evidence


2.2.1 Ciliates on ostracods
Weitschat and Guhl (1994) reported on stalked peritrichid ciliates (Cilio-
phora) attached on the inner part of the shell and on the epipodal appendages
of ostracods found within an Early Triassic ammonite from Spitsbergen.
Wilkinson et al. (2007) and Siveter et al. (2015) referred this association to
ectoparasitism. Given the definition of parasitism used here, we do not
consider these ciliates parasitic on the ostracods because they were inter-
preted to be filter feeding and not nutritionally dependent on their host
(see Weitschat and Guhl, 1994, Figures 9 and 10).

2.3 Modern evidence only


2.3.1 Non-crustacean parasites
Crustaceans serve as intermediate or definitive hosts to an enormous
range of unicellular and metazoan parasites and such infestations can be of
economic significance in large-scale commercial aquaculture. The produc-
tion of farmed marine penaeid crustaceans, for example, can be adversely
impacted by infections with various protists, including microsporidians,
Fossil Crustaceans as Parasites and Hosts 247

haplosporidians and gregarines (Lightner, 1993). The Acanthocephala were


once treated as a distinct phylum, but are now considered close relatives of
the rotifers within a single phylum, the Syndermata (Herlyn et al., 2003;
Wey-Fabrizius et al., 2013). Acanthocephalans typically use vertebrates as
final hosts, but many use crustaceans as intermediate hosts (e.g. Near,
2002; Weber et al., 2013). In marine species, amphipods are the commonest
intermediate host taxon, but others such as brachyuran crabs and mysida-
ceans are also used (Taraschewski, 2000). Nematodes are highly diverse
and numerous parasitic nematodes have heteroxenic life cycles that include
crustaceans as hosts (McClelland, 2002). The newly hatched larva of the
sealworm Pseudoterranova decipiens, for example, is ingested by benthic cope-
pods, amphipods or mysidaceans that in turn may be eaten by larger crusta-
ceans or small fish before being consumed by the mammalian final host.
Similarly, larval Anisakis are found in pelagic crustaceans such as copepods
and euphausiaceans that are fed on by larger nekton (fish and squid) that
form the diet of the final host. Cestodes typically use vertebrates as the final
host, but they have complex life cycles often involving a succession of two
or more intermediate hosts. Crustacean groups known to serve as either first
or second intermediate hosts include copepods, amphipods, mysidaceans,
branchiopods, barnacles, stomatopods and decapods (Dollfus, 1976). The
fish tapeworm of humans, D. latum, uses freshwater copepods as an interme-
diate host. Digenetic trematodes are also typically endoparasitic and have
similarly complex life cycles involving free-living stages as well as a succes-
sion of hosts (Cribb, 2005). While the first intermediate host is usually a
mollusk, the second can often be a crustacean that either consumes the
infective cercarial larva or is penetrated by it. In contrast, most monogeneans
are ectoparasites of aquatic vertebrates and have direct life cycles, but
Udonella is hyperparasitic on caligid copepods (Okawachi et al., 2012).
Most of these protist and platyhelminth parasites are microscopic and inhabit
the internal organs of their crustacean hosts; their impact on the host is un-
likely to leave a readily recognizable trace should the host be fossilized (see
Littlewood and Donovan, 2003; Littlewood, 2006).

2.3.2 Crustacean parasites


Crustaceans are used as hosts by parasitic crustaceans including all members
of the Tantulocarida as well as some copepods, isopods and cirripedes.
Various crustaceans serve as hosts to copepods of the family Nicothoidae
and it has been suggested that these tiny parasites are egg mimics since
they are commonly found in the marsupium of peracarids, in the brood
248 Adiël A. Klompmaker and Geoff A. Boxshall

pouch of myodocopan ostracods, or on the egg mass of decapods (Boxshall


and Halsey, 2004). This mimicry is presumed to reduce the risk of removal
by the host’s grooming behaviour. These egg mimics have little detectable
impact on the host, although a few other nicothoids that inhabit the bran-
chial chambers of their hosts are known to cause swellings or cysts (Boxshall
and Lincoln, 1983a, Figure 1(a)). Such cysts have the potential to be visible
on fossils of the hosts; however, no examples of fossil crustaceans infested by
nicothoids have yet been reported.
Marine crustaceans including copepods, ostracods and peracarid mala-
costracans are hosts to tantulocaridans. The Tantulocarida, established in
1983 (Boxshall and Lincoln, 1983b), are minute ectoparasites. The life
cycle includes an asexual and a sexual cycle that share a free living, infective
stage, the tantulus larva that is typically about 100 mm in length. In the
asexual cycle, the tantulus attaches to its crustacean host, sheds its post-
cephalic trunk and develops into a sac-like asexual female within which
develops the next generation of tantulus larvae (Boxshall and Lincoln,
1987). In the sexual cycle, after attaching to the host, the post-cephalic
trunk of the tantulus swells to form a sac within which a single adult
male or female develops. These free-swimming adults are the mating
stages, but are non-feeding (Huys et al., 1993). These tiny parasites attach
externally to their crustacean hosts using an oral disc only a few microns in
diameter, and use an oral stylet to make a single puncture in the host’s
cuticle. They are unlikely to be fossilized and cause only a minimal lesion
that would be on such a minute scale as to be unrecognizable as of tantu-
locaridan origin.
Ostracods, amphipods, euphausiaceans, copepods, barnacles, mysids and
other isopods are all known to serve as hosts to parasitic isopods of the su-
perfamily Cryptoniscoidea (e.g. Vannier and Abe, 1993, for ostracods).
The larval stages of cryptoniscoids are often ectoparasitic, while adults
may be ecto- or endoparasitic. In many cases, parasites have little discernable
impact on the host and no examples of fossil taxa from these groups have
been reported as harbouring isopod parasites. Some evidence in the fossil
record may be expected for barnacles because Hosie (2008, Figure 12(a))
showed a cryptoniscoid-induced swelling in the stalk of the modern barna-
cle Smilium zancleanum. Infection of modern decapods by members of the
Entoniscidae has been reported to cause asymmetrical carapaces (Miyashita,
1941; Shields and Kuris, 1985). Examples of this could also be expected from
the fossil record (Klompmaker et al., 2014), but may be difficult to differen-
tiate from swellings caused by bopyroids.
Fossil Crustaceans as Parasites and Hosts 249

3. CRUSTACEANS AS PARASITES OF
NON-CRUSTACEAN HOSTS
3.1 Fossil evidence
3.1.1 Ascothoracidan barnacles in invertebrates
These small barnacles (up to 8 mm in size) parasitize a variety of modern
marine hosts, particularly echinoderms (but not regular echinoids and sea cu-
cumbers) and cnidarians (gorgonians, zoanthids, scleractinians and antipatha-
rians) (e.g. Moyse, 1983; Grygier, 1996; Grygier and Høeg, 2005). Many are
endoparasitic in the main body cavity of asteroids and echinoids, but ectopa-
rasitic on anthozoans, crinoids and ophiuroids (references in Krawczy nski and
Wilson, 2011, but see Grygier and Høeg, 2005). Several ascothoracidans
(Ascothorax, Ulophysema and Dendrogaster) cause castration of their echinoderm
hosts (Grygier and Høeg, 2005). Grygier and Høeg (2005) noted that little is
known about their feeding behaviour, but basal forms have piercing and
sucking mouthparts while the more transformed genera such as Ulophysema
and Dendrogaster employ absorptive feeding, similar to rhizocephalans.
Body fossils of these barnacles are unknown, but structures attributed
to ascothoracidans have been reported from the fossil record (Figure 4;
Table 1). Voigt (1959) first recognized galls or cysts named Endosacculus
moltkiae in the Late Cretaceous octocoral Moltkia minuta that were attri-
buted to ascothoracidans because of morphological similarity to cysts of
modern representatives. Moyse (1983), for example, showed morphologi-
cally similar galls on modern octocorals. Subsequently, Voigt (1967) found
further Late Cretaceous evidence for infestation of octocorals, presumably
by an ascothoracidan, and a cyst in the early Palaeocene (Danian) hydroco-
ral Astya crassus was suggested to have been produced by parasitic crusta-
ceans (either copepods or ascothoracidans) by Bernecker and Weidlich
(2005), although their photograph (Figure 7(c-1)) does not allow identifi-
cation of the parasitic structure. Voigt (1959) was of the opinion that his
examples did not represent parasitism because the host was not particularly
damaged and continued to grow after infestation. He considered it did not
represent a true endoparasite deriving its food from the host; rather, he sug-
gested that commensalism or ‘syn€ okie’ (inhabitation of host and taking part
in meals of the host sensu K€ orner, 1979) was more appropriate. However,
research on modern ascothoracidans suggests that they are parasitic (see
Grygier and Høeg, 2005).
Fossils attributed to ascothoracidans have also been found in an echinoid
and possibly in a brachiopod. Madsen and Wolff (1965) described the
250 Adiël A. Klompmaker and Geoff A. Boxshall

Figure 4 Growth anomalies in octocorals and a brachiopod attributed to infestation by


ascothoracidan barnacles. (aed) Galls or cysts (ichnotaxon Endosacculus moltkiae) in
the octocoral Moltkia minuta from the Late Cretaceous (Maastrichtian) of the
Netherlands (a, b, d) and the Late Cretaceous (Campanian) of Sweden (c). ((b) ¼ holo-
type of E. moltkiae). (e) Possible ascothoracidan infestation (marked) in the brachiopod
Moorellina negevensis from the Middle Jurassic (late Callovian) of Israel. Scale bar (e)
0.2 mm; scale bar for others not available. ((aed) from Voigt (1959), pl. 25.7e25.9 and
26.2; (e) from Krawczyn ski and Wilson (2011), Figure 2(a))

presence of holes in a Late CretaceousePalaeocene echinoid and attributed


them to two specimens of Ulophysema based on similar holes in echinoids
produced by modern Ulophysema. There may be further evidence from echi-
noids: Bromley (2004) suspected that the two round holes in an Early Creta-
ceous echinoid attributed to boring gastropods by Kier (1981) may be the
work of ascothoracidans. Restudy of the specimen is called for. Krawczy
nski
and Wilson (2011) described a gall, possibly of ascothoracidan origin, in a
Fossil Crustaceans as Parasites and Hosts
Table 1 Structures attributed to ascothoracidan barnacles from the fossil record
Host taxon Name cyst/parasite Age Locality and country Stratigraphic unit References

Moltkia minuta Endosacculus Late Cretaceous The Netherlands ‘Maastrichter Voigt (1959)
(Octocorallia) moltkiae (late Tuffkreide’
Maastrichtian)
M. minuta E. moltkiae Late Cretaceous If€
o (Schonen), ‘Tr€
ummerkreide’, Voigt (1959)
(Octocorallia) (Campanian) Sweden rocks with
Actinocamax
mamillatus
Isis sp. Endosacculus (?) Late Cretaceous South Emba, Rocks with Voigt (1967)
(Octocorallia) najdini (late Kazakhstan Belemnella
Maastrichtian) arkhangelskii
Echinocorys Ulophysema Late Cretaceous Limfjord in North e Madsen and Wolff
(Echinoidea) ePalaeocene Jutland, Denmark (1965)
(?Maastrichtian
eDanian)*
Moorellina negevensis e Middle Jurassic Matmor Hills in Matmor Fm, Krawczy
nski and
(Brachiopoda) (late Callovian) Hamakhtesh lamberti Zone Wilson (2011)
Hagadol, Israel
* pers. comm. Sten Jakobsen to AAK.

251
252 Adiël A. Klompmaker and Geoff A. Boxshall

brachiopod from the Middle Jurassic of Israel. No ascothoracidan-infested


brachiopods are known today.
The fossil record of these traces has not yet been investigated exhaus-
tively given the limited stratigraphic range (Table 1). Also, Grygier and
Høeg (2005) suggested that galls formed by members of the Petrarcidae
(e.g. Grygier, 1990) could be recognized in fossil scleractinian corals.

3.1.2 Copepods in echinoderms


Thirteen families of copepods parasitize exclusively echinoderms, and some
representatives of several other families also use echinoderms as hosts. While
most of these copepods are ectoparasites, members of the Chordeumiidae
live within cysts of host origin or within the genital bursae of their ophiuroid
hosts and can cause swellings that are visible externally. Two families living
on echinoid hosts are known to induce their hosts to form galls. The three
species of Calverocheridae form galls in their host’s spines (Stock, 1968),
while the two species of Pionodesmotidae form galls within the test of
the host (Bonnier, 1898; Anton et al., 2013). The hosts of both families
are deep-sea echinothuriid echinoids that are atypical sea urchins in having
a soft and flexible body. Calverocherids induce the host to create galls in
spines on the adoral surface; these galls are typically asymmetrical and located
near the middle of the spine. The gall has one pore located at its distal end.
Stock (1968) described the swelling as consisting of loose calcareous material
around an ampulla inside which a single copepod is found. Pionodesmotids
are rarely reported: more than a century passed between the establishment of
the type species and the discovery of a second. Pionodesmotes domhainfharrai-
geanus lives within endocysts located in the ambulacral region on the adoral
side of the host echinoid’s test (Anton et al., 2013). These cysts have a single
opening and are not visible externally. Exocysts caused by the asterocherid
copepod Cystomyzon dimerum containing multiple openings are formed on
modern hydrocorals (Zibrowius, 1981). Fossil endocysts have not been
found yet, but exocysts on the surface of echinoid tests, referred to as the
ichnotaxon Castexia douvillei, and the so-called ‘Halloween pumpkin-
mask’ cysts resembling the exocysts in modern hydrocorals (Zibrowius,
1981) (Figure 5), have both been suggested to be treatable as trace fossils
by Radwa nska and Radwa nski (2005). Halloween pumpkin-mask cysts
are also known from fossil crinoids (Radwa nska and Radwa nski, 2005).
Other, older interpretations of the trace maker of the latter cysts exist
including trematodes, myzostomid annelids and ascothoracid barnacles
(e.g. Mehl et al., 1991; Richter, 1991; Neumann and Hostettler, 2004).
Fossil Crustaceans as Parasites and Hosts 253

Figure 5 Examples of Halloween pumpkin-mask cysts (aee) and cysts ascribed to the
ichnotaxon Castexia douvillei (fei), all inferred to be caused by copepod parasites. (a, b)
The modern hydrocoral Stylaster papuensis from Tagula Island, Louisiade Archipelago,
southwest Pacific, showing cysts induced by the copepod Cystomyzon dimerum. (c,
d) Copepod cyst in the gonopore of an echinoid test of Acrosalenia spinosa from the
Middle Jurassic (Bathonian) of Bicqueley, Lorraine, France. (e) Three cysts in a crinoid
stem of Isocrinus tuberculatus from the Early Jurassic (Sinemurian) of Germany. (fei)
Two cysts in an echinoid test of Collyrites dorsalis from the Middle Jurassic (Callovian)
of Mortagne-au-Perche, Orne, France. Smallest cyst (h) located in an ambulacral co-
lumn, whereas the larger one (i) is positioned in an interambulacral column. Scale
bars unavailable in source articles. ((a, b) from Zibrowius (1981), pl. 1.8 and 2.4; (c, d)
from Radwan ska and Poirot (2010), Figure 4; (e) from sole figure in Weinfurtner (1989);
(fei) from Radwanska and Radwan ski (2005), Figure 11.1)

A list of echinoderms inferred to be infested by copepods resulting in cysts or


swellings is presented in Table 2. Thus far, only Jurassic occurrences, chiefly
from Europe, are known.

3.1.3 Copepods in fish


A single species of parasitic copepod from a fish host is known from body
fossils: Kabatarina patersoni was recovered from two skulls of an
Table 2 Echinoderms inferred to be infested by copepods resulting in cysts or swellings
Locality and Stratigraphic

254
Host taxon Name cyst Age country unit (if known) References

Acrosalenia spinosa Halloween Middle Jurassic Bicqueley, Lorraine, Caillasse a Anabacia, Radwa
nska and
(Echinoidea) pumpkin-mask (early France convergens Poirot (2010)
e test Bathonian) Subzone
A. spinosa Halloween Middle Jurassic Luc-sur-Mer, Basse- Caillasses de Basse- Radwa
nska and
(Echinoidea) pumpkin-mask (late Bathonian) Normandie, Ecarde Poirot (2010)
e test France
Acrosalenia Halloween Middle Jurassic Calvados, France Richter (1991)
hemicidaroides pumpkin-mask (Bathonian)
(Echinoidea)
e test
Cidaroida Halloween Late Jurassic Germany Goldfuss (1829)
(Echinoidea) pumpkin-mask
e test
Plegiocidaris Halloween Late Jurassic (early Ludwag Quarry, Mehl et al. (1991)

Adiël A. Klompmaker and Geoff A. Boxshall


coronata pumpkin-mask Kimmeridgian) Oberfranken,
(Echinoidea) Germany
e test
P. coronata Halloween Late Jurassic Teuchatz Quarry, Mehl et al. (1991)
(Echinoidea) pumpkin-mask (middle Oberfranken,
e test Kimmeridgian) Germany
P. coronata Halloween Late Jurassic Obermain-Alb, Hildner (2014)
(Echinoidea) pumpkin-mask (Kimmeridgian) Germany
e test
Paracidaris Halloween Late Jurassic Aargau, Switzerland Birmenstorfer Mehl et al. (1991)
laeviuscula pumpkin-mask (middle Schichten
(Echinoidea) Oxfordian) (Argovium)
e test
?P. laeviuscula Halloween Late Jurassic Aargau, Switzerland transversarium Zone Zbinden (1991)

Fossil Crustaceans as Parasites and Hosts


(Echinoidea) pumpkin-mask (middle
e test Oxfordian)
Hemicidaris Halloween Late Jurassic Bernese Jura, bimammatum Zone Zbinden (1991)
intermedia pumpkin-mask? (middle Switzerland
(Echinoidea) Oxfordian)
e test
Stomechinus Halloween Late Jurassic Reuchenette, Neumann and
perlatus pumpkin-mask? (Oxfordian) Switzerland Hostettler (2004)
(Echinoidea)
e test
Plegiocidaris Halloween Late Jurassic (late Bielawy Quarry, Radwa
nska and
monilifera pumpkin-mask Oxfordian) Couiavia region, Radwa nski
(Echinoidea) Poland (2005)
e test
Plegiocidaris Halloween Late Jurassic Poitou, France Nicolleau and
crucifera pumpkin-mask (Oxfordian) Vadet (1995)
(Echinoidea)
e test
H. intermedia Halloween Late Jurassic Chasseral-Kette, Mehl et al. (1991)
(Echinoidea) pumpkin-mask (middle Switzerland
e test Oxfordian)
?H. intermedia Halloween Late Jurassic (early Ma1ogoszcz, Holy Upper Oolite Radwa
nska and
(Echinoidea) pumpkin-mask Kimmeridgian) Cross Mountains, Member Radwa nski
e ?test Poland (2005)
P. crucifera e Late Jurassic Near Cze˛ stochowa, Radwa
nska and
(Echinoidea) (middle Poland Radwa nski
e spine Oxfordian) (2005)

255
(Continued)
Table 2 Echinoderms inferred to be infested by copepods resulting in cysts or swellingsdcont'd
Locality and Stratigraphic

256
Host taxon Name cyst Age country unit (if known) References
Echinoidea e Halloween Middle Jurassic Israel Matmor Borszcz et al. (2014)
spine pumpkin-mask (Callovian) Formation
Isocrinus Halloween Early Jurassic Autobahn A9, near Weinfurtner (1989)
tuberculatus pumpkin-mask (Sinemurian)* G€oggelsbuch near
(Crinoidea) e Allersberg,
stem Germany
Millericrinus Halloween Late Jurassic France De Loriol (1886)
horridus pumpkin-mask (Oxfordian)
(Crinoidea) e
stem
Collyrites dorsalis Castexia douvillei Middle Jurassic Marolles, Sarthe, Lambert (1927)
(Echinoidea) (late Callovian) France
e test
C. dorsalis C. douvillei Middle Jurassic Courgeou, Orne, Mercier (1936)

Adiël A. Klompmaker and Geoff A. Boxshall


(Echinoidea) (late Callovian) France
e test
C. dorsalis C. douvillei Middle Jurassic Mortagne-au- Radwanska and
(Echinoidea) (late Callovian) Perche, Orne, Radwa nski
e test France (2005)
Collyrites acuta C. douvillei Middle Jurassic Caucasus, Solovyev (1961) and
(Echinoidea) (¼Canceripustula (Callovian) Daghestan and Solovyev and
e test nocens) Turkmenistan Markov (2013)
Collyrites ellipticus C. douvillei Middle Jurassic Wrwsowa Knobby layer Radwanska and
(Echinoidea) (Callovian) (Cze˛ stochowa), Radwa nski
e test Polish Jura, (2005)
Poland
Fossil Crustaceans as Parasites and Hosts
C. ellipticus C. douvillei Middle Jurassic Picrzchno near Knobby layer Radwa nska and
(Echinoidea) (Callovian) Klobuck, Polish Radwa nski
e test Jura, Poland (2005)
C. ellipticus C. douvillei Middle Jurassic Zawiercie, Polish Knobby layer Radwa nska and
(Echinoidea) (Callovian) Jura, Poland Radwa nski
e test (2005)
Cladocyclus None e body fossil Early to Late Serra do Araripe, Santana Formation Cressey and
gardneri (fish) copepod Cretaceous Ceara, Brazil (nodules mostly Patterson (1973)
e gill Kabatarina (Aptian from Romualdo and Cressey and
chambers pattersoni eCenomanian)x Member) Boxshall (1989)
Ventalepis e Late Devonian Pavari site on Ketleri Formation Luksevics et al.
ketleriensis (Famennian) Ciecere River, (2009)
(fish) e scale Latvia
Holoptychius sp. e Late Devonian Pavari site on Ketleri Formation Luksevics et al.
(fish) e scale (Famennian) Ciecere River, (2009)
Latvia
Psammolepis e Middle Devonian J~
oksi (Kalmetum€agi) Gauja Formation Luksevics et al.
venyukovi (fish) (Givetian) locality on Piusa (2009)
e scale River, Estonia
* Infested crinoid species and its age determined using Simms (1989).
x
Stages derived using Martill (2007).

257
258 Adiël A. Klompmaker and Geoff A. Boxshall

actinopterygian fish, Cladocyclus gardneri, which were preserved in calcareous


nodules (Table 2; Figure 6). These originate primarily from the Romualdo
Member of the Santana Formation (Cretaceous: AptianeCenomanian) (see
Martill, 2007) of the Serra do Araripe, Ceara in northern Brazil (Cressey and
Paterson, 1973; Cressey and Boxshall, 1989). It is likely that their environ-
ment was marine. Kabatarina is classified in the family Dichelesthiidae, a fa-
mily that also contains two extant genera of fish parasites, both monotypic
(Boxshall and Halsey, 2004). Kabatarina is the sole fossil representative of
the order Siphonostomatoida.
Fish are by far the most commonly reported hosts for modern parasitic
copepods, with over 2000 species known, and they can have a marked
impact on the host. The feeding activity and attachment mechanisms of
ectoparasitic copepods can cause erosion of the host’s epidermis and stimu-
late tissue proliferation and fibrotic responses in affected host tissues
(Boxshall, 1977, 2005c). Lesions apparent on Devonian fish have been
tentatively attributed to infestation by various parasites, including copepods
(Table 2). Luksevics et al. (2009, p. 342), based on the similar size and di-
mensions of the round fossulae in scales and skeletal elements of Devonian
sarcopterygian and placoderm fishes, suggested an ‘interpretation of these
structures as being produced by parasites similar to copepod crustaceans’.
In particular, Luksevics et al. (2009) inferred that perforations of scales

Figure 6 The copepod Kabatarina pattersoni found in the fish Cladocyclus gardneri from
the Early Cretaceous of Brazil. (a) Lateral view of cephalothorax showing mouthparts
and anterior swimming legs. (b) Reconstruction of adult female ((a) from Cressey and
Boxshall (1989), pl. 1.2; (b) from Cressey and Patterson (1973), Figure 1, Science 180,
1283e1285, reprinted with permission from AAAS). Scale bars 100 mm.
Fossil Crustaceans as Parasites and Hosts 259

may result from the feeding activity of ectoparasitic crustaceans, such as co-
pepods. However, other causative agents are possible so the involvement of
copepods cannot be confirmed. Lesions in fishes can have diverse causes,
making it difficult to identify in fossil fishes without preserved pathogen
remains (Petit, 2010; Petit and Khalloufi, 2012).

3.1.4 Gall crabs (Cryptochiridae) in corals


Members of the Cryptochiridae inhabit corals, living in either pits or in true
galls created by modifying the coral (e.g. Abelson et al., 1991; Carricart-
Ganivet et al., 2004; Figure 7). Verrill (1867) was the first to call a crypto-
chirid species parasitic after studying specimens of Hapalocarcinus marsupialis
modifying its coral host to form galls. In contrast, Hiro (1937) referred to
Hapalocarcinus and Cryptochirus as commensal to their coral hosts. Subse-
quently, Kropp (1986) convincingly demonstrated that three cryptochirid
species (H. marsupialis, Utinomiella dimorpha and Cryptochirus coralliodytes
(¼ Favicola rugosa)) feed on coral mucus, debris and/or coral pieces instead
of filter feeding on particles from the water. Kropp (1986) had reservations
about calling them parasites, as he considered that mucus feeding by crypto-
chirids would not represent a metabolic drain on the corals. Johnsson et al.
(2006) concurred and suggested they should not be called parasites. The key
question is to what extent mucus feeding constitutes a sufficient negative

Figure 7 Modern crescentic cryptochirid crab dwelling of Troglocarcinus corallicola in


the coral host Colpophyllia natans from Curaçao (Modified from Van der Meij (2014),
Figure 1(b)). Fossil evidence of similar-shaped dwellings has been found recently.
260 Adiël A. Klompmaker and Geoff A. Boxshall

impact on the resources of the coral to justify calling these crabs parasitic,
which is typically not specified in definitions of parasitism. Simon-Blecher
and Achituv (1997) suggested that C. coralliodytes is a parasite of faviid corals
because specimens settle on a coral polyp resulting in the death of that polyp,
they inhibit coral growth rate (as shown quantitatively) and create depres-
sions in the coral skeleton around the pits. Simon-Blecher et al. (1999)
found that the same species encouraged algal and fungal growth in and
around the pit, and that the crabs feed on coral mucus, possibly leading to
an energy loss for the coral. They also speculated that mucus removed by
crabs could decrease the efficiency of feeding by the polyps. Based on qua-
litative observations, Mohammed and Yassien (2013) considered that the
gall-forming species H. marsupialis and Opecarcinus aurantius would have no
effect on coral growth. Carricart-Ganivet et al. (2004) analysed the gut con-
tents of Troglocarcinus corallicola and found green pigments, filamentous algae,
foraminifera and some suspended material, suggesting that this species ob-
tained its food at least partially from material deposited in the depression
around its cavity. Recently, Nogueira et al. (2014) showed a positive, signi-
ficant correlation between cryptochirid (Kroppcarcinus siderastreicola) cavity
size and size of the depression around the coral Siderastrea stellata, high-
lighting the negative effect of that cryptochirid on coral growth. Recent
articles refer to cryptochirids by the all-inclusive term obligate symbionts
(e.g. Badar o et al., 2012; Van der Meij, 2014, 2015; Van der Meij and
Schubart, 2014; Canario et al., 2015), but Wei et al. (2013) mentioned
both obligate symbionts and parasites. Obvious benefits for the coral are
unknown. In conclusion, there is continuing debate as to whether crypto-
chirids are commensal or parasitic on their host corals, but the argument for
parasitism is strongest for C. coralliodytes. The fact that multiple species feed
on coral mucus and can impact coral growth indicate that at least some
cryptochirids can be considered as coral parasites.
Their fossil record was unknown until very recently, when Portell and
Klompmaker (2014) reported on characteristic crescentic domiciles attri-
buted to cryptochirids in Pleistocene corals from Florida. Subsequently,
De Angeli and Ceccon (2015) claimed the first cryptochirid body fossil,
Montemagrechirus tethysianus, from the Eocene (Ypresian) of Italy. The small
size (w5 mm) and the environment (coral reef) match the attributes of mo-
dern cryptochirids. The overall shape (longer than wide) of the carapace also
fits the Cryptochiridae, but several major differences are evident in the fron-
tal part (the posterior part exhibits few characters). First, the rostrum contains
two pronounced spines in Montemagrechirus that are not present in any of the
Fossil Crustaceans as Parasites and Hosts 261

52 currently known modern species. Moreover, the orbits of the fossil spe-
cies are directed anterolaterally, which is highly unusual for modern crypto-
chirids (only Pseudocryptochirus viridis and two deep water species of
Cecidocarcinus show a somewhat similar orbital position, 5.7%). Of further
note is the well-calcified shell, which is why Montemagrechirus was preserved.
The shell of modern cryptochirids is usually soft and can even be transparent
(Utinomi, 1944; Guinot et al., 2013, for H. marsupialis) or the posterior part
of the carapace is soft (Potts, 1915, for Cryptochirus), although Utinomi
(1944) mentioned that the carapace of Cryptochirus is rather hard. The some-
what transparent nature of cryptochirid carapaces may be observed also in
Badar o et al. (2012) and Van der Meij and Schubart (2014). Portell and
Klompmaker (2014) did not report any remains of cryptochirid shell associ-
ated with the domiciles and no other fossil cryptochirid body fossils have
been found to date, consistent with the relatively soft nature of the carapace.
The anterolateral position of the orbits in combination with the narrowing
of the carapace posteriorly in Montemagrechirus suggests that eyes would have
protruded anterolaterally from the carapace, which may not be most suitable
condition to inhabit a pit or tunnel (indeed, eyes and orbits are typically
directed forward in cryptochirids). This morphology would argue against
parasitism in corals together with the well-calcified shell that would not
be needed for a truly parasitic form living in a pit or gall. Therefore, we
do not consider this taxon further. Whether Montemagrechirus represents an
early cryptochirid (perhaps prior to becoming parasitic) or should be classi-
fied in a different family should be addressed anew.

3.1.5 Pentastomida
Molecular evidence consistently places the Pentastomida as sister group to
the Branchiura, and these two taxa together are referred to as the Ichthyos-
traca which, together with the Mystacocarida and Ostracoda, forms the Oli-
gostracan branch of the Pancrustacea (e.g. Regier et al., 2010; Oakley et al.,
2013; Giribet and Edgecombe, 2013). However, controversy still surrounds
the classification of the Pentastomida, since fossil and some morphological
data have been interpreted as evidence that they represent stem-lineage de-
rivatives of the Euarthropoda and, therefore, cannot be considered crusta-
ceans. Waloszek et al. (2006) concluded that the phylogenetic position of
the Pentastomida remained unresolved, but Castellani et al. (2011) consi-
dered their derivation from the Euarthropodan stem-lineage to be the
most plausible hypothesis. As molecular phylogenetics has moved away
from target-gene approaches towards phylogenomics and transcriptomics
262 Adiël A. Klompmaker and Geoff A. Boxshall

(see Giribet and Edgecombe, 2013), analyses of the flood of new sequence
data have continued to support a sister relationship between the Pentasto-
mida and the Branchiura within the Oligostraca (e.g. Regier et al., 2010;
Sanders and Lee, 2010; Oakley et al., 2013). There is not yet, however,
any stable estimate of divergence times between them: Sanders and Lee
(2010), using data from five genes, estimated a divergence time of 519 Ma
(but with confidence limits of 292e616 Ma), whereas Oakley et al.
(2013) arrived at a maximum estimate of 424 Ma for this split using new
transcriptome data sets plus existing nuclear genome, mitochondrial
genome, EST and ribosomal DNA data. The Oakley et al. (2013) estimate
of divergence time for the Pentastomida and Branchiura conflicts with the
suggested presence of pentastomids in the Cambrian.
Regardless of their correct taxonomic placement, we address parasitism of
this group here. Modern pentastomids, or tongue worms, are obligate para-
sites at all stages of their life cycle (B€
ockeler, 2005). Their definitive hosts are
frequently snakes, but also amphibians, turtles, crocodilians, lizards, birds and
mammals including humans; many of these same taxa, as well as freshwater
and marine fishes and insects, also serve as intermediate hosts (B€ ockeler,
2005; Poore, 2012; Christoffersen and De Assis, 2013). Modern pentastomids
are bloodsuckers (sea birds, reptiles) or feed on mucus and sloughed cells
(mammals) (B€ ockeler, 2005). Currently, 144 species and subspecies are
known (Christoffersen and De Assis, 2013); Poore (2012) listed 124 species.
From the fossil record, only 10 named species are known (Table 3;
Figure 8), all of them from the early Palaeozoic and most are based on phos-
phatized specimens. Andres (1989) was the first to suggest the resemblance of
fossil specimens to pentastomids based on Early Ordovician specimens from

Oland (Sweden), but Walossek and M€ uller (1994), Walossek et al. (1994),
Waloszek et al. (2006) and Castellani et al. (2011) described all known
CambrianeOrdovician species thus far, all presumably representing larvae.
Most fossil species are stratigraphically confined to Upper CambrianeLower
Ordovician strata (Table 3). Recently, putative adult pentastomids from the
Silurian Herefordshire Lagerst€atte were reported to be attached to ostra-
cods: such ectoparasitism is unknown in modern pentastomids (Siveter
et al., 2015). A targeted search for specimens in Palaeozoic or Mesozoic
fine-grained limestone deposits of deep-water origin should, eventually,
yield additional species. Such Konservat-Lagerst€atten, in which phosphati-
zation can take place, have yielded most of the species known thus far
(Walossek and M€ uller, 1994; Walossek et al., 1994; see Maas et al., 2006,
for discussion on preservation). Specimens may have been overlooked
Table 3 The species of Pentastomida known from the fossil record
Species Age Locality and country Stratigraphic unit References

Fossil Crustaceans as Parasites and Hosts


Aengapentastomum Late CambrianeEarly Stora Backor quarry, Ceratopyge Waloszek et al. (2006)
andresi Ordovician V€asterg€
otland, Sweden Limestone
(late Tremadocian) (reworked)
Boeckelericambria Late Cambrian Roadcut between Zone 5C Walossek and
pelturae (late Furongian) Haggården and M€uller (1994)
Marieberg, Kinnekulle,
V€asterg€
otland, Sweden
Haffnericambria Late Cambrian Trolmen at the Kinnekulle, Zone 5C Walossek and
trolmeniensis (late Furongian) V€asterg€
otland, Sweden M€uller (1994)
Heymonsicambria Late Cambrian Kinnekulle area, Zone 5 Castellani et al. (2011)
ahlgreni (late Furongian) V€asterg€
otland, Sweden
Heymonsicambria Late Cambrian Trolmen at the Kinnekulle, Zone 5C Walossek and
gossmannae (late Furongian) V€asterg€
otland, Sweden M€uller (1994)
Heymonsicambria Late Cambrian Trolmen at the Kinnekulle, Zone 5 Walossek and
kinnekullensis (late Furongian) V€asterg€
otland, Sweden M€uller (1994)
Heymonsicambria Late CambrianeEarly Stora Backor quarry, Ceratopyge Walossek and
repetskii Ordovician (late V€asterg€
otland, Sweden Limestone M€uller (1994)
Tremadocian) (?reworked)
Heymonsicambria Late Cambrian Trolmen at the Kinnekulle, Zone 5C Walossek and
scandica (late Furongian) V€asterg€
otland, Sweden M€uller (1994)
Heymonsicambria CambrianeOrdovician Green Point, Martin Point Member, Walossek et al. (1994)
taylori boundary Newfoundland, Canada Green Point
Formation
Invavita piratica* mid-Silurian (Wenlock, Herefordshire, England Volcanoclastic deposit, Siveter et al. (2015)
late Sheinwoodian to Wenlock Series
early Homerian)

263
* Putative pentastomid.
264 Adiël A. Klompmaker and Geoff A. Boxshall

Figure 8 Larvae from fossil pentastomids. (a) Lateral view of Heymonsicambria scandica
from the Late Cambrian (late Furongian) of V€astergo €tland, Sweden. (b) Ventral view of
Heymonsicambria taylori from the Late CambrianeEarly Ordovocian of Canada. (c)
Ventrolateral view of Boeckelericambria pelturae from the Late Cambrian (late Furon-
gian) of V€astergo €tland, Sweden. Scale bars 50 mm. ((a) from Walossek and M€ uller
(1994), Figure 4(a), reproduced by permission of The Royal Society of Edinburgh and Dieter
Waloszek from Transactions of the Royal Society of Edinburgh: Earth Sciences 85, 1e37; (b)
from Walossek et al. (1994), pl. 1.3).

because of their small size: the largest described CambrianeOrdovician


specimen has a length of 780 mm (the Canadian species), making them diffi-
cult to recognize with the naked eye, although the suggested Silurian spe-
cimens range from 1 to almost 4 mm in length, mainly due to a long trunk
(Siveter et al., 2015). Moreover, their generally soft bodies except for the
cuticle do not fossilize easily.
Surprisingly, all known fossil species are found in marine deposits,
whereas modern species are mostly known from terrestrial animals. Given
their very similar morphology to today’s representatives (evolutionary stasis),
Walossek and M€ uller (1994) suggested that modern representatives must
have derived from marine ancestors instead of from terrestrial arthropods.
early Palaeozoic specimens were most likely parasitic and Walossek and
M€ uller (1994) and Walossek et al. (1994) inferred that the Cambriane
Ordovician ones may have lived internally in cavities such as their host’s
gill chambers, as deduced from the limb morphology. If correct, the larvae
were already highly adapted to their hosts in the earliest Palaeozoic
Fossil Crustaceans as Parasites and Hosts 265

(Waloszek et al., 2006). Walossek and M€ uller (1994) noted that all
CambrianeOrdovician pentastomids were associated with conodonts.
Thus, it could be speculated that these early vertebrates were hosts to penta-
stomids (compare Sanders and Lee, 2010; De Baets et al., 2015). Considering
the small size of early Palaeozoic pentastomids, which are interpreted as adults
by Sanders and Lee (2010; but compare Castellani et al., 2011 for a different
view) and resemblance to larvae of extant pentastomids that infect interme-
diate hosts such as fish, it has been suggested by these authors that the small
Palaeozoic forms confined their entire life cycle to small fish-like vertebrates
present around that time (Sanders and Lee, 2010). They may have transi-
tioned to a more continental habitat along with their hosts in the Late Devo-
nian, when the amphibians arose from lobe-finned fish (e.g. Daeschler et al.,
2006), giving rise to reptiles later in the Palaeozoic. The discovery of Silurian
pentastomids on marine ostracods suggests that early pentastomids used
invertebrate taxa (arthropods) as hosts (Siveter et al., 2015). This does, how-
ever, not preclude a mid-to-late Palaeozoic terrestrialization of pentastomids
during the vertebrate radiation on land, as Siveter et al. (2015) also pointed
out. Alternatively, an invertebrate route onto land cannot be entirely
excluded either. Further pentastomid fossil evidence, preferably still associ-
ated with their host, is needed to test these speculative hypotheses.

3.2 Equivocal fossil evidence


3.2.1 Barnacle borings attributed to Acrothoracica in marine
invertebrates
In their overview of the fossil record of parasitism in marine invertebrates,
Baumiller and Gahn (2002, Table 1) listed several instances of acrothoracican
barnacle borings in Jurassic and Cretaceous belemnites (Seilacher, 1968),
Devonian platyceratid gastropods (Baird et al., 1990) and Devonian brachio-
pods (Rodriguez and Gutschick, 1977). Boucot and Poinar (2010)
mentioned many more examples of acrothoracican borings in Palaeo-
zoiceCenozoic marine invertebrates. Taylor and Wilson (2003) even
mentioned the presence of such borings in gastropods from the Ordovician,
but did not refer to them as parasites. Baumiller and Gahn’s (2002) claim of
parasitism appears ambiguous. For example, Tomlinson (1969, p. 11) stated
that ‘the (acrothoracican) cirriped can, at most, be considered a modest shell-
weakening pest, and in general does little if any harm to the host. All species
of the order collect food without taking from or giving anything of value to
the host’. Furthermore, Baird et al. (1990, p. 233) inferred ‘that the relation-
ship between the barnacles and the host gastropods was one of obligate
266 Adiël A. Klompmaker and Geoff A. Boxshall

commensalism, perhaps bordering on parasitism’. Seilacher (1969) and


Ba1uk and Radwa nski (1991) further discussed the commensal nature of
acrothoracican barnacle borings in various fossil invertebrates. Although
difficulties exist to identify commensalism in the fossil record (see Zapalski,
2011), commensalism appears the best descriptor of this symbiotic relation-
ship, in accord with Patton (1967). Therefore, it is not considered further
here.

3.2.2 Barnacles (Pyrgomatidae) in corals


Santos et al. (2012) recently described evidence for parasitism in the middle
Miocene coral Tarbellastrea reussiana by the barnacle Ceratoconcha aff. C. cos-
tata (Figure 9). The coral provided a habitat and protection from predators,
and a filter feeding location. Negative effects on the host included inhibition
of skeletal growth, modification of coral wall geometry and the secretion of
additional material in response to stress. Santos et al. (2012) indicated that
these barnacles do not feed on coral tissue, but instead use their thoracic cirri
for filter feeding purposes. Probably for this reason, Malay and Michonneau
(2014, Table 1) classified extant species of Ceratoconcha as planktotrophic
rather than parasitic. The barnacle does undoubtedly hijack part of the
food of the coral, especially for the polyps surrounding the barnacles, but
it is not nutritionally dependent on the corals themselves. Therefore, we

Figure 9 Several cross-sections of specimens of the pyrgomatid barnacle Ceratoconcha


aff. C. costata next to an example of the trace fossil Imbutichnus costatum (arrow) resul-
ting from the overgrowth of the coral Tarbellastrea reussiana of this barnacle from the
middle Miocene (LanghianeSerravalian) of Portugal. Scale bar 5 mm. (Modified from
Santos et al. (2012), Figure 3(d))
Fossil Crustaceans as Parasites and Hosts 267

consider that this relationship is better described as commensalism. Members


of this genus are known from the Oligocene to the present (Ross and
Newman, 2002), although the Oligocene occurrence was based on a per-
sonal communication only. Santos et al. (2012) indicated that pyrgomatids
were associated with corals at least since the Miocene, while they may
have appeared in the early Cenozoic, based on molecular analyses (Tsang
et al., 2014). Savignium, Pyrgoma, Cantellius and Nobia may represent an in-
termediate step from filter feeding to full parasitism. After analysing carbon
isotopic composition of these barnacles and their coral hosts, Achituv et al.
(1997) suggested that these taxa may depend on the host as a source of food
to a great extent, which was supported by the presence of coral tissue in the
stomach contents of one species. However, their cirral nets are probably still
used for filter feeding and they may have fed on expelled zooxanthellae as
well (Achituv et al., 1997). Although Ceratoconcha and the other taxa are
not widely accepted to be parasitic, there is a group within the Pyrgomatidae
that is accepted as such. Ross and Newmann (1995) erected the Hoekiini
tribe for species feeding exclusively on tissue of the coral Hydnophora with
their biting trophi (mouthparts), while the aberrant cirri do not serve to cap-
ture food. Ross and Newman (1969) suggested that the retention of vestigial
appendages in Hoekia monticulariae could mean that its coral-eating habit has
evolved recently, or that the juveniles still use them. They speculated that
the parasitic relationship developed within the last 10 Ma. A late Cenozoic
origin is corroborated by Malay and Michonneau (2014), who suggested
that hoekiines evolved quite recently given that this tribe represented a rela-
tively short branch in their phylogeny. Their fossil record is non-existent
thus far. Their relatively thin and fragile wall will not be easily preserved,
but the calcareous portion of the basis with an amoeboid outline has a better
preservation potential (Ross and Newman, 2002). Moreover, it may be
more readily recognized given the shape.

3.2.3 Isopods (Cymothooidea) in fishes and squids


Boucot and Poinar (2010) cited Bowman (1971) describing a case of possible
parasitism. Bowman (1971, p. 540) documented the isopod Palaega lamnae
from the Upper Cretaceous (Coniacian) Lower Austin Chalk of Texas
(USA) and tentatively noted that ‘the isopod could have been associated
with the Lamna [shark] either as a parasite or as a scavenger on the corpse,
but there is no evidence to support either role except for their occurrence
together’. Wieder and Feldmann (1992) questionably included this isopod
in Palaega until more complete material is discovered and concurred with
268 Adiël A. Klompmaker and Geoff A. Boxshall

the note of Bowman (1971). If placement within the Cirolanidae proves


correct, then it could be inferred that P. lamnae is a scavenger since modern
cirolanids are free-living scavengers and predators (Smit et al., 2014). The
fossil record of Cirolanidae is relatively well known (e.g. Wieder and
Feldmann, 1992; Hyzný et al., 2013) and provides some evidence of
scavenging. For example, fossil cirolanids were found within the carcass
of an actinopterygian fish from the mid-Cretaceous (Albian) of Australia
(Wilson et al., 2011), suggesting a scavenging life mode. Another inte-
resting observation is the occurrence of Palaega pisana in association with
a Pliocene ‘whale fall community’, which would have provided ample
feeding opportunities for predators and scavengers such as Palaega (Pasini
and Garassino, 2012a). A similar claim was made by Pasini and Garassino
(2012b) for Palaega sp. and P. steatopigia from the Pliocene. Claims of para-
sitic cirolanid species from the fossil record lack evidence (Palaega williamso-
nensis: Rathbun, 1935; Cirolana enigma: Wieder and Feldmann, 1992) with
the possible exception of Polz et al. (2006), who suggested that the Late
Jurassic (late Kimmeridgian) Palaega nusplingensis may have been parasitic
on account of its attachment to the ventral side of teuthoid squid
(Figure 10). However, they could not exclude the possibility that the
specimen was feeding on the squid carcass. The same applies to at least
20 cymothooids attached to a water bug from the Late Jurassic (Tithonian)
of Solnhofen, Germany, where scavenging was deemed more likely (Polz,
2004). In summary, there is no convincing evidence of parasitism by fossil
Cirolanidae including Bowman’s P. lamnae.

Figure 10 Holotype of Palaega nusplingensis from the Late Jurassic (late Kimmeridgian)
of Germany. This specimen is mentioned to be attached to the ventral side of a teu-
thoid squid. Scale bar 1 mm (Modified from Polz et al. (2006), pl. 3.2).
Fossil Crustaceans as Parasites and Hosts 269

Other modern members of the superfamily Cymothooidea are parasitic


such as Cymothoidae, but their fossil record is non-existent (Smit et al.,
2014). Etter (2014, p. 939) considered that there is no definite fossil record
of predatory or parasitic Aegidae, Tridentellidae, Corallanidae and
Cymothoidae. The first three families are mentioned to be commensal to
and/or micropredators of modern fish (Smit et al., 2014). Unless members
of these families and the Cirolanidae changed their lifestyle throughout
evolutionary history from parasitic to non-parasitic, their presence in the
fossil record does not suggest parasitism. To our knowledge, members of
the modern Cymothoidae do not leave distinct traces that would help to
identify them as parasites in the fossil record.

3.2.4 Crabs (Trapeziidae) and corals


Boucot and Poinar (2010, p. 36) listed these crabs as being potentially para-
sitic to corals, feeding on coral mucus. They were classified as obligate ecto-
parasites by Knudsen (1967). Trapeziid crabs are known to extend back to
the Eocene (Schweitzer, 2005; Beschin et al., 2007; De Angeli and Ceccon,
2013). However, their relationship is better described as mutualism (e.g.
Stimson, 1990; Schweitzer, 2005; McKeon and Moore, 2014) because
corals derive benefit as these crabs remove material from and defend the
corals from starfish predators (e.g. Glynn, 1983a,b). Therefore, this relation-
ship is not addressed further here.

3.3 Modern evidence only


3.3.1 Copepods
Nearly half of all known marine copepod species live in close symbiotic as-
sociation with other organisms. Although in many cases the precise nature of
the association between the copepod and its host has not been elucidated,
copepods are known to be parasitic on hosts representing virtually every
phylum of aquatic metazoans. Boxshall and Halsey (2004) noted records
of associated copepods from Porifera, Cnidaria, Platyhelminthes, Nemertea,
Sipunculida, Echiura, Annelida, Mollusca, Arthropoda, Bryozoa, Phoro-
nida, Brachiopoda, Echinodermata, Hemichordata and Chordata (including
urochordates as well as vertebrates). The largest copepods (body length of up
to 80 mm) are parasites of cetaceans and large fishes, but most are small (body
length 1.0e2.0 mm) and, given their lack of a calcified cuticle, have limited
fossilization potential. Parasitic copepods are known to cause visible lesions,
such as swellings, cysts or galls on certain hosts, but such macrolesions are
relatively rare. In the majority of species, the impact on the host is more
270 Adiël A. Klompmaker and Geoff A. Boxshall

subtle and highly unlikely to be detected in a fossil. Cysts and galls attributed
to infestation by parasitic copepods are known for some modern hard-
bodied host taxa such as scleractinian corals (Zibrowius, 1981; Dojiri,
1988; Buhl-Mortensen and Mortensen, 2004; Kim and Yamashiro, 2007),
echinoids (Stock, 1968, 1981, see also Section 3.1.2) and crustaceans
(Boxshall and Lincoln, 1983a), but not from others, such as the Mollusca.
No trace fossil attributed to a parasitic copepod has yet been reported
from soft-bodied hosts, such as polychaete worms or tunicates because of
the host’s low preservation potential.

3.3.2 Tantulocarida
About 35 species of tantulocaridans are currently known, either from para-
sitic stages attached to marine crustaceans, or from their free-living infective
larvae collected from the meiofauna in marine sediments from shallow
coastal waters to the deep-sea (Mohrbeck et al., 2010). Larvae are typically
about 100 mm in length. Whereas asexual adult females can attain a
maximum size of about 1.4 mm, sexual adults of both sexes are less than
1 mm in length (Boxshall and Lincoln, 1987; Huys et al., 1993). Their small
size and lack of a calcified cuticle indicate a poor fossilization potential.

3.3.3 Branchiura
Branchiuran fish lice are all external parasites of fishes (for example Figure 11),
with the exception of Dolops ranarum that uses freshwater amphibians as hosts.
Three genera, Chonopeltis, Dolops and Dipteropeltis, are exclusively freshwater
in distribution, while the largest genus, Argulus, occurs on both marine and
freshwater fishes. With the exception of Dolops that attaches by means of

Figure 11 Two adult females of Argulus japonicus parasitic on a goldfish showing a


typical lesion with a surface swelling and haemorrhaging.
Fossil Crustaceans as Parasites and Hosts 271

paired clawed limbs, branchiurans typically attach to their host using paired
suckers derived from the maxillules (Gresty et al., 1993). They feed while
on the host, but adult females detach in order to lay eggs on submerged
hard substrates. In Argulus, the eggs hatch directly into an infective larva
equipped with cephalic swimming appendages that atrophy in subsequent
parasitic stages. Once erroneously treated as a subgroup of the Copepoda,
the Branchiura is now recognized as a distinct subclass of Oligostraca. No fos-
sils are known, but the estimated suggested sister-group relationship of Bran-
chiura with the Pentastomida (Sanders and Lee, 2010; Regier et al., 2010;
Oakley et al., 2013) suggests a Cambrian origin (see Section 3.1.5). The
feeding activity of branchiurans on the epidermis of their hosts can cause le-
sions including epidermal erosion, localized swelling and haemorrhaging
(Figure 11), but such traces, even if fossilized, are very general and would
be hard to attribute to a specific parasite taxon.

3.3.4 Ostracoda
Ostracods are a diverse modern group with a rich and extensive fossil record,
but the existence of parasitic forms is equivocal. A few ostracods live in close
symbiotic associations. Members of the podocopan family Endocytheridae,
for example, inhabit the burrows of freshwater decapods, but are probably
commensals according to Hobbs and Peters (1977), and will not be treated
further. It has also been argued that the myodocopan cypridinids Vargula
parasitica and Sheina orri are parasitic on the gills and in the nasal cavities of
various elasmobranch and actinopterygian fishes (e.g. Harding, 1966), but
Cohen (1983) considered that no myodocopids were truly parasitic and
she suggested that these ostracods were scavengers on injured or unhealthy
fish. Bennett et al. (1997) examined S. orri on the gills of a shark and docu-
mented local distortion of gill lamellae accompanied by tissue damage at all
sites of ostracod attachment, but they were unable to confirm that the ostra-
cods had ingested shark tissue. The ostracods were typically found in small
pockets between adjacent gill filaments. Bennett et al. (1997) concluded
that S. orri was parasitic, but such soft tissue features would have a low fossi-
lization potential.

3.3.5 Facetotecta
The adults of the Facetotecta are as yet unknown, but the free-swimming
larval stages are widely distributed in coastal marine waters. Facetotectan
larvae, the so-called Y-larvae, have their body enclosed within a bivalved
carapace (Glenner et al., 2008), but it is not calcified and no fossils are known
as a result. The hosts of the presumed parasitic adults are unknown.
272 Adiël A. Klompmaker and Geoff A. Boxshall

3.3.6 Thoracica
A scattered assortment of pedunculate thoracican barnacles has switched
from their ancestral suspension feeding to a parasitic mode of life. Anelasma
squalicola is the sole representative of the family Anelasmatidae and is typi-
cally found embedded in the skin of deep-sea sharks behind the dorsal fin
(Rees et al., 2014; Leung, 2014). A recent molecular clock study (Rees
et al., 2014) places the split of Anelasma from Capitulum at 126.5 Ma. Rhizo-
lepas species are parasitic on marine polychaete hosts (Day, 1939). Both
Rhizolepas and Anelasma have an atrophied suspension-feeding apparatus
and depend, at least in part, on nutrients absorbed from their host via an
embedded rootlet system. Heteralepadomorph barnacles typically live in as-
sociation with hosts, many inhabiting the gills of decapods, and they typi-
cally have reduced or absent valves, so their fossilization potential is
presumably lower than for typical barnacles. Most heteralepadomorph bar-
nacles are likely commensal, but Koleolepas species participate in a tripartite
symbiosis, living on gastropod shells inhabited by hermit crabs and carrying
sea anemones. They are typically attached beneath the pedal disc of the
anemone and feed by cropping its tentacles (Yusa and Yamato, 1999).
The shell plates are reduced or even absent in these thoracicans and their
fossilization potential is presumably lower than typical sessile or stalked
barnacles.

3.3.7 Malacostraca
The whale lice (Cyamidae) are dorsoventrally flattened amphipods found on
various cetacean hosts. Cyamids were traditionally treated as a wholly para-
sitic infraorder of the amphipod suborder Caprellidea (cf. Martin and Davis,
2001), but are now treated only as a family of parasites contained within a
much large clade (infraorder Corophiida) of free-living amphipods that
are predominantly detritivores (Myers and Lowry, 2003; Lowry and Myers,
2013). Whale lice cannot swim and lack any planktonic dispersal phase in
their life cycle; it is therefore presumed that infestation spreads only during
intraspecific bodily contact between host individuals.
Gnathiid isopods have free-living, non-feeding adults that inhabit ca-
vities in muddy sediments, in dead barnacles or in sponges, but the juveniles,
praniza larvae, feed on the blood of elasmobranch and actinopterygian fishes.
Engorged praniza larvae drop off their hosts between blood meals, rather as a
terrestrial tick. They have a very distinctive appearance, but their thin non-
calcified cuticle, which allows for their gorging feeding behaviour, is likely
to limit their fossilization potential.
Fossil Crustaceans as Parasites and Hosts 273

4. OVERVIEW FOSSIL EVIDENCE AND FUTURE


RESEARCH
Fossil evidence of parasitism in crustaceans is confined primarily to
decapod crustaceans (Figure 12). Swellings in decapod crustaceans caused
by infestation by isopods are by far the best represented evidence of para-
sitism in or by crustaceans. Ninety-three infested species and hundreds of
infested specimens have been found (e.g. Wienberg Rasmussen et al.,
2008; Klompmaker et al., 2014; herein). Nearly every other example of
parasitism is represented by only few specimens and/or is known from a
limited confirmed stratigraphic range (black in Figure 12). The stratigraphic
range of many other examples can be greatly extended because of their
modern existence, assuming that parasitism occurred in the time in between
the fossil and modern occurrences. This applies especially for examples of
parasitism based on body fossils. Platyhelminth hooks in crustaceans and
the inferred Devonian occurrence of copepods on fish are not without
doubt, however (see above). Excellent preservation certainly has helped
in the case of Palaeozoic pentastomids and the Cretaceous fish parasite,
the copepod Kabatarina. These examples have in common that the evidence
for parasitism is small in size compared to the other examples in Figure 12.
The length of the only copepod body fossil in a fish is w1.0 mm, the length
of larval pentastomids is 0.78 mm and platyhelminth hooks are only
0.4 mm long (Cressey and Boxshall, 1989; Walossek et al., 1994; Upe-
niece, 2001). This small size explains the spotty occurrence of this type of
parasitism in combination with rare circumstances allowing preservation
of such fossils. The lack of targeted research may also be a factor.
Many opportunities for additional research exist. Even though swellings
in decapod crustaceans are the most common evidence of parasitism
involving crustaceans, examples of K. crusta are as yet not documented for
many stratigraphic stages and even some epochs. Moreover, very limited
quantitative data exist per locality, highlighting the need for additional
research by systematically collecting all available specimens. Whether abun-
dant species are more heavily infested than rare species, as was suggested by
Klompmaker et al. (2014) for the Koskobilo (Spain) locality, should be
investigated further by focusing on a locality where specimens exhibiting
K. crusta are common. The discovery of additional infested taxa will help
to refine and re-evaluate the trends shown in Figure 2. Less quantitative,
but equally useful, additional evidence can be gained from the reporting
of new examples of parasitism discussed here given the spotty distribution
274 Adiël A. Klompmaker and Geoff A. Boxshall
Figure 12 Stratigraphic ranges of examples of parasitism in or by crustaceans as recognized herein. Equivocal fossil evidence (3.2) is not

Fossil Crustaceans as Parasites and Hosts


included. Grey bars represent probable occurrences based on modern or bracketing fossil occurrences in that host. The resolution used
for determining first and last occurrences is the stage level if known (see text and tables for details). Horizontal stippled lines mark boun-
daries between the Phanerozoic eras. A question mark refers to uncertain evidence of parasitism of an example that has confirmed occur-
rences. Categories in bold, italic font are based on body fossil evidence of parasites, whereas the rest is based on structures (swellings, cysts,
galls, domiciles and intersex evidence) induced by the inferred parasite. The Devonian occurrences of copepods in fish are based on round
traces on scales instead of copepod body fossils as in the Cretaceous. Timescale (left part) produced with TSCreator 6.3 (http://www.
tscreator.org).

275
276 Adiël A. Klompmaker and Geoff A. Boxshall

of confirmed occurrences shown in Figure 12. Systematic surveying of exist-


ing collections and the collection of additional material in the field are likely
to yield results. It is conceivable that the stratigraphic ranges of many exam-
ples could be expanded considerably. This would help to pinpoint the
earliest occurrence of parasitism for each group and may also serve as calibra-
tion points for molecular studies of particular parasite clades. Furthermore,
not much is known about the prevalence of parasitism in nearly every
example mentioned in Figure 12 except for isopod infestations in decapods.
Such data may help to evaluate the commonness of parasitism and its
possible effect on the evolution of the infested group.
A strategy to increase the number of fossil examples of parasitism
involving crustaceans is to identify modern traces of parasitism that do not
yet have a fossil record. Many modern parasites stimulate localized tissue res-
ponses in their hosts and some of the examples identified herein involve in-
ternal skeletal elements or hard external body coverings that have potential
for fossilization and could be the focus of targeted research. We recognize
that size is an issue so special emphasis should be placed on submillimetre
and millimetre scale features. Other traces of parasitism are rather generic,
local small pits, and will remain difficult to attribute to a particular parasite
taxon. It is apparent from Figure 12 that there is a long history of parasitism
involving crustaceans. Divergence times indicating the putative origin of
parasitism are available for some crustaceans, for example, the thoracican
barnacles (Rees et al., 2014; Tsang et al., 2014), but similar studies are
lacking for other large taxa such as the copepods. There is a continuing
need for such molecular studies to complement palaeontological research
focused on fossil parasites and their traces.

ACKNOWLEDGEMENTS
We are very grateful to Kenneth De Baets and Tim Littlewood for the invitation to
contribute to this volume, for their useful comments, and the former also for useful literature
suggestions. We also thank John Jagt and Andrei Solovjev for providing useful literature;
G€unter Schweigert for useful literature and Figure 5(e), Barry van Bakel for discussion about
the specific placement of the Goniodromites specimen; Dieter Waloszek for Figure 8(a) and (c);
Sten Jakobsen for information on Echinocorys from Denmark; Rodney Feldmann for the sug-
gestion to work on this paper; Andreas Kroh for museum numbers for the Ernstbrunn spe-
cimens; Jesper Milan for the loan of a specimen from the Geomuseum Faxe; Roger Portell,
Sean Roberts and Alex Kittle for assistance with UF collections; René Fraaije for a museum
number of Oertijdmuseum De Groene Poort; Sancia van der Meij for Figure 6; Cristina
Robins for reading and commenting on part of the text. Mark Wilson and Joachim Haug
also kindly commented on earlier draft of this manuscript; their comments are much appre-
ciated. Our gratitude goes to all these persons. Furthermore, the following publishers,
Fossil Crustaceans as Parasites and Hosts 277

organisations, journals and persons are acknowledged for permission to reproduce various
images: PLOS ONE, The Paleontological Society, Springer, AAAS, Acta Geologica Polo-
nica, Contributions to Zoology, Micro Press, Rosenstiel School of Marine and Atmospheric
Science of the University of Miami, Dieter Waloszek, NRC Research Press, The Royal
Society of Edinburgh, The Palaeontological Association and Staatliches Museum f€
ur Natur-
kunde. This work was in part supported by the Jon L. and Beverly A. Thompson Endow-
ment Fund, an Arthur James Boucot Research Grant (Paleontological Society) and a
COCARDE Workshop Grant (ESF) to A.A.K. This is University of Florida Contribution
to Paleobiology 687.

REFERENCES
Abelson, A., Galil, B.S., Loya, Y., 1991. Skeletal modification in stony corals caused by
indwelling crabs: hydrodynamical advantages for crab feeding. Symbiosis 10, 233e248.
Achituv, Y., Brickner, I., Erez, J., 1997. Stable carbon isotope ratios in Red Sea barnacles
(Cirripedia) as an indicator of their food source. Mar. Biol. 130, 243e247.
Andres, D., 1989. Phosphatisierte Fossilien aus dem unteren Ordoviz von S€ udschweden.
Berl. Geowiss. Abh. A 106, 9e19.
Anton, R.F., Stevenson, A., Schwabe, E., 2013. Description of a new abyssal copepod asso-
ciated with the echinoid Sperosoma grimaldii Koehler, 1897. Spixiana 36, 201e210.
Bachmayer, F., 1955. Die fossilen Asseln aus den Oberjuraschichten von Ernstbrunn in Nie-
der€ €
osterreich und von Stramberg in M€ahren. Sitzungsber. Osterr. Akad. Wiss. Math.
Naturwiss. Kl. 164 (1), 255e273.
Badaro, M.F., Neves, E.G., Castro, P., Johnsson, R., 2012. Description of a new genus of
Cryptochiridae (Decapoda: Brachyura) associated with Siderastrea (Anthozoa: Scleracti-
nia), with notes on feeding habits. Sci. Mar. 76, 517e526.
Baer, J.G., 1951. Ecology of Animal Parasites. University of Illinois Press, Urbana.
Baird, G.C., Brett, C.E., Tomlinson, J.T., 1990. Host-specific acrothoracid barnacles on
Middle Devonian platyceratid gastropods. Hist. Biol. 4, 221e244.
Ba1uk, W., Radwa nski, A., 1991. A new occurrence of fossil acrothoracican cirripedes: Try-
petesa polonica sp. n. in hermitted gastropod shells from the Korytnica Basin (Middle
Miocene; Holy Cross Mountains, Central Poland), and its bearing on behavioral evolu-
tion of the genus Trypetesa. Acta Geol. Pol. 41, 1e36.
Baumiller, T.K., Gahn, F.J., 2002. Fossil record of parasitism on marine invertebrates
with special emphasis on the platyceratid-crinoid interaction. Paleontol. Soc. Pap. 8,
195e210.
Beck, J.T., 1980. The effects of an isopod castrator, Probopyrus pandalicola, on the sex charac-
ters of one of its caridean shrimp hosts, Palaemonetes paludosus. Biol. Bull. 158, 1e15.
Bennett, M.B., Heupel, M.R., Bennett, S.M., Parker, A.R., 1997. Sheina orri (Myodocopa:
Cyprdinidae), an ostracod parasitic on the gills of the epaulette shark, Hemiscyllium
ocellatum (Elasmobranchii: Hemiscyllidae). Int. J. Parasitol. 27, 275e281.
Bernecker, M., Weidlich, O., 2005. Azooxanthellate corals in the Late Maastrichtian-Early
Paleocene of the Danish basin: bryozoan and coral mounds in a boreal shelf setting.
In: Freiwald, R., Roberts, J.M. (Eds.), Cold-water Corals and Ecosystems. Springer,
Berlin, Heidelberg, pp. 3e25.
Bertling, M., Braddy, S., Bromley, R.G., Demathieu, G.D., Genise, J., Mikulas, R.,
Nielsen, J.K., Nielsen, K.S.S., Rindsberg, A., Schlirf, M., Uchman, A., 2006. Names
for trace fossils: a uniform approach to concepts and procedures. Lethaia 39, 265e286.
Beschin, C., Busulini, A., De Angeli, A., Tessier, G., 2007. I Decapodi dell’Eocene inferiore
di Contrada Gecchelina (Vicenza e Italia settentrionale) (Anomura e Brachyura). Mus.
Archeol. Sci. Nat. G. Zannato Montecchio Magg. Vicenza 2007, 5e76.
278 Adiël A. Klompmaker and Geoff A. Boxshall

Beschin, C., Busulini, A., Tessier, G., 2015. Nuova segnalazione di crostacei associati a coralli
nell’Eocene inferiore dei Lessini orientali (Vestenanova e Verona). Lav. Soc. Veneziana
Sci. Nat. 40, 47e109.
Bishop, G.A., 1974. A sexually aberrant crab (Dakoticancer overanus Rathbun, 1917) from the
Upper Cretaceous Pierre Shale of South Dakota. Crustaceana 26, 212e218.
Bishop, G.A., 1983. A second sexually aberrant specimen of Dakoticancer overanus Rathbun,
1917, from the Upper Cretaceous Dakoticancer assemblage, Pierre Shale, South Dakota
(Decapoda, Brachyura). Crustaceana 44, 23e26.
B€
ockeler, W., 2005. Pentastomida (tongue worms). In: Rohde, K. (Ed.), Marine
Parasitology. CABI Publishing, Wallingford, and Csiro Publishing, Collingwood,
pp. 235e240.
Bonnier, J., 1898. Note sur le Pionodesmotes phormosomae, copepode parasite du Phormosoma
uranus. In: Resultats des Campagnes Scientifiques accomplies sur son Yacht par Albert
Ier, Prince souverain de Monaco, vol. 12, pp. 61e66, pl. 10.
Borszcz, T., Wilson, M.A., Binkowski, M., 2014. Jurassic Halloween: discovery of exocysts on
echinoid spines from the Callovian of Israel. In: Palaeontological Association 58th Annual
Meeting. Leeds, England. Programme and Abstracts. doi:10.13140/2.1.3272.5765.
Boucot, A.J., 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier,
Amsterdam, Oxford, New York and Tokyo.
Boucot, A.J., Poinar Jr., G.O., 2010. Fossil Behavior Compendium. CRC Press, Boca
Raton.
Bowman, T.E., 1971. Palaega lamnae, new species (Crustacea: Isopoda) from the Upper
Cretaceous of Texas. J. Paleontol. 45, 540e541.
Boxshall, G.A., 1977. The histopathology of infection by Lepeophtheirus pectoralis (M€ uller,
1776) (Copepoda: Caligidae). J. Fish Biol. 10, 411e415.
Boxshall, G.A., 2005a. Branchiura (fish lice). In: Rohde, K. (Ed.), Marine Parasitology. CABI
Publishing, Wallingford, and Csiro Publishing, Collingwood, pp. 145e147.
Boxshall, G.A., 2005b. Tantulocarida (tantulocarids). In: Rohde, K. (Ed.), Marine
Parasitology. CABI Publishing, Wallingford, and Csiro Publishing, Collingwood,
pp. 147e149.
Boxshall, G.A., 2005c. Copepoda (copepods). In: Rohde, K. (Ed.), Marine Parasitology.
CABI Publishing, Wallingford, and Csiro Publishing, Collingwood, pp. 123e135.
Boxshall, G.A., Halsey, S.H., 2004. An Introduction to Copepod Diversity. The Ray Soci-
ety, London.
Boxshall, G.A., Lincoln, R.J., 1983a. Some new parasitic copepods (Siphonostomatoida:
Nicothoidae) from deep-sea asellote isopods. J. Nat. Hist. 17, 891e900.
Boxshall, G.A., Lincoln, R.J., 1983b. Tantulocarida, a new class of Crustacea ectoparasitic on
other crustaceans. J. Crustacean Biol. 3, 1e16.
Boxshall, G.A., Lincoln, R.J., 1987. The life cycle of the Tantulocarida (Crustacea). Philos.
Trans. R. Soc. Lond. Ser. B 315, 267e303.
Boyko, C.B., Williams, J.D., 2009. Crustacean parasites as phylogenetic indicators in decapod
evolution. In: Martin, J.W., Crandall, K.A., Felder, D.L. (Eds.), Decapod Crustacean
Phylogenetics, Crustacean Issues, 18. CRC Press, Boca Raton, pp. 197e220.
Boyko, C.B., Williams, J.D., 2011. Parasites and other symbionts of squat lobsters. In:
Poore, G.C.B., Ahyong, S.T., Taylor, J. (Eds.), The Biology of Squat Lobsters,
Crustacean Issues 19. CRC Press, Boca Raton, pp. 271e295.
Boyko, C.B., Moss, J., Williams, J.D., Shields, J.D., 2013. A molecular phylogeny of Bopy-
roidea and Cryptoniscoidea (Crustacea: Isopoda). Syst. Biodivers. 11, 495e506.
Brockerhoff, A.M., 2004. Occurrence of the internal parasite Portunion sp. (Isopoda: Entonis-
cidae) and its effect on reproduction in intertidal crabs (Decapoda: Grapsidae) from New
Zealand. J. Parasitol. 90, 1338e1344.
Fossil Crustaceans as Parasites and Hosts 279

Bromley, R.G., 2004. A stratigraphy of marine bioerosion. Geol. Soc. Lond. Spec. Publ. 228,
455e479.
Bromley, R.G., Beuck, L., Ruggiero, E.T., 2008. Endolithic sponge versus terebratulid
brachiopod, Pleistocene, Italy: accidental symbiosis, bioclaustration and deformity. In:
Wisshak, M., Tapanila, L. (Eds.), Current Developments in Bioerosion. Springer, Berlin,
Heidelberg, pp. 361e368.
Buhl-Mortensen, L., Mortensen, P.B., 2004. Symbiosis in deep-water corals. Symbiosis 37,
33e61.
Bursey, C.R., 1978. Histopathology of the parasitization of Munida iris (Decapoda: Galathei-
dae) by Munidion irritans (Isopoda: Bopyridae). Bull. Mar. Sci. 28, 566e570.
Canario, R., Badaro, M.F., Johnsson, R., Neves, E.G., 2015. A new species of Troglocarcinus
(Decapoda: Brachyura: Cryptochiridae) symbiotic with the Brazilian endemic coral Mus-
sismilia (Anthozoa: Scleractinia: Mussidae). Mar. Biol. Res. 11, 76e85.
Carricart-Ganivet, J.P., Carrera-Parra, L.F., Quan-Young, L.I., García-Madrigal, M.S.,
2004. Ecological note on Troglocarcinus corallicola (Brachyura: Cryptochiridae) living in
symbiosis with Manicina areolata (Cnidaria: Scleractinia) in the Mexican Caribbean. Coral
Reefs 23, 215e217.
Castellani, C., Maas, A., Waloszek, D., Haug, J.T., 2011. New pentastomids from the Late
Cambrian of Swedenddeeper insight of the ontogeny of fossil tongue worms. Palaeon-
togr. A 293, 95e145.
Cericola, M.J., Williams, J.D., 2015. Prevalence, reproduction and morphology of the para-
sitic isopod Athelges takanoshimensis Ishii, 1914 (Isopoda: Bopyridae) from Hong Kong
hermit crabs. Mar. Biol. Res. 11, 236e252.
Christoffersen, M.L., De Assis, J.E., 2013. A systematic monograph of the Recent Pentasto-
mida, with a compilation of their hosts. Zool. Meded. 87, 1e206.
Cohen, A.C., 1983. Rearing and postembryonic development of the myodocopid ostracode
Skogsbergia lerneri from the coral reefs of Belize and the Bahamas. J. Crustacean Biol. 3,
235e256.
C
onsole-Gonella, C., Marquillas, R.A., 2014. Bioclaustration trace fossils in epeiric shallow
marine stromatolites: the Cretaceous-Palaeogene Yacoraite Formation, Northwestern
Argentina. Lethaia 47, 107e119.
Conway Morris, S., 1990. Parasitism. In: Briggs, D.E.G., Crowther, P.R. (Eds.), Palaeobio-
logy: A Synthesis. Blackwell Science, Oxford, pp. 376e381.
Cressey, R., Boxshall, G., 1989. Kabatarina pattersoni, a fossil parasitic copepod (Dichelesthii-
dae) from a Lower Cretaceous fish. Micropaleontology 35, 150e167.
Cressey, R., Patterson, C., 1973. Fossil parasitic copepods from a Lower Cretaceous fish. Sci-
ence 180 (4092), 1283e1285.
Cribb, T.H., 2005. Digenea (endoparasitic flukes). In: Rohde, K. (Ed.), Marine Parasitology.
CSIRO Publishing, Collingwood & CABI Publishing, Wallingford, pp. 76e87.
Daeschler, E.B., Shubin, N.H., Jenkins, F.A., 2006. A Devonian tetrapod-like fish and the
evolution of the tetrapod body plan. Nature 440 (7085), 757e763.
Dale, W.E., Anderson, G., 1982. Comparison of morphologies of Probopyrus bithynis,
P. floridensis, and P. pandalicola larvae reared in culture (Isopoda, Epicaridea). J. Crusta-
cean Biol. 2, 392e409.
Darrell, J.G., Taylor, P.D., 1993. Macrosymbiosis in corals: a review of fossil and potentially
fossilizable examples. Cour. Forschunginst. Senckenberg 164, 185e198.
Day, J.H., 1939. A new cirripede parasite e Rhizolepas annelidicola, nov. gen. et sp. Proc.
Linn. Soc. Lond. 151, 64e79.
De Angeli, A., Ceccon, L., 2013. Tetraliidae and Trapeziidae (Crustacea, Decapoda, Brachyura)
from the Early Eocene of Monte Magre (Vicenza, NE Italy). Atti Soc. Italiana Sci. Nat.
Museo Civico Storia Nat. Milano 154, 25e40.
280 Adiël A. Klompmaker and Geoff A. Boxshall

De Angeli, A., Ceccon, L., 2015. Nuovi crostacei brachiuri dell’Eocene di Monte Magre
(Vicenza, Italia settentrionale). Lav. Soc. Veneziana Sci. Nat. 40, 119e138.
De Baets, K., Dentzien-Dias, P.C., Upeniece, I., Verneau, O., Donoghue, P.C.J., 2015.
Constraining the deep origin of parasitic flatworms and host-interactions with fossil
evidence. Adv. Parasitol. 90, 93e135.
De Loriol, P., 1886. Crinoides. Feuilles 1-3. In: Masson, G. (Ed.), Paleontologie française, ou
description des fossiles de la France, Serie 1, Animaux invertebres, Terrain jurassique, vol.
11 (2). Paris, pp. 1e47.
Dojiri, M., 1988. Isomolgus desmotes, new genus, new species (Lichomolgidae), a gallicolous
poecilostome copepod from the scleractinian coral Seriatopora hystrix Dana in
Indonesia, with a review of gall-inhabiting crustaceans of anthozoans. J. Crustacean
Biol. 8, 99e109.
Dollfus, R., 1976. E numération des Cestodes du plankton et des invertébrés marins. Ann.
Parasitol. Hum. Comp. 51, 207e220.
Donovan, S.K., 2015. When is a fossil not a fossil? When it is a trace fossil. Lethaia 48,
145e146.
Dreyer, H., W€agele, J.W., 2001. Parasites of crustaceans (Isopoda: Bopyridae) evolved from
fish parasites: molecular and morphological evidence. Zoology 103, 157e178.
Dumbauld, B.R., Chapman, J.W., Torchin, M.E., Kuris, A.M., 2011. Is the collapse of mud
shrimp (Upogebia pugettensis) populations along the Pacific coast of North America caused
by outbreaks of a previously unknown bopyrid isopod parasite (Orthione griffenis)? Estu-
aries Coasts 34, 336e350.
Edgecombe, G.D., 2010. Arthropod phylogeny: an overview from the perspective of
morphology, molecular data and the fossil record. Arthropod Struct. Dev. 39, 74e87.
Etter, W., 2014. A well-preserved isopod from the Middle Jurassic of southern Germany and
implications for the isopod fossil record. Palaeontology 57, 931e949.
Feldmann, R.M., 1998. Parasitic castration of the crab, Tumidocarcinus giganteus Glaessner,
from the Miocene of New Zealand: coevolution within the Crustacea. J. Paleontol.
72, 493e498.
Feldmann, R.M., Schweitzer, C.E., 2010. The oldest shrimp (Devonian: Famennian) and
remarkable preservation of soft tissue. J. Crustacean Biol. 30, 629e635.
Feldmann, R.M., Tshudy, D.M., Thomson, M.R.A., 1993. Late Cretaceous and Paleocene
decapod crustaceans from James Ross Basin, Antarctic Peninsula. Paleontol. Soc. Mem.
28, 1e41.
Fraaije, R.H.B., 2014. Diverse Late Jurassic anomuran assemblages from the Swabian Alb and
evolutionary history of paguroids based on carapace morphology. Neues Jahrb. Geol.
Pal€aontol. Abh. 273, 121e145.
Fraaije, R.H.B., Klompmaker, A.A., Artal, P., 2012. New species, genera and a family of her-
mit crabs (Crustacea, Anomura, Paguroidea) from a mid-Cretaceous reef of Navarra,
northern Spain. Neues Jahrb. Geol. Pal€aontol. Abh. 263, 85e92.
Fraaije, R.H.B., Van Bakel, B.W.M., Jagt, J.W.M., Skupien, P., 2013a. First record of pagu-
roid anomurans (Crustacea) from the Tithonian-lower Berriasian of Stramberk, Moravia
(Czech Republic). Neues Jahrb. Geol. Pal€aontol. Abh. 269, 251e259.
Fraaije, R.H.B., Artal, P., Van Bakel, B.W.M., Jagt, J.W.M., Klompmaker, A.A., 2013b. An
array of sixth abdominal tergite types of paguroid anomurans (Crustacea) from the mid-
Cretaceous of Navarra, northern Spain. Neth. J. Geosci. 92, 109e117.
Franţescu, O.D., 2014. Fossil mudshrimps (Decapoda: Axiidea) from the Pawpaw formation
(Cretaceous: Albian), northeast Texas, USA. Bull. Mizunami Fossil Mus. 40, 13e22.
Fredensborg, B.L., Poulin, R., 2005. Larval helminths in intermediate hosts: does
competition early in life determine the fitness of adult parasites? Int. J. Parasitol. 35,
1061e1070.
Fossil Crustaceans as Parasites and Hosts 281

Giribet, G., Edgecombe, G.D., 2013. The Arthropoda: a phylogenetic framework. In:
Minelli, A., Boxshall, G.A., Fusco, G. (Eds.), Arthropod Biology and Evolution.
Springer Verlag, Berlin, Heidelberg, pp. 17e40.
Glenner, H., Hebsgaard, M.B., 2006. Phylogeny and evolution of life history strategies of the
parasitic barnacles (Crustacea, Cirripedia, Rhizocephala). Mol. Phylogenet. Evol. 41,
528e538.
Glenner, H., Høeg, J.T., Grygier, M.J., Fujita, Y., 2008. Induced metamorphosis in crusta-
cean y-larvae: towards a solution to a 100-year-old riddle. BMC Biol. 2008 (6), 21.
Glynn, P.W., 1983a. Crustacean symbionts and the defense of corals: co-evolution on the
reef? In: Nitecki, M.H. (Ed.), Coevolution. University of Chicago Press, Chicago and
London, pp. 111e178.
Glynn, P.W., 1983b. Increased survivorship in corals harboring crustacean symbionts. Mar.
Biol. Lett. 4, 105e112.
Goldfuss, A., 1829. Petrefacta Germaniae, Vol. 2. D€ usseldorf, Arnz, pp. 77e164.
Gonzalez, M.T., Acu~ na, E., 2004. Infestation by Pseudione humboldtensis (Bopyridae) in the
squat lobsters Cervimunida johni and Pleuroncodes monodon (Galatheidae) off northern
Chile. J. Crustacean Biol. 24, 618e624.
Gresty, K.A., Boxshall, G.A., Nagasawa, K., 1993. The fine structure and function of the
cephalic appendages of the branchiuran parasite, Argulus japonicus Thiele. Philos. Trans.
R. Soc. Lond. Ser. B 339, 119e135.
Grygier, M.J., 1990. Introcornia (Crustacea: Ascothoracida: Petrarcidae) parasitic in an aherma-
typic coral from Saint Paul Island, Indian Ocean. Vie Milieu 40 (4), 313e318.
Grygier, M.J., 1996. Ascothoracida. In: Forest, J. (Ed.), Traité de Zoologie 7(2), Crustacés:
Géneralités (suite) et Systématique (1re partie). Masson, Paris, pp. 433e452.
Grygier, M.J., Høeg, J.T., 2005. Ascothoracida (ascothoracids). In: Rohde, K. (Ed.), Marine
Parasitology. CABI Publishing, Wallingford, and Csiro Publishing, Collingwood,
pp. 149e154.
Gueriau, P., Charbonnier, S., Clément, G., 2014. First decapod crustaceans in a Late Devo-
nian continental ecosystem. Palaeontology 57, 1203e1213.
Guinot, D., Tavares, M., Castro, P., 2013. Significance of the sexual openings and sup-
plementary structures on the phylogeny of brachyuran crabs (Crustacea, Decapoda,
Brachyura), with new nomina for higher-ranked podotreme taxa. Zootaxa 3665,
1e414.
Harding, J.P., 1966. Myodocopan ostracods from the gills and nostrils of fishes. In: Barnes, H.
(Ed.), Some Contemporary Studies in Marine Science. Allen & Unwin, London,
pp. 369e374.
Hayashi, R., Chan, B.K.K., Simon-Blecher, N., Watanabe, H., Guy-Haim, T.,
Yonezawa, T., Levy, Y., Shuto, T., Achituv, Y., 2013. Phylogenetic position and evolu-
tionary history of turtle and whale barnacles (Cirripedia: Balanomorpha: Coronuloidea).
Mol. Phylogenet. Evol. 67, 9e14.
Herlyn, H., Piskurek, O., Schmidtz, J., Ehlers, U., Zischler, H., 2003. The phylogeny of the
Syndermata (Rotifera: Monogonota, Bdelloidea, Seisonidea; Acanthocephala
Palaeacanthocephala, Eoacanthocephala, Archiacanthocephala). Mol. Phylogenet.
Evol. 26, 155e164.
Hernaez, P., Martínez-Guerrero, B., Anker, A., Wehrtmann, I.S., 2010. Fecundity and
effects of bopyrid infestation on egg production in the Caribbean sponge-dwelling
snapping shrimp Synalpheus yano (Decapoda: Alpheidae). J. Mar. Biol. Assoc. U.K. 90,
691e698.
Hessler, R.R., 1969. Peracarida. In: Moore, R.C. (Ed.), Treatise on Invertebrate
Paleontology, Arthropoda, vol. 4 (1). University of Kansas and Geological Society of
America, Lawrence, pp. R360eR393.
282 Adiël A. Klompmaker and Geoff A. Boxshall

Hildner, R., 2014. Hochbautechnik im Muschelkalk-Meer e Encrinus-Stiele unter der Lupe.


Fossilien 2014 (3), 26e31.
Hiro, F., 1937. Studies on the animals inhabiting reef corals. I. Hapalocarcinus and Cryptochirus.
Palao Trop. Biol. Stn. Stud. 1, 137e154.
Hobbs Jr., H.H., Peters, D.J., 1977. The entocytherid ostracods of North Carolina. Smithson.
Contrib. Zool. 247, 1e73.
Høeg, J.T., 1995. The biology and life cycle of the Rhizocephala (Cirripedia). J. Mar. Biol.
Assoc. U.K. 75, 517e550.
Høeg, J.T., Glenner, H., Shields, J.D., 2005. Cirripedia Thoracica and Rhizocephala
(barnacles). In: Rohde, K. (Ed.), Marine Parasitology. CABI Publishing, Wallingford,
and Csiro Publishing, Collingwood, pp. 154e165.
Hosie, A.M., 2008. Four new species and a new record of Cryptoniscoidea (Crustacea:
Isopoda: Hemioniscidae and Crinoniscidae) parasitizing stalked barnacles from New
Zealand. Zootaxa 1795, 1e28.
Housa, V., 1963. Parasites of Tithonian decapod crustaceans (Stramberk, Moravia).
 redniho Ustavu
Ust  Geol. Sb. UUG 28, 101e114.
Housa, V., Vasícek, Z., 2004. Ammonoidea of the Lower Cretaceous deposits (Late Berriasian,
Valanginian, Early Hauterivian) from Stramberk, Czech Republic. GeoLines 18, 7e57.
Huys, R., Boxshall, G.A., Lincoln, R.J., 1993. The Tantulocaridan life cycle: the circle
closed? J. Crustacean Biol. 13, 432e442.
Hyzný, M., Bruce, N.L., Schl€ ogl, J., 2013. An appraisal of the fossil record for the Cirolanidae
(Malacostraca: Peracarida: Isopoda: Cymothoida), with a description of a new cirolanid
isopod crustacean from the Early Miocene of the Vienna Basin (Western Carpathians).
Palaeontology 56, 615e630.
Hyzný, M., Gasparic, R., Robins, C.M., Schl€ ogl, J., 2014. Miocene squat lobsters (Deca-
poda, Anomura, Galatheoidea) of the Central Paratethys e a review, with description
of a new species of Munidopsis. In: Fraaije, R.H.B., Hyzný, M., Jagt, J.W.M.,
Krobicki, M., Van Bakel, B.W.M. (Eds.), Proceedings of the 5th Symposium on Meso-
zoic and Cenozoic Decapod Crustaceans, Krakow, Poland, 2013: A Tribute to Pal
Mihaly M€ uller. Scripta Geologica 147, 241e267.
Jakobsen, S.L., Collins, J.S.H., 1997. New middle Danian species of anomurans and
brachyuran crabs from Fakse, Denmark. Bull. Geol. Soc. Den. 44, 89e100.
Johnsson, R., Neves, E., Franco, G.M.O., Da Silveira, F.L., 2006. The association of two gall
crabs (Brachyura: Cryptochiridae) with the reef-building coral Siderastrea stellata Verrill,
1868. Hydrobiologia 559, 379e384.
Jones, A., 2013. Population Dynamics of Dakoticancer overanus from the Pierre Shale, South
Dakota. MS Thesis, Kent State University, Kent.
Jones, W.T., Feldmann, R.M., Schweitzer, C.E., Schram, F.R., Behr, R.A., Hand, K.L.,
2014. The first Paleozoic stenopodidean from the Huntley Mountain formation (Devo-
nian-Carboniferous), north-central Pennsylvania. J. Paleontol. 88, 1251e1256.
Kier, P.M., 1981. A bored Cretaceous echinoid. J. Paleontol. 55, 656e659.
Kiessling, W., Fl€ ugel, E., Golonka, J., 1999. Paleoreef maps: evaluation of a comprehensive
database on Phanerozoic reefs. AAPG Bull. 83, 1552e1587.
Kim, I.-H., Yamashiro, H., 2007. Two species of poecilostomatoid copepods inhabiting galls
on scleractinian corals in Okinawa, Japan. J. Crustacean Biol. 27, 319e326.
Kinne, O., 1980. Diseases of marine animals: general aspects. In: Kinne, O. (Ed.), Diseases of
Marine Animals, Volume I: General Aspects, Protozoa to Gastropoda. Biologische
Anstalt Helgoland, Hamburg, pp. 13e73.
Klompmaker, A.A., 2013. Extreme diversity of decapod crustaceans from the mid-Creta-
ceous (late Albian) of Spain: implications for Cretaceous decapod paleoecology. Creta-
ceous Res. 41, 150e185.
Fossil Crustaceans as Parasites and Hosts 283

Klompmaker, A.A., Artal, P., Van Bakel, B.W.M., Fraaije, R.H.B., Jagt, J.W.M., 2011. Etyid
crabs (Crustacea, Decapoda) from mid-Cretaceous reefal strata of Navarra, northern
Spain. Palaeontology 54, 1199e1212.
Klompmaker, A.A., Feldmann, R.M., Robins, C.M., Schweitzer, C.E., 2012. Peak diversity
of Cretaceous galatheoids (Crustacea, Decapoda) from northern Spain. Cretaceous Res.
36, 125e145.
Klompmaker, A.A., Schweitzer, C.E., Feldmann, R.M., Kowalewski, M., 2013a. The influ-
ence of reefs on the rise of Mesozoic marine crustaceans. Geology 41, 1179e1182.
Klompmaker, A.A., Ortiz, J.D., Wells, N.A., 2013b. How to explain a decapod crustacean
diversity hotspot in a mid-Cretaceous coral reef. Palaeogeogr. Palaeoclimatol. Palaeoe-
col. 374, 256e273.
Klompmaker, A.A., Artal, P., Van Bakel, B.W.M., Fraaije, R.H.B., Jagt, J.W.M., 2014. Pa-
rasites in the fossil record: a Cretaceous fauna with isopod-infested decapod crustaceans,
infestation patterns through time, and a new ichnotaxon. PLoS One 9 (3), e92551.
http://dx.doi.org/10.1371/journal.pone.0092551.
Klompmaker, A.A., Hyzný, M., Portell, R.W., Kowalewski, M., 2015. Growth, inter- and
intraspecific variation, palaeobiogeography, taphonomy and systematics of the Cenozoic
ghost shrimp Glypturus. J. Syst. Palaeontol., online ahead of print 1e28. http://
dx.doi.org/10.1080/14772019.2015.1009505.
Knaust, D., 2012. Trace-fossil systematics. In: Knaust, D., Bromley, R.G. (Eds.), Trace Fossils
as Indicators of Sedimentary Environments, Developments in Sedimentology, 64.
Elsevier, Amsterdam, pp. 79e101.
Knudsen, J.W., 1967. Trapezia and Tetralia (Decapoda, Brachyura, Xanthidae) as obligate
ectoparasites of pocilloporid and acroporid corals. Pac. Sci. 21, 51e57.
Kornecki, K.M., 2014. Cretaceous Confluence in the Coon Creek Formation of Mississippi
and Tennessee, USA: Taphonomy and Systematic Paleontology of a Decapod Konsen-
trat-Lagerst€atte. MS Thesis, Kent State University, Kent.
K€orner, H., 1979. Tiersymbiosen und €ahnliche Formen der Vergesellschaftung. Von Dieter
Matthes. 241 Seiten, 87 Abbildungen. Gustav Fischer Verlag, Stuttgart- New York
1978. Biol. Unserer Zeit 9 (5), 158.
Krawczy nski, C., Wilson, M., 2011. The first Jurassic thecideide brachiopods from the Mid-
dle East: a new species of Moorellina from the Upper Callovian of Hamakhtesh Hagadol,
southern Israel. Acta Geol. Pol. 61, 71e77.
Kropp, R.K., 1986. Feeding biology and mouthpart morphology of three species of corals
gall crabs (Decapoda: Cryptochiridae). J. Crustacean Biol. 6, 377e384.
Lambert, J., 1927. Parasite inconnu et nouveau du callovien. Extraits P.V. Séances Soc. Linn.
Bordeaux Actes la Soc. Linn. Bordeaux 79, 59.
Lauridsen, B.W., Bjerager, M., 2014. Danian cold-water corals from the Baunekule facies,
Faxe Formation, Denmark: a rare taphonomic window of a coral mound flank
habitat. Lethaia 47, 437e455.
Lauridsen, B.W., Bjerager, M., Surlyk, F., 2012. The middle Danian Faxe Formationdnew
lithostratigraphic unit and a rare taphonomic window into the Danian of Denmark. Bull.
Geol. Soc. Den. 60, 47e60.
Lester, R.J.G., 2005. Isopoda (isopods). In: Rohde, K. (Ed.), Marine Parasitology. CABI
Publishing, Wallingford, and Csiro Publishing, Collingwood, pp. 138e144.
Leung, T.L., 2014. Evolution: how a barnacle came to parasitise a shark. Curr. Biol. 24,
R564eR566.
Levine, N.D., 1988. The Protozoan Phylum Apicomplexa, vol. 1. CRC Press Inc., Boca
Raton, 203 pp.
Lightner, D.V., 1993. Diseases of cultured penaeid shrimp. In: McVey, J.P. (Ed.), CRC Hand-
book of Mariculture. Crustacean Aquaculture. CRC Press, Boca Raton, pp. 393e486.
284 Adiël A. Klompmaker and Geoff A. Boxshall

Lins, L.S.F., Ho, S.Y.W., Wilson, G.D.F., Lo, N., 2012. Evidence for Permo-Triassic colo-
nization of the deep sea by isopods. Biol. Lett. 8, 979e982.
Littlewood, D.T.J., 2006. The evolution of parasitism in flatworms. In: Maule, A.G.,
Marks, N.J. (Eds.), Parasitic Flatworms: Molecular Biology, Biochemistry, Immunology
and Physiology. CABI Pub., Wallingford, Cambridge, pp. 1e36.
Littlewood, D.T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
Lowry, J.K., Myers, A.A., 2013. A phylogeny and classification of the Senticaudata subord.
nov. (Crustacea: Amphipoda). Zootaxa 3610, 1e80.
Luksevics, E., Lebedev, O., Mark-Kurik, E., Karataj ut_e-Talimaa, V., 2009. The earliest evi-
dence of hosteparasite interactions in vertebrates. Acta Zool. 90 (suppl. 1), 335e343.
L€
utzen, J., 2005. Amphipoda (amphipods). In: Rohde, K. (Ed.), Marine Parasitology. CABI
Publishing, Wallingford, and Csiro Publishing, Collingwood, pp. 165e169.
Maas, A., Braun, A., Dong, X.P., Donoghue, P.C.J., M€ uller, K.J., Olempska, E.,
Repetski, J.E., Siveter, D.J., Stein, M., Waloszek, D., 2006. The ‘Orsten’dmore than
a Cambrian Konservat-Lagerst€atte yielding exceptional preservation. Palaeoworld 15,
266e282.
Madsen, F.J., Wolff, T., 1965. Evidence of the occurrence of Ascothoracica (parasitic cirri-
peds) in Upper Cretaceous. Medd. Dan. Geol. Foren. 15, 556e558.
Malay, M.C.M.D., Michonneau, F., 2014. Phylogenetics and morphological evolution of
coral-dwelling barnacles (Balanomorpha: Pyrgomatidae). Biol. J. Linn. Soc. 113, 162e179.
Markham, J.C., 1986. Evolution and zoogeography of the Isopoda Bopyridae, parasites of
Crustacea Decapoda. In: Gore, G.H., Heck, K.L. (Eds.), Crustacean Issues 4. Crustacean
Biogeography. A.A. Balkema, Rotterdam, pp. 143e164.
Martill, D.M., 2007. The age of the Cretaceous Santana Formation fossil Konservat Lager-
st€atte of north-east Brazil: a historical review and an appraisal of the bio-
chronostratigraphic utility of its palaeobiota. Cretaceous Res. 28, 895e920.
Martin, J.W., Davis, G., 2001. An updated classification of the recent Crustacea. Nat. Hist.
Mus. Los Angeles Sci. Ser. 19, 1e124.
McClelland, G., 2002. The trouble with sealworms (Pseudoterranea decipiens species complex,
Nematoda): a review. Parasitology 125, 183e203.
McDermott, J.J., 1991. Incidence and host-parasite relationship of Leidya bimini (Crustacea,
Isopoda, Bopyridae) in the brachyuran crab Pachygrapsus transversus from Bermuda.
Ophelia 33, 71e95.
McKeon, C.S., Moore, J.M., 2014. Species and size diversity in protective services offered by
coral guard-crabs. PeerJ 2, e574. http://dx.doi.org/10.7717/peerj.574.
McKinney, F.K., 2009. Bryozoan-hydroid symbiosis and a new ichnogenus, Caupokeras. Ich-
nos 16, 193e201.
Mehl, J., Mehl, D., Hackel, W., 1991. Parasitare Zystenbildungen an jurassischen Cidariden
und das Porospongia-Problem. Berl. Geowiss. Abh. A134, 227e261.
Mercier, J., 1936. Zoothylacies d’E  chinide fossile provoquées par un Crustacé: Castexia dou-
villei nov. gen., nov. sp. Bull. Soc. Géol. France 5 (6), 149e154.
Miyashita, Y., 1941. Observations on an entoniscid parasite of Eriocheir japonicus de Haan,
Entionella fluviatilis n. g., n. sp. J. Zool. 9, 251e267.
Mohammed, T.A.A., Yassien, M.H., 2013. Assemblages of two gall crabs within coral species
northern Red Sea, Egypt. Asian J. Sci. Res. 6, 98e106.
Mohrbeck, I., Arbizu, P., Glatzel, T., 2010. Tantulocarida (Crustacea) from the Southern
Ocean deep sea, and the description of three new species of Tantulacus Huys, Andersen
& Kristensen, 1992. Syst. Parasitol. 77, 131e151.
Moyse, J., 1983. Isidascus bassindalei gen. nov., sp. nov. (Ascothoracida: Crustacea) from
north-east Atlantic with a note on the origin of barnacles. J. Mar. Biol. Assoc. U.K.
63, 161e180.
Fossil Crustaceans as Parasites and Hosts 285

Myers, A.A., Lowry, J.K., 2003. A phylogeny and a new classification of the Corophidea
Leach, 1814 (Amphipoda). J. Crustacean Biol. 23, 443e485.
Near, T.J., 2002. Acanthocephalan phylogeny and the evolution of parasitism. Integr. Comp.
Biol. 42, 668e677.
Neumann, C., Hostettler, B., 2004. Evolutions€ okologie parasitischer Interaktionen am Beispiel
mesozoischer Ascothoracica und Echinodermata. In: Reich, M., Hagdorn, H., Reitner, J.
(Eds.), 3. Arbeitstreffen deutschsprachiger Echinodermenforscher in Ingelfingen, 29. bis 31.
oktober 2004, Arbeiten und Kurzfassungen. Universit€atsdrucke, G€ ottingen, pp. 39e40.
Nicolleau, P., Vadet, A., 1995. Les oursins des marnes a spongiaires de l’Oxfordien du
Poitou. In: Branger, P., Nicolleau, P., Vadet, A. (Eds.), Les ammonites et les oursins
de l’Oxfordien du Poitou (facies a spongiaires de l’Oxfordien moyen et supérieur).
Musées de la ville de Niort, Aiffres, pp. 55e81.
Nogueira, M.M., Menezes, N.M., Johnsson, R., Neves, E., 2014. The adverse effects of
cryptochirid crabs (Decapoda: Brachyura) on Siderastrea stellata Verril, 1868 (Anthozoa:
Scleractinia): causes and consequences of cavity establishment. Cah. Biol. Mar. 55,
155e162.
Oakley, T.H., Wolfe, J.M., Lindgren, A.R., Zaharoff, A.K., 2013. Phylotranscriptomics to
bring the understudied into the fold: monophyletic Ostracoda, fossil placement, and pan-
crustacean phylogeny. Mol. Biol. Evol. 30, 215e233.
O’Brien, J., Van Wyk, P., 1985. Effects of crustacean parasitic castrators (epicaridean iso-
pods and rhizocephalan barnacles) on growth of crustacean hosts. In: Wenner, A.M.
(Ed.), Crustacean Issues 3. Factors in Adult Growth. A. A. Balkema, Rotterdam,
pp. 191e218.
Okawachi, H., Ohtsuka, S., Norshida, B.I., Venmathi Maran, B.A., Ogawa, K., 2012. Sea-
sonal occurrence and microhabitat of the hyperparasitic monogenean Udonella fugu on
the caligid copepod Pseudocaligus fugu infecting the grass puffer Takifugu niphobles in
the Seto Inland Sea, Japan. Ocean Sci. J. 47, 181e187.
Palmer, T.J., Wilson, M.A., 1988. Parasitism of Ordovician bryozoans and the origin of
pseudoborings. Palaeontology 31, 939e949.
Pasini, G., Garassino, A., 2012a. Palaega pisana n. sp. (Crustacea, Isopoda, Cirolanidae) from
the Pliocene of Orciano Pisano, Pisa (Toscana, Central Italy). Atti Soc. Italiana Sci. Nat.
Museo Civico Storia Nat. Milano 153, 3e11.
Pasini, G., Garassino, A., 2012b. First record of cirolanids (Crustacea, Isopoda, Cirolanidae)
from the middle Pliocene of Parma and Reggio Emilia Province (Emilia Romagna, N
Italy). Atti Soc. Italiana Sci. Nat. Museo Civico Storia Nat. Milano 153, 13e20.
Patton, W.K., 1967. Commensal Crustacea. In: Proceedings Symposium Crustacea, Marine
Biological Association of India, Part 3, pp. 1228e1243.
Pérez-Losada, M., Høeg, J.T., Crandall, K.A., 2004. Unraveling the evolutionary radiation
of the thoracican barnacles using molecular and morphological evidence: a comparison of
several divergence time estimation approaches. Syst. Biol. 53, 244e264.
Petit, G., 2010. Skin nodules in fossil fishes from Monte Bolca (Eocene, Northern Italy).
Geodiversitas 32, 157e163.
Petit, G., Khalloufi, B., 2012. Paleopathology of a fossil fish from the Solnhofen Lagerst€atte
(Upper Jurassic, southern Germany). Int. J. Paleopathol. 2, 42e44.
Petric, M., Ferri, J., Mladineo, I., 2010. Growth and reproduction of Munida rutllanti (Deca-
poda: Anomura: Galatheidae) and impact of parasitism by Pleurocrypta sp. (Isopoda:
Bopyridae) in the Adriatic Sea. J. Mar. Biol. Assoc. U.K. 90, 1395e1404.
Polz, H., 2004. Asselansammlung auf einer Wasserwanze aus den Solnhofener Plattenkalken.
Archaeopteryx 22, 51e60.
Polz, H., Schweigert, G., Maisch, M.W., 2006. Two new species of Palaega (Isopoda: Cym-
othoida: Cirolanidae) from the Upper Jurassic of the Swabian Alb, South Germany.
Stuttg. Beitr. Naturkd. Ser. B 362, 1e17.
286 Adiël A. Klompmaker and Geoff A. Boxshall

Poore, G.C.B., 2012. The nomenclature of the Recent Pentastomida (Crustacea), with a list
of species and available names. Syst. Parasitol. 82, 211e240.
Portell, R.W., Klompmaker, A.A., 2014. First evidence of coral-inhabiting gall crabs (Cryp-
tochiridae) from the fossil record. North Am. Paleontol. Conv. Abstr. Paleontol. Soc.
Spec. Publ. 13, 91.
Potts, F.A., 1915. Hapalocarcinus, the gall-forming crab, with some notes on the related
Cryptochirus. Papers from the Department of Marine Biology of the Carnegie Institution
of Washington 8, 33e69.
Radwa nski, A., 1972. Isopod-infected prosoponids from the Upper Jurassic of Poland. Acta
Geol. Pol. 22, 499e506.
Radwa nska, U., Radwa nska, A., 2005. Myzostomid and copepod infestation of Jurassic echi-
noderms: a general approach, some new occurrences, and/or re-interpretation of previ-
ous reports. Acta Geol. Pol. 55, 109e130.
Radwa nska, U., Poirot, E., 2010. Copepod-infested Bathonian (Middle Jurassic) echinoids
from northern France. Acta Geol. Pol. 60, 549e555.
Rasmussen, E., 1973. Systematics and ecology of the Isefjord marine fauna (Denmark).
Ophelia 11, 1e495.
Rathbun, M.J., 1935. Fossil Crustacea of the Atlantic and Gulf Coastal Plain. Geol. Soc. Am.
Spec. Pap. 2 (i-viii), 1e160.
Rayner, G.W., 1935. The Falkland species of the crustacean genus Munida. Discovery Rep.
Cambridge 10, 209e245.
Rees, D.J., Noever, C., Høeg, J.T., Ommundsen, A., Glenner, H., 2014. On the origin of a
novel parasitic-feeding mode within suspension-feeding barnacles. Curr. Biol. 24,
1429e1434.
Regier, J.C., Shultz, J.W., Zwick, A., Hussey, A., Ball, B., Wetzer, R., Martin, J.W.,
Cunningham, C.W., 2010. Arthropod relationships revealed by phylogenomic analysis
of nuclear protein-coding sequences. Nature 463, 1079e1083.
Reinhard, E.G., 1956. Parasitic castration of Crustacea. Parasitology 5, 79e107.
Richter, A.E., 1991. Seeigel mit Myzostomiden-Parasiten. Fossilien 8 (3), 134.
Robins, C.M., Feldmann, R.M., Schweitzer, C.E., 2012. The oldest Munididae (Decapoda:
Anomura: Galatheoidea) from Ernstbrunn, Austria (Tithonian). Ann. Naturhist. Mus.
Wien Ser. A 114, 289e300.
Robins, C.M., Feldmann, R.M., Schweitzer, C.E., 2013. Nine new genera and 24 new spe-
cies of the Munidopsidae (Decapoda: Anomura: Galatheoidea) from the Jurassic Ernst-
brunn Limestone of Austria, and notes on fossil munidopsid classification. Ann.
Naturhist. Mus. Wien Ser. A 115, 167e251.
Roccatagliata, D., Lovrich, G.A., 1999. Infestation of the false king crab Paralomis granulosa
(Decapoda: Lithodidae) by Pseudione tuberculata (Isopoda: Bopyridae) in the Beagle Chan-
nel, Argentina. J. Crustacean Biol. 19, 720e729.
Rodriguez, J., Gutschick, R.C., 1977. Barnacle borings in live and dead hosts from the Loui-
siana Limestone (Famennian) of Missouri. J. Paleontol. 51, 718e724.
Rohde, K., 2005. Definitions, and adaptations to a parasitic way of life. In: Rohde, K. (Ed.),
Marine Parasitology. CABI Publishing, Wallingford, and Csiro Publishing, Colling-
wood, pp. 1e6.
Ross, A., Newman, W.A., 1969. A coral-eating barnacle. Pac. Sci. 23, 252e256.
Ross, A., Newman, W.A., 1995. A coral-eating barnacle, revisited (Cirripedia,
Pyrgomatidae). Contrib. Zool. 65, 129e175.
Ross, A., Newman, W.A., 2002. Coral barnacles: Cenozoic decline and extinction in
the Atlantic/East Pacific versus diversification in the Indo-West Pacific. In: Proceedings
of the 9th International Coral Reef Symposium, Bali, Indonesia, vol. 1, pp. 135e139.
Sanders, K.L., Lee, M.S., 2010. Arthropod molecular divergence times and the Cambrian
origin of pentastomids. Syst. Biodiversity 8, 63e74.
Fossil Crustaceans as Parasites and Hosts 287

Santos, A., Mayoral, E., Baarli, B.G., Silva, D., Carlos, M., Cach~ao, M., Johnson, M.E., 2012.
Symbiotic association of a pyrgomatid barnacle with a coral from a volcanic middle
Miocene shoreline (Porto Santo, Madeira Archipelago, Portugal). Palaeontology 55,
173e182.
Schneider, S., Harzhauser, M., Kroh, A., Lukeneder, A., 2013. Ernstbrunn Limestone and
Klentnice beds (Kimmeridgian Berriasian; Waschberg-danice Unit; NE Austria and SE
Czech Republic): state of the art and bibliography. Bull. Geosci. 88, 105e130.
Schweitzer, C.E., 2005. The Trapeziidae and Domeciidae (Decapoda: Brachyura: Xanthoi-
dea) in the fossil record and a new Eocene genus from Baja California Sur, Mexico.
J. Crustacean Biol. 25, 625e636.
Schweitzer, C.E., Feldmann, R.M., 2008. Revision of the genus Laeviprosopon Glaessner,
1933 (Decapoda: Brachyura: Homolodromioidea: Prosopidae) including two new
species. Neues Jahrb. Geol. Pal€aontol. Abh. 250, 273e285.
Seilacher, A., 1968. Swimming habits of belemnites, recorded by boring barnacles. Palaeo-
geogr. Palaeoclimatol. Palaeoecol. 4, 279e285.
Seilacher, A., 1969. Paleoecology of boring barnacles. Am. Zool. 9, 705e719.
Selden, P.A., Huys, R., Stephenson, M.H., Heward, A.P., Taylor, P.N., 2010. Crustaceans
from bitumen clast in Carboniferous glacial diamictite extend fossil record of copepods.
Nat. Commun. 1, 50.
Shields, J.D., Kuris, A.M., 1985. Ectopic infections of Portunion conformis (Isopoda: Entonis-
cidae) in Hemigrapsus spp. J. Invertebr. Pathol. 45, 122e124.
Shields, J.D., Ward, L.A., 1998. Tiarinion texopallium, new species, an entoniscid isopod
infesting majid crabs (Tiarinia spp.) from the Great Barrier Reef, Australia. J. Crustacean
Biol. 18, 590e596.
Simms, M.J., 1989. British Lower Jurassic crinoids. Monogr. Palaeontogr. Soc. Lond. 142
(581), 1e103, pls. 1-15.
Simon-Blecher, N., Achituv, Y., 1997. Relationship between the coral pit crab Cryptochirus
coralliodytes Heller and its host coral. J. Exp. Mar. Biol. Ecol. 215, 93e102.
Simon-Blecher, N., Chemedanov, A., Eden, N., Achituv, Y., 1999. Pit structure and trophic
relationship of the coral pit crab Cryptochirus coralliodytes. Mar. Biol. 134, 711e717.
Siveter, D.J., Briggs, D.E.G., Siveter, D.J., Sutton, M.D., 2015. A 425-million-year-old Silu-
rian pentastomid parasitic on ostracods. Curr. Biol. 25, 1632e1637.
Smit, N.J., Bruce, N.L., Hadfield, K.A., 2014. Global diversity of fish parasitic isopod crus-
taceans of the family Cymothoidae. Int. J. Parasitol. Parasites Wildl. 3, 188e197.
Smith, A.E., Chapman, J.W., Dumbauld, B.R., 2008. Population structure and energetics of
the bopyrid isopod parasite Orthione griffenis in mud shrimp Upogebia pugettensis. J. Crus-
tacean Biol. 28, 228e233.
Solovjev, A.N., 1961. The parasite Canceripustula nocens in a Late Jurassic echinoid. Paleontol.
Zh. 4, 115e119.
Solovjev, A.N., Markov, A.V., 2013. The role of echinoids in shaping environments. Pale-
ontol. J. 47, 480e484.
Stimson, J., 1990. Stimulation of fat-body production in the polyps of the coral Pocillopora
damicornis by the presences of mutualistic crabs of the genus Trapezia. Mar. Biol. 106,
211e218.
Stock, J.H., 1968. The Calvocheridae, a family of copepods inducing galls in sea-urchin
spines. Bijdr. Dierkd. 38, 85e90.
Stock, J.H., 1981. Associations of Hydrocorallia Stylasterina with gall-inhabiting Copepoda
Siphonostomatoidea from the south-west Pacific. Part II. On six species belonging to
four new genera of the copepod family Asterocheridae. Bijdr. Dierkd. 51, 287e312.
Takahashi, T., Matsuura, S., 1994. Laboratory studies on molting and growth of the shore
crab, Hemigrapsus sanguineus de Haan, parasitized by a rhizocephalan barnacle. Biol.
Bull. 186, 300e308.
288 Adiël A. Klompmaker and Geoff A. Boxshall

Tapanila, L., 2008a. Direct evidence of ancient symbiosis using trace fossils. In: From Evo-
lution to Geobiology: Research Questions Driving Paleontology at the Start of a
New Century, Paleontological Society Short Course, pp. 271e287.
Tapanila, L., 2008b. The endolithic guild: an ecological framework for residential cavities in
hard substrates. In: Wisshak, M., Tapanila, L. (Eds.), Current Developments in
Bioerosion. Springer, Berlin/Heidelberg, pp. 3e20.
Tapanila, L., Ekdale, A.A., 2007. Early history of symbiosis in living substrates: trace-fossil
evidence from the marine record. In: Miller III, W. (Ed.), Trace Fossils: Concepts, Pro-
blems, Prospects. Elsevier, Amsterdam, pp. 345e355.
Taraschewski, H., 2000. Host-parasite interactions in Acanthocephala: a morphological
approach. Adv. Parasitol. 46, 1e179.
Taylor, P.D., Wilson, M.A., 2003. Palaeoecology and evolution of marine hard substrate
communities. Earth Sci. Rev. 62, 1e103.
Tomlinson, J.T., 1969. The burrowing barnacles: (Cirripedia: Order Acrothoracica). U.S.
Natl. Mus. Bull. 296, 1e162.
Tsang, L.M., Chu, K.H., Nozawa, Y., Chan, B.K.K., 2014. Morphological and host speci-
ficity evolution in coral symbiont barnacles (Balanomorpha: Pyrgomatidae) inferred from
a multi-locus phylogeny. Mol. Phylogenet. Evol. 77, 11e22.
Tucker, B.W., 1930. On the effects of an Epicaridan Parasite, Gyge branchialis, on Upogebia
littoralis. Q. J. Microsc. Sci. 2 (293), 1e118.
Tucker, A.B., Feldmann, R.M., Powell, C.L., 1994. Speocarcinus berglundi n. sp. (Decapoda:
Brachyura), a new crab from the Imperial Formation (late Miocene-late Pliocene) of
southern California. J. Paleontol. 68, 800e807.
Upeniece, I., 1999. Fossil record of parasitic helminths in fishes. In: Abstracts of the “5th
International Symposium on Fish Parasites”, 154.
Upeniece, I., 2001. The unique fossil assemblage from the Lode Quarry (Upper Devonian,
Latvia). Mitt. Mus. Naturkd. Berlin e Geowiss. Reihe 4, 101e119.
Upeniece, I., 2011. Palaeoecology and Juvenile Individuals of the Devonian Placoderm and
Acanthodian Fishes from Lode Site, Latvia. Unpublished Doctoral Thesis, University of
Latvia.
Utinomi, H., 1944. Studies on the animals inhabiting reef corals. III. A revision of the family
Hapalocarcinidae (Brachyura), with some remarks on their morphological peculiarities.
Palao Trop. Biol. Stn. Stud. 2, 687e731, pls. 3e5.
Van der Meij, S.E., 2014. Host species, range extensions, and an observation of the mating
system of Atlantic shallow-water gall crabs (Decapoda: Cryptochiridae). Bull. Mar. Sci.
90, 1001e1010.
Van der Meij, S.E., 2015. Host relations and DNA reveal a cryptic gall crab species (Crus-
tacea: Decapoda: Cryptochiridae) associated with mushroom corals (Scleractinia:
Fungiidae). Contrib. Zool. 84, 39e57.
Van der Meij, S.E., Schubart, C.D., 2014. Monophyly and phylogenetic origin of the gall
crab family Cryptochiridae (Decapoda : Brachyura). Invertebr. Syst. 28, 491e500.
Vannier, J., Abe, K., 1993. Functional morphology and behavior of Vargula hilgendorfii
(Ostracoda: Myodocopida) from Japan, and discussion of its crustacean ectoparasites: pre-
liminary results from video recordings. J. Crustacean Biol. 13, 51e76.
Van Wyk, P.M., 1982. Inhibition of the growth and reproduction of the porcellanid
crab Pachycheles rudis by the bopyrid isopod, Aporobopyrus muguensis. Parasitology 85,
459e473.
Verrill, A.E., 1867. Remarkable instances of crustacean parasitism. Am. J. Sci. 44 (2), 126.
Vinn, O., Wilson, M.A., M~ otus, M.A., Toom, U., 2014. The earliest bryozoan parasite:
Middle Ordovician (Darriwilian) of Osmussaar Island, Estonia. Palaeogeogr. Palaeocli-
matol. Palaeoecol. 414, 129e132.
Fossil Crustaceans as Parasites and Hosts 289

Voigt, E., 1959. Endosacculus moltkiae n. g. n. sp., ein vermutlicher fossiler Ascothoracide (Ento-
mostr.) als Cystenbildner bei der Oktokoralle Moltkia minuta. Pal€aontol. Z. 33, 211e223.
Voigt, E., 1967. Ein vermutlicher Ascothoracide (Endosacculus (?) najdini n. sp.) als Bewohner
einer kretazischen Isis aus der UdSSR. Pal€aontol. Z. 41, 86e90.
Walossek, D., M€ uller, K.J., 1994. Pentastomid parasites from the Lower Palaeozoic of
Sweden. Trans. R. Soc. Edinburgh: Earth Sci. 85, 1e37.
Walossek, D., Repetski, J.E., M€ uller, K.J., 1994. An exceptionally preserved parasitic
arthropod, Heymonsicambria taylori n. sp. (Arthropoda incertae sedis: Pentastomida),
from Cambrian-Ordovician boundary beds of Newfoundland, Canada. Can. J. Earth
Sci. 31, 1664e1671.
Waloszek, D., Repetski, J.E., Maas, A., 2006. A new Late Cambrian pentastomid and a review of
the relationships of this parasitic group. Trans. R. Soc. Edinburgh: Earth Sci. 96, 163e176.
Weber, M., Wey-Fabrizius, A.R., Podsiadlowski, L., Witek, A., Schill, R.O., Sugar, L.,
Herlyn, H., Hankeln, T., 2013. Phylogenetic analyses of endoparasitic Acanthocephala
based on mitochondrial genomes suggest secondary loss of sensory organs. Mol. Phylo-
genet. Evol. 66, 182e189.
Wei, T.P., Chen, H.C., Lee, Y.C., Tsai, M.L., Hwang, J.S., Peng, S.H., Chiu, Y.W., 2013.
Gall polymorphism of coral-inhabiting crabs (Decapoda, Cryptochiridae): a new
perspective. J. Mar. Sci. Technol. 21, 304e307.
Weinfurtner, G., 1989. Seeliliensteilglieder mit Parasiten. Fossilien 6 (2), 64e65.
Weitschaft, W., Guhl, W., 1994. Erster Nachweis fossiler Ciliaten. Pal€aontol. Z. 68, 17e31.
Wey-Fabrizius, A.R., Podsiadlowski, L., Herlyn, H., Hankeln, T., 2013. Platyzoan mito-
chondrial genomes. Mol. Phylogenet. Evol. 69, 365e375.
Wieder, R.W., Feldmann, R.M., 1992. Mesozoic and Cenozoic fossil isopods of North
America. J. Paleontol. 66, 958e972.
Wienberg Rasmussen, H., Jakobsen, S.L., Collins, J.S.H., 2008. Raninidae infested by para-
sitic Isopoda (Epicaridea). Bull. Mizunami Fossil Mus. 34, 31e49.
Wilkinson, I., Wilby, P., Williams, P., Siveter, D., Vannier, J., 2007. Ostracod carnivory
through time. In: Elewa, A.M.T. (Ed.), Predation in Organisms, a Distinct
Phenomenon. Springer, Berlin and Heidelberg, pp. 39e57.
Williams, J.D., Boyko, C.B., 2012. The global diversity of parasitic isopods associated with
crustacean hosts (Isopoda: Bopyroidea and Cryptoniscoidea). PLoS One 7 (4), e35350.
http://dx.doi.org/10.1371/journal.pone.0035350.
Wilson, G.D., Paterson, J.R., Kear, B.P., 2011. Fossil isopods associated with a fish skeleton
from the Lower Cretaceous of Queensland, Australiaedirect evidence of a scavenging
lifestyle in Mesozoic Cymothoida. Palaeontology 54, 1053e1068.
Wilson, M.A., Vinn, O., Palmer, T.J., 2014a. Bivalve borings, bioclaustrations and symbiosis
in corals from the Upper Cretaceous (Cenomanian) of southern Israel. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 414, 243e245.
Wilson, M.A., Reinthal, E.A., Ausich, W.I., 2014b. Parasitism of a new apiocrinitid
crinoid species from the Middle Jurassic (Callovian) of southern Israel. J. Paleontol.
88, 1212e1221.
Yusa, Y., Yamato, S., 1999. Cropping of sea anemone tentacles by a symbiotic barnacle. Biol.
Bull. 197, 315e318.
Zapalski, M.K., 2011. Is absence of proof a proof of absence? Comments on commensalism.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 302, 484e488.
Zbinden, A., 1991. Seeigel mit Myzostomiden. Fossilien 8 (5), 266.
Zibrowius, H., 1981. Associations of Hydrocorallia Stylasterina with gall-inhabiting Cope-
poda Siphonostomatoidea from the south-west Pacific; Part 1. On the stylasterine hosts,
including two new species, Stylaster papuensis and Crypthelia cryptotrema. Bijdr. Dierkd. 51,
268e286.
CHAPTER SEVEN

A Prejudiced Review of Ancient


Parasites and Their Host
Echinoderms: CSI Fossil Record
or Just an Excuse for
Speculation?
Stephen K. Donovan
Department of Geology, Naturalis Biodiversity Center, Leiden, The Netherlands
E-mail: Steve.Donovan@naturalis.nl

Contents
1. Introduction 292
2. Interpretations and Confidence 294
2.1 Problems of interpretation 295
2.2 Limits of confidence 297
3. Some Examples 298
3.1 A coralecrinoid association from the Mississippian 299
3.2 A growth deformity in a Mississippian crinoid 302
3.3 Epizoobionts infesting a Mississippian crinoid 304
3.4 Platyceratid gastropods infesting Upper Palaeozoic crinoids 306
3.5 Site selectivity of pits in echinoid tests, Upper Cretaceous 313
4. Discussion 319
5. Conclusions 323
Acknowledgements 323
References 323

Abstract
Recognizing the presence of a parasite and identifying it is a relatively straightforward
task for the twenty-first century parasitologist. Not so the pursuit of ancient parasites in
fossil organisms, a much more difficult proposition. Herein, Boucot’s seven-tiered
scheme of reliability classes is applied as a measure of confidence of the recognition
of putative parasitism in two echinoderm classes, Upper Palaeozoic crinoids and a
Cretaceous echinoid (high confidence is 1, low confidence 7). Of the five examples,
the parasitic(?) organism is preserved in only two of them. A zaphrentoid coral on
the camerate crinoid Amphoracrinus may have robbed food from the arms (Category
1 or 2B). A pit in what appears to be a carefully selected site on the disparid crinoid
Synbathocrinus is associated with a growth deformity of the cup (Category 4). Multiple
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.05.003 All rights reserved. 291
292 Stephen K. Donovan

pits in an Amphoracrinus theca are also associated with a deformed cup, but it is more
difficult to interpret (Category 4 or 7). Some specimens of the camerate crinoid
Neoplatycrinites have circular grooves or depressions posteriorly, presumably produced
by coprophagic/parasitic platyceratid gastropods (Category 1). Site selectivity of pits in
the echinoid Hemipneustes places them preferentially adjacent to respiratory tube feet
(Category 4). From these examples it is deduced that sparse infestations of borings or
epizoozoic organisms permit a more confident interpretation of organism/organism in-
teractions; dense accumulations, possibly following multiple spatfalls, mask such
patterns.

1. INTRODUCTION
Consider the task of the parasitologist. There may or may not be an
actual physical manifestation of the attentions of a parasite in a human being,
sheep, favourite pet dog or fish bought from a fishmonger, but its presence is
easily confirmed (or otherwise) using the wide range of biomedical, biochem-
ical and gene sequencing tools currently available. A stool is collected, blood
sampled or the fish cut open, and the hunt is on. Not only can the parasite be
identified, but its life cycle elucidated and its genes sequenced. In the twenty-
first century, all this can be done before lunch.
The hunter of fossil parasites, in my own case, in ancient echinoderms,
has none of these advantages. We may confidently speculate that many or
most parasites in ancient echinoderms left no evidence of their presence and
we will never know that they were ever there. Such navel contemplation is
the easy part. Commonly, where there is evidence of some interaction
between an organism, parasite or not, and the host, this is preserved as
some invasive structure (boring) into the echinoderm’s endoskeleton.
This may or may not be associated with a growth deformation of the
host. Borings without growth deformities must be regarded as not proven
parasitism, as such traces of interaction can occur after the death of the echi-
noderm, either soon after (Donovan, 2014; Donovan et al., 2014a) or mil-
lions of years later (Donovan and Lewis, 2011; Donovan, 2013). Associated
growth deformities show that the boring was engendered while the echino-
derm was alive and able to react to the invasion. What commonly does not
get preserved is the producer of the boring, which almost invariably lacks a
mineralized (and thus easily preserved) skeleton, so our interpretation of its
purpose (parasitism, or failed or successful predation, or something else)
must be made in ignorance of the simple fact, whodunit? The parasitologist
looks for parasites; the palaeoparasitologist works by necessity without
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 293

parasites, except for rare examples such as parasites with preservable skele-
tons such as parasitic snails (see below) or rare mineralized producers of pits
such as Phosphannulus M€ uller et al., 1974 (Welch, 1976; Werle et al., 1984;
Boucot, 1990, pp. 32e34).
So, what can we hope to achieve? As already intimated, many palaeon-
tological investigations of echinoderms and putative parasites begin with the
recognition of pits and boreholes in the echinoderm endoskeleton. Pits, bor-
ings and related structures in ancient substrates, such as tracks, trails, burrows
and coprolites, are trace fossils, representing evidence of ancient organic ac-
tivity. The study of trace fossils is ichnology. Only rarely is the producing
organism preserved adjacent to a trace except in certain settings, such as
the shells of boring bivalves which may be preserved in their borehole in
limestone (for example, Donovan and Jagt, 2013a). It is a mantra of ichnol-
ogy, the study of trace fossils, that a given organism may have produced a
range of trace fossils, representing different activities, and that a particular
morphology of trace fossil may have been produced by more than one group
of organisms, involved in similar activities. Trace fossils are given Latinized
binomens, but these do not refer to an organism per se, but to the sedimen-
tary structure that is the trace fossil. That is, their classification as sedimentary
structures is analogous to, but not part of, Linnean classification. It is a com-
mon ‘game’ in ichnology to speculate on the identity of producing organ-
isms, but it is incorrect to be too dogmatic about such assertions (Donovan,
2010). Pits in live echinoderms commonly produce growth reactions in the
endoskeleton, either inhibiting growth or causing swellings of various mor-
phologies, but these are pathologies of the echinoderm and are not part of
the trace fossil (Donovan, 2015). Further, trace fossils are named on
morphology and not substrate (Pickerill, 1994); for example, it was
suggested that borings in the echinoderm test that cause a range of growth
deformations should be referred to Tremichnus Brett, 1985, but all morpho-
logically similar small round holes are more correctly referred to Sedilichnus
M€ uller, 1977, a senior synonym of the widely used Oichnus Bromley, 1981
(Zonneveld and Gingras, 2014; see also Pickerill and Donovan, 1998;
Donovan and Pickerill, 2002).
If this sounds daunting, then add to this the confusion introduced by the
taphonomy, the study of the preservation of our infested echinoderm. There
are some organisms which have simple skeletons that are commonly pre-
served whole, such as gastropods, shelled cephalopods, tube-dwelling
worms, foraminiferans and stony corals, or disarticulated into just a few
pieces, including bivalve molluscs, brachiopods and ostracods. Most other
294 Stephen K. Donovan

groups of organisms have complicated, multielement skeletons that fall apart


soon after death; one individual can be preserved as tens or hundreds of
individual pieces. The latter include the vertebrates, arthropods, plants and
echinoderms. Commonly, evidence of a possible parasitic interaction is
found in a fragment of echinoderm that cannot be classified with any con-
fidence any closer than class or order. And, of course, it may be that the para-
sitic infestation weakened the skeleton so that it is more prone to
disarticulation postmortem than otherwise.
So, palaeoparasitology of echinoderms most commonly involves recog-
nizing pathological growth deformities produced in response to structures
generated by parasitic(?) organisms. The producing organisms are not them-
selves commonly preserved or otherwise identifiable (although two specific
examples where the parasite is preserved are discussed below) and the echi-
noderm may be preserved as only a fragment of endoskeleton that can only
be classified to class level, such as a fragment of crinoid column. If we accept
this as the starting point for this paper, then we can, at least, describe some
good examples that lead to reasonable interpretations.
The morphological terminology of the crinoid endoskeleton is explained
in Ubaghs (1978) and Moore et al. (1978); for that of the echinoid test, see
Smith and Kroh (2011). Specimens discussed below are deposited in the
following institutions: Department of Earth Sciences, the Natural History
Museum, London, England (BMNH); Naturalis Biodiversity Center,
Leiden, the Netherlands (RGM); and Natuurhistorisch Museum Maastricht,
Maastricht, the Netherlands (NHMM).

2. INTERPRETATIONS AND CONFIDENCE


A great diversity of disease-carrying and parasitic organisms infest
extant echinoderms (Jangoux, 1987a,b,c,d), although they are rarely para-
sitic themselves (Rouse in Rhode, 2005, pp. 248e250). It is only in exam-
ples where infestations of ancient echinoderms caused some recognizable
deformity in skeletal growth, or the exceedingly rare cases in which the
infesting organism itself is fossilized, that an indication of disease or parasitism
can be identified in the fossil record (Conway Morris, 1981; Boucot, 1990;
Littlewood and Donovan, 2003). For example, parasites may produce
distinctive galls or cysts in echinoid radioles (Warén and Moolenbeek,
1989), asteroid arms, crinoid pinnules and cirri, and ophiuroids discs
(Grygier, 1988). Examples of non-cyst-like structures produced by
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 295

purported parasites which are considered below infest Upper Palaeozoic


crinoid thecae and Cretaceous echinoid tests.

2.1 Problems of interpretation


Interpretations of exotic interactions between ancient (or even extant) or-
ganisms are commonly problematic and rarely leads to hard conclusions,
apart from examples such as simple encrustations or boring invasions with
numerous modern analogues. For an example of the problem of interpreting
ancient interactions with echinoderms which is not parasitism, but certainly
exotic, consider the lectotype specimen of the primitive mollusc Phthipodo-
chiton thraivensis (Reed, 1911) (see Sutton and Sigwart, 2012; Figure 1

Figure 1 Phthipodochiton thraivensis (Reed, 1911), BMNH G47258, lectotype, from the
Upper Ordovician of south-west Scotland. (After Donovan et al. (2010), Figures 2 and 3(c),
respectively.) (a) The specimen showing the valves of the chiton, preserved as natural
moulds and curving up from the lower centre to the right, past the label and towards
the top left. Scale bar represents 10 mm. (b) High-resolution X-ray microtomography
image of the complete gut contents (valves removed). Scale bar represents 1 mm.
296 Stephen K. Donovan

herein) from the Upper Ordovician Lady Burn Starfish Beds of south-west
Scotland. Mark Sutton and Julia Sigwart imaged the concealed structure of
this specimen using high-resolution X-ray microtomography (for full
details, see Donovan et al., 2010, pp. 935e936). This revealed a string of
shelly material preserved in a position that would correspond to the gut in
an extant chiton. The most prominent component of this string is a group
of nine crinoid columnals of more or less similar morphology that are inter-
preted as being derived from an individual of the camerate Macrostylocrinus
cirrifer Ramsbottom, 1961 (Donovan et al., 2011b), a common taxon at
Lady Burn.
Multiple lines of evidence indicate that this association is not a hydrody-
namic accumulation (Donovan et al., 2010). The columnals represent
ingested hard parts that were passing through the spiral gut of the chiton
at the time of death. Chitons are commonly thought of as simple grazers,
feeding on encrusting algae on rock surfaces, but many living chitons live
on animal matter (Fulton, 1975; Latyshev et al., 2004) and examination of
gut contents in living species shows a variety of food preferences, in some
species highly specialized (Sirenko, 2000).
Almost nothing is known regarding the predators and scavengers of
Ordovician crinoids (Meyer and Ausich, 1983; Baumiller and Gahn,
2003). Determination of predation and scavenging of Ordovician crinoids
has been mainly inferential, although there is suggestive evidence for pred-
atory decapitation in some disparids (Donovan and Schmidt, 2001), perhaps
a Category 3 association sensu Boucot (1990; see below). Category 1 asso-
ciations of crinoids as prey are rare throughout the fossil record, and the
example discussed herein is the oldest and the most unexpected.
Crinoid columnals may have been consumed by this chiton due to pre-
dation, scavenging or ingestion of sediment rich in crinoid bioclasts. The last
seems unlikely; unless the chiton was particularly unselective, it would more
likely have harvested organic matter from finer grained sedimentary parti-
cles. Predation or scavenging is thus more probable, but equally plausible.
The speculation of predatory versus scavenging behaviour may appear trivial
in this example, but it is analogous to the sometimes heated discussions of the
habits of Tyrannosaurus rex Osborn, 1905 e predator or scavenger e which
have engendered a diverse correspondence (for a recent discussion, see
Erickson, 2014). Whichever, such crinoidivory in this ancient Scottish
mollusc is unexpected, being unknown from extant chitons; like T. rex,
we lack a modern analogue to provide an answer.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 297

To give a further echinodermal example of an organismeorganism


interaction that defies entirely confident interpretation, consider the
well documented record of otherwise well-preserved Palaeozoic crinoids
that retain evidence of regeneration of one or more arms (see Baumiller
and Gahn, 2003, 2004). These are commonly interpreted as evidence of
successful predation on the arms of crinoids, but as of yet there is no
entirely satisfactory explanation of why anything should have been eating
such a poorly nutritious part of the animal. Extant crinoids carry their go-
nads mainly in their arms and pinnules (Breimer, 1978, pp. T18eT19),
making these nutritious organs the principal targets for predators. If eaten,
an arm can be regenerated; the comatulid Antedon bifida (Pennant, 1777) is
gravid throughout the year, presumably to tempt predators and divert
them from the visceral mass, although the reproductive season is only a
month (Nichols, 1994). Palaeozoic crinoids lacked this adaptation and it
is assumed that the gonads were included within the cup, although some
taxa may have adapted their large anal sacs to bear the gonads (Lane,
1984). Both I and other authorities (V.J. Syverson, October 2013, personal
communication) are of the opinion that predation on the arms of Palaeo-
zoic crinoids may be evidence of predation, not on the innutritious arms of
crinoids, but on some nutritious organism(s) or group(s) of organism(s) that
lived on the arms. This is entirely speculative as I am not aware of reports of
likely prey organisms being preserved on the arms; they may have been
unmineralized and thus unlikely to be fossilized. The evidence is good
for the predation of the arms of Palaeozoic crinoids, but not good for
the reason why.

2.2 Limits of confidence


The two examples given above e one particular, one general e demonstrate
something of the difficulties of making a precise interpretation of the inter-
actions between echinoderms and other organisms even when the data is
good. What is needed at this juncture is some qualitative measure that re-
searchers can apply to their deliberations to give their readers a ‘feel’ of
the level of confidence with which they are made. I suggest one way
forward, admittedly preliminary, is an adaptation of Boucot’s (1990,
pp. 9e10) original reliability classes which he applied to myriad examples
that he examined for evidence of behaviour and coevolution in the fossil re-
cord. The reader is referred to the original publication for more explanation
and many more examples, as this tabulation is merely a brief abstract of a
298 Stephen K. Donovan

much greater body of argument and evidence. Briefly, Boucot’s categories


are summarized thus:

Category 1 The rare examples where the evidence is incontrovertible.


Category 2A Organisms preserved in close association, but not actually in
position.
Category 2B Arguments based on functional analogy with closely related taxa.
Category 3 Fairly certain, based on known behaviour of living analogues.
Category 4 Less certain due to the producer of evidence being unknown.
Category 5A High degree of uncertainty about the maker.
Category 5B Examples where modern biogeographic evidence is crucial.
Category 6 Fairly speculative, such as functional determinations of wholly
extinct taxa.
Category 7 Highly speculative, little reliability.

Of the two examples given above, I would place the Ordovician crinoi-
divorous chiton in Boucot’s Category 2A and would change it to Category 1
if a modern analogue was discovered. Predation on the arms of Palaeozoic
crinoids is best assigned to Category 4 or, perhaps, 5A.

3. SOME EXAMPLES
The structure of this paper was determined by personal preference. It
could have written as the echinodermal parts of Conway Morris (1981)
brought up-to-date, a revision of the parasitism parts of various reviews of
echinoderm taphonomy (such as Meyer and Ausich, 1983; Donovan,
1991b) or a parasitism paper modelled on the review of predation on cri-
noids by Baumiller and Gahn (2003). Rather, I have something a little
different in approach from all of these, preferring to focus on a few selected
examples from my own experience rather than a review that attempted to be
too broad and, in consequence, lacked detail.
Actual examples of putative parasitic infestations of fossil echinoderms
are desirable at this stage to demonstrate something of the range of evidence
available and how it has been interpreted. Most of the examples lean heavily
on my own research, which explains why they preponderantly concern my
beloved crinoids. I consider it preferable to discuss the familiar in this review
than try to spread the text too thinly across the many extinct and extant clas-
ses of echinoderms. My intention here is to demonstrate broad principles
that can be extrapolated to a reader’s own investigations.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 299

3.1 A coralecrinoid association from the Mississippian


(Figure 2)
Material, locality and horizon: BMNH EE5797 (Figure 2) is from Salt-
hill Quarry, Clitheroe, Lancashire [SD 7550 4265], England, one of the
most important localities for Mississippian (Lower Carboniferous) echino-
derms in north-west England (Donovan et al., 2003; Kabrna, 2011). The
crinoid theca was found in the Cover Mudstone, present at the top of the
lower Viséan (¼ upper Chadian) Salthill Cap Beds of the Bellman Lime-
stone Member.
Description: (Based on Donovan et al., 2005, pp. 43e44.) BMNH
EE5797 is an undeformed calyx (Figure 2(a) and (b)) of the monobathrid
camerate crinoid Amphoracrinus gilbertsoni (Miller in Phillips, 1836). Preserva-
tion of camerate crinoids as their golf ball-like thecae, lacking most if any of
the contiguous arms and stem, is common at this locality. The most notice-
able variance is that Wright (1955, p. 195) noted ‘. posterior interbrachial
area [¼ CD interray] much wider than the others .’ BMNH EE5797 has
an AB interray that is almost as wide as the CD interray at the level of the
base of the free arms; the other three interrays are narrower and all of about
the same width.
The solitary rugose coral may be a zaphrentoid (Mitchell, 2003). It is
small in comparison to the crinoid calyx. It is preserved in transverse
section (Figure 2(c)) and is apparently less than 2.0 mm high, although it
has obviously been depressed into the crinoid calyx as indicated by a series
of more or less concentric fractures. More precise systematic identification
has not been possible on the basis of this single section. The coral is situated
in the AB interray, in close contact with the anterior branch of the B
ray arm.
Discussion: There are a number of lines of evidence to indicate that the
solitary coral was attached to the crinoid calyx in life (Donovan et al., 2005),
although an actual attachment area per se is not clearly apparent and,
assuming it to be present, could only be exposed by destructive techniques.
The corallite is small (Figure 2(c)). Although solitary rugose corals were typi-
cally unattached in adulthood, as larvae (and, plainly, in subsequent early
growth stages) they were capable of cementation to a variety of hard sub-
strates (see, for example, Hubbard, 1970, p. 203).
The coral corallite is preserved more or less perpendicular to the surface of
the crinoid calyx and has been pushed into it, being surrounded by concentric
cracks which have broken across plates, probably due to post-depositional
300 Stephen K. Donovan

Figure 2 (aec) BMNH EE5797, Amphoracrinus gilbertsoni (Miller, 1821, in Phillips,


1836), encrusted by a solitary rugose coral. (After Donovan et al. (2005), Figure 1.) Scale
bars represent 10 mm. (a) Theca of crinoid. A ray (¼ anterior) central; note the coral in
the AB interray (left) at the same level as the base of the free arms. The anal pyramid is
subapical and situated about mid-way between the apex of the tegmen and the arms
on the opposite side of the specimen to the A ray (¼ CD interray). Note that the AB
interray is broader than the EA interray (right of A ray arm). (b) Apical view. A ray (¼
anterior) towards top of page, AB interray upper right, CD interray (posterior) towards
bottom of page. Note regular pentagonal outline of undeformed specimen. (c) The
corallite of the coral. Note how the curved breakage of the crinoid around the coral
follows the outline of the coral theca and has not resulted in disarticulation of plates.
(def) BMNH E71430, Synbathocrinus conicus Phillips, 1836, dorsal cup and proximal
column, with single borehole of Sedilichnus paraboloides (Bromley, 1981). Scale bar
represents 5 mm. (d) Lateral view with E ray central. Note boring and oral surface
sloping anteriorly (that is, to left). (e) EA interray central, showing the prominent boring
in the R:R:B plate triple junction (compare with Figure 3(a)). (f) D ray, showing slope of
oral surface towards EA interray (compare with Figure 3(b)).
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 301

compactional strain. Although the position of the coral on the crinoid could
be merely coincidence, the unusual width, which we interpret as growth
rather than postmortem deformation (Figure 2(b)), of the AB interray is
reminiscent of the sorts of deformities that are known to be induced in
echinoderms by interactions with encrusting organisms (Meyer and Ausich,
1983).
All these lines of evidence support the supposition of an original
biotic relationship between the host crinoid and its attached coral in
life; that is, the living coral was an epizoozoan (sensu Taylor and Wilson,
2002) on the living crinoid. In the Mississippian of the British Isles, it is
more common to find crinoids, particularly their columns, encrusted by
colonial tabulate corals than by either solitary or colonial rugose corals
(see, for example, Donovan and Lewis, 1999). The close association of
A. gilbertsoni with a solitary rugose coral would thus be worthy of comment
under any circumstances, but it is the position with respect to the crinoid
calyx and the presumed life orientation of the latter that excites particular
comment. In attaching to the calyx of a crinoid, the solitary coral has
gained an advantage from its elevated position would have been above
the turbid bottom layers of the water column, so water currents would
have been essentially sediment free. The importance of feeding is further
emphasized by the corallite being hard against the B ray arm in the AB
interray. The A ray is anterior and, if A. gilbertsoni formed a parabolic filtra-
tion fanlike modern rheophilic crinoids (Macurda and Meyer, 1974), the
coral would have been directed into the clean water currents, being posi-
tioned up-current from the anal pyramid. The situation in close contact
with the B ray arm is at least suggestive that the coral may have actively
harvested food with its tentacles from the adoral groove of the crinoid’s
arm. Although obviously speculative, this supposition is supported by the
unusual width of the AB interray, a growth deformity which probably
resulted from a reaction to the coral and prevented it from interfering
with the feeding activity of the A ray arm, too. The available evidence is
at least highly suggestive that this coral was both a filter feeder and a parasite
on the crinoid.
Reliability: Category 1 or 2B. Boucot (1990, p. 9) stated (Category 2B)
that ‘With organisms belonging to extinct higher taxa [such as rugose corals
and camerate crinoids] functional analysis of behavior from morphology is
clearly less reliable’. But other groups of crinoids and stony corals are extant
and well known, so Category 1 is at least plausible.
302 Stephen K. Donovan

3.2 A growth deformity in a Mississippian crinoid


(Figures 2(def) and 3)
Material, locality and horizon: BMNH E71430 (Figures 2(d) and 3) is
from Salthill Quarry, Clitheroe, Lancashire [SD 7550 4265], England
(Donovan et al., 2003; Kabrna, 2011; Section 3.1 herein). The crinoid theca
was found in the lower Viséan (¼ upper Chadian) Salthill Cap Beds of the
Bellman Limestone Member.
Description: (Based on Donovan, 1991a, p. 2.) The specimen is a dorsal
cup, with the two most proximal columnals of the column, of the disparid
crinoid Synbathocrinus conicus Phillips, 1836 (Wright, 1952, pp. 134e136, pl.
36, Figures 10, 21 and 22). The cup has been bored once, in the EA interray,
at the triple suture between the E and A radials, and the basal in the EA
interray (Figures 2(d) and 3(a, b)). The boring is circular with a rounded
margin; the cavity is a conical, flat-bottomed pit which does not break
through into the body cavity and is assigned to Sedilichnus paraboloides
(Bromley, 1981). The cup is not swollen around the excavation, unlike
the common reaction to borings seen on some crinoid columns from the
same site, but the cup is less developed on the bored side, the oral surface
sloping towards the EA interray (Figures 2(f) and 3(b)).

Figure 3 BMNH E71430, Synbathocrinus conicus Phillips, 1836. (a, b) Camera lucida
drawings of the dorsal cup and proximal column. (After Donovan (1991a), Figure 2.)
Scale bar represents 5 mm. (a) EA interray central (compare with Figure 2(e)), showing
the position of the boring. (b) DE interray central, showing how the oral surface slopes
down towards the region of infestation, presumably a growth reaction to the distur-
bance. Key: RR ¼ radial circlet; BB ¼ basal circlet; COL ¼ most proximal part of column.
(c) Schematic oral view based on BMNH E71430. (After Donovan (1991a), Figure 3.)
Carpenter rays (AeE) indicated, corresponding to the positions of the arms; X is the po-
sition of the anal series (¼ posterior). Small arrows mark positions of cryptic B:B:R plate
triple sutures; large arrows mark positions of prominent R:R:B plate triple sutures. Key:
1 ¼ in the same position of the anal series; 2 ¼ adjacent to the anal series; 3 ¼ distant
from the anal series; * ¼ position of boring. Current flow would have been from the top
of the page.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 303

Discussion: Borings in the cup of Synbathocrinus are rare. Brett (1978)


suggested that pit-forming organisms on Palaeozoic crinoids were host spe-
cific; I have looked at hundreds of specimens of this genus, from the Missis-
sippian of North America and Europe, and the Permian of Timor, for almost
25 years and have failed to find a second bored specimen.
The excavating organism selected its site with some precision. That
the boring was made in a live crinoid is undoubted, with growth pro-
ceeding with greater vigour on the side opposite to the pit (Figures 2(f)
and 3(b)). In life, all sides of the cup would have been equally available
for infestation by borers and encrusters with the crown elevated above
the substrate. The boring was made precisely at the triple suture between
two radial (R) and a basal plates (B) (Figures 2(d) and 3(a)). The three
B:B:R triple junctions, in the A, C and E rays are cryptic and can only
tentatively be distinguished; in contrast, the five R:R:B sutures are readily
apparent (Figures 2(def) and 3(c)). Sutures between plates, bonded
together by collagenous ligament fibres, would have been easier to bore
into, either mechanically or chemically, than the solid calcite of the plates;
therefore, the position of S. paraboloides is interpreted as a choice made for
ease of infestation by the boring organism, the identity of which remains
unknown.
Why bore into the EA interray, rather than at one of the four other triple
plate sutures? The sutures labelled 3 in Figure 3(c), in the EA and AB inter-
rays, are the most up-current (¼ anterior) plate triple junctions and, thus,
they are also the potential borehole sites most removed from the anal series
(X). This suggests that the pit-forming organism was not a coprophage. The
most probable life strategy of the pit-former was as a hard substrate dweller
that fed by filtration (compare with Brett, 1978, 1985). It would be attached
to the highest ‘fixed’ point of the crinoid, just below the oral surface and the
arms, and always orientated into the prevalent current by the crinoid. It is
presumed to have removed part of the suspended particulate food that
would otherwise have been ingested by the crinoid. Whether this contrib-
uted to the deformation of the cup or it was mainly a reaction to the forma-
tion of the pit must remain uncertain.
Reliability: Category 4. ‘For some types of behavioral evidence there is
an even larger degree of uncertainty about the maker of this evidence. This is
particularly true for many trace fossils’ (Boucot, 1990, p. 9). No determina-
tion of the borer can be made apart from that it was small, unmineralized and
probably a filter feeding invertebrate.
304 Stephen K. Donovan

3.3 Epizoobionts infesting a Mississippian crinoid (Figure 4)


Material, locality and horizon: BMNH EE8728 (Figure 4) is from Salthill
Quarry, Clitheroe, Lancashire [SD 7550 4265], England (Donovan et al.,
2003; Kabrna, 2011; Section 3.1 herein). The crinoid theca was found in
the Cover Mudstone, present only at the top of the lower Viséan (¼ upper
Chadian) Salthill Cap Beds of the Bellman Limestone Member.

Figure 4 Multiple Sedilichnus paraboloides (Bromley, 1981) infesting the dorsal cup of
the Lower Carboniferous crinoid Amphoracrinus gilbertsoni (Miller in Phillips, 1836),
BMNH EE8728. (After Donovan et al. (2006), Figure 1.) (a) Base of dorsal cup, D-ray to-
wards top of page. (b) Enlargement of (a), showing how pits are concentrated on plates,
not crossing sutures. (c) Enlarged lateral view of (mainly) dorsal cup, E-ray central,
showing sub-horizontal arrangement of closely spaced pits. (d) Theca in lateral view,
same orientation as (c). Note absence of pits above the line of the arm facets. Specimen
whitened by ammonium chloride. Scale bars represent 10 mm (a, d) or 5 mm (b, c).
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 305

Description: (Based on Donovan et al., 2006, pp. 43e44.) The reader


has already been introduced to A. gilbertsoni and S. paraboloides (see above).
This description will concentrate on the pattern of infestation of pits on the
crinoid theca (Figure 4).
The theca preserves the dorsal cup, fixed arms and tegmen. Sedilichnus
paraboloides is found on all plates of the dorsal cup (basal and radial circlets),
not including the articular facet for the column (Figure 4(a) and (b)); on
fixed brachial plates (Figure 4(c)); and on interbrachial plates up to just above
the mid-height of the facet for the free arms (Figure 4(c), upper left). There
are no pits on the tegmen above this level (Figure 4(d)). The hemispherical
pits may be close packed together on individual plates, but are rarely in con-
tact; they do not occur on the depressed sutures between thecal plates, that
is, those of the dorsal cup and proximal fixed brachials (Figure 4(c)). Pits
(n > 50, all c.1 mm diameter) are common throughout the defined area
and ‘bald’ areas seen on some parts of the specimen are covered with pits
elsewhere in analogous positions in relation to the arms. The base of the
cup is angled to the long axis of the theca and slopes up towards the CD
interray.
Discussion: This infestation occurred while the crinoid was alive. Most
obviously, pits are distributed on the theca, below the level of arms,
through 360 , but are not found on articular facet of the column. Thus,
the embedding organisms had access to the entire lower half of the theca,
but the column facet would have been covered by the column in a living
crinoid. The absence of pits on the tegmen above the level of the arms is
most easily explained by the crinoid itself keeping this region free of any
infestation by the action of the tube feet. These are situated on the adoral
surface of the free arms, enabling them to ‘clean’ the tegmen, but not the
region below arm level. Presumably, the proximal column and aboral sur-
faces of the free arms, which are not preserved, could also have been
infested. Further, there is evidence for a growth deformity, that is, the
sloping base of the theca, which is analogous to the specimen of S. conicus
described above (Section 3.2).
The palaeoecology of the pit-forming organism is, in part, decipherable,
even if its identity remains obscure. There is the obvious segregation of
S. paraboloides on the substrate (see above), indicating that they were unable
to survive the attentions of the tube feet. The pits are concentrated on the
more elevated parts of the lower half of the theca, that is, they occur on the
more concave, central areas of plates (Figure 4(b) and (c)). The plate sutures,
which would presumably be easier regions for embedment (Section 3.2), are
306 Stephen K. Donovan

avoided; these sutures are sunken between the plates, so it is probable that
the small difference in height between sutures and plate centres was impor-
tant to the producers. This strongly supports a suggestion that they were fil-
ter feeding. Interestingly, S. paraboloides is distributed on the lower part of
the theca through 360 , and is apparently not selective with respect to the
life orientation of the crinoid, there being no obvious preference for anterior
(A-ray) or posterior (CD-interray) (but see Section 4, ‘Discussion’, below).
Pits are close packed, but in only very few instances overlap, suggesting that
this represents a gregarious accumulation of an organism that thrived on
elevated calcareous substrates, rather than successive pits formed by one or
a few individuals (compare with Donovan and Lewis, 2010).
Reliability: Category 4 (compare with Section 3.2) or 7, ‘. so highly
speculative as to have little reliability at all .’ (Boucot, 1990, p. 9). The
sloping base of the theca is analogous to the example of S. conicus, above,
in which a single pit is interpreted as engendering the aberration. In the pre-
sent example, with the theca heavily covered by borings below arm level,
there seems little connection between a growth deformity and an infestation
through 360 . The pits would have gained advantages attached to an elevated
cup, but any evidence that the association may have been parasitic is masked
by the dense accumulation of pits. The pit producer may have been conspe-
cific with that in Section 3.2; the specimens are from the same locality.

3.4 Platyceratid gastropods infesting Upper Palaeozoic


crinoids (Figures 5 and 6)
Material, locality and horizon: All specimens are members of the mono-
bathrid camerate genus Neoplatycrinus Wanner, 1916, from the Permian of
West Timor. RGM B9, Neoplatycrinus major Wanner (Figure 5(a) and (b)),
from Basleo (Charlton et al., 2002, Figure 2), Noil Tonini, West Timor (pre-
sumably about sites 6e8 of Webster, 1998). RGM ST.32842[1], Neoplatycri-
nus dilatatus Wanner, 1916 (Figure 5(cef)), comes from Toenino. RGM
T.4439[1], Neoplatycrinus sp. cf. N. dilatatus Wanner (Figure 5(g)), no associ-
ated locality data. RGM T.3851[1], N. dilatatus Wanner (Figure 6(aec)), no
associated locality data.
Description: (After Donovan and Webster, 2013, pp. 990e991.) The
gross morphologies of the Neoplatycrinus crinoid species from West Timor
were described in Wanner (1916); herein, only the modifications to the ge-
ometry of the form of the CD (posterior) interray in a few specimens are
described. Two broad morphologies are identified, one centred below
(Figure 5(a)) and the other centred on the periproct. RGM B9 has a theca
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 307

Figure 5 Homing scars of platyceratid gastropods (Lacrimichnus? isp.) on Permian


platycrinitid crinoids from West Timor. (After Donovan and Webster (2013), Figure 1.)
(a, b) RGM B9, Neoplatycrinus dilatatus Wanner, 1916, posterior (CD interray central)
and anterior (A ray central) views of theca, respectively. Note deeply sunken homing
scar. (cef) N. dilatatus Wanner, 1916, RGM ST.32842[1]. (c) Posterior view, depressed
CD interray central. (d) Lateral view, B ray slightly left of central, note posterior flat-
tening of theca (right). (e) Anterior view, A ray central. (f) Basal view, posterior towards
top of page. (g) RGM T.4439[1], Neoplatycrinus sp. cf. N. dilatatus Wanner, 1916, CD inter-
ray viewed slightly obliquely from below and showing deep depression in radials. Spec-
imens uncoated. All scale bars represent 10 mm.
308 Stephen K. Donovan

Figure 6 Platycrinitid crinoids and platyceratid gastropods. (aec) Neoplatycrinus dilata-


tus Wanner, 1916, RGM T.3851[1], Permian of West Timor. (After Donovan and Webster
(2013), Figure 3.) (a) Posterior view (CD interray central) showing depressed area around
periproct, Lacrimichnus? isp., and on the C and D radials, extending onto basal circlet.
(b) Anterior view, A ray central. (c) Lateral view, BC interray central, note posterior flat-
tened in this view (left) and sloping towards base. (def) Platyceratid gastropod pre-
served on the tegmen of Platycrinites s.s. wachsmuthi (Wanner, 1916). (After Webster
and Donovan (2012), Figure 1(aec).) (d) Posterior view. (e) B and C interray. (f) Oral
view. Specimens uncoated. Scale bars represents 10 mm.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 309

that is typically bulbous anteriorly (Figure 5(b)), but, viewed from the side
(BC and DE interrays), the CD interray is flattened and slopes down to the
basals, slightly protruding just below the level of the arms. The radials of the
CD interray bear a prominent sub-circular groove, c. 8.9e9.7 mm in diameter,
situated below the periproct and more on the D than C radial (Figure 5(a)); a
comparable structure was illustrated by Wanner (1916, pl. 99, Figure 1(a)).
The central area is flattened on the C side and onto the D, but otherwise
slightly raised. Sutures between the radials are poorly apparent in this region.
Outside the groove, the test is sculpted by two (D) or three (C) sub-concen-
tric raised ridges. These structures do not encroach onto the basal circlet. This
suggests that they were not produced by the shell of the epizoozoic snail
(sensu Taylor and Wilson, 2002), which would not have been so careful as
to avoid adjacent plates in this manner. Rather, they are more likely a stereom
response to gastropod infestation (compare with, for example, Grygier, 1988;
Eckert, 1988; Grygier and Nomura, 1998; amongst many others).
The other three thecae (Figures 5(ceg) and 6(aec)) have a similar
structure to each other, albeit a little different from RGM B9. These three
thecae are distinctly truncated posteriorly (Figures 5(d, f) and 6(c); as is
also RGM B9). The radials in the CD interray form a depressed region
below the periproct, like an inverted triangle with bowed sides, with the
lateral extent determined by the positions of the arm facets. These depressed
regions are approximately parallel-sided to convex below the arms (Figures
5(c, g) and 6(a)), becoming constricted towards the base of the thecae.
The plate triple junction between the C and D radials, and the supporting
CDeDeDE basal are depressed. RGM T.3851[1] (Figure 6(aec)) has the
most rounded depressed region with a shallow, sub-circular groove appar-
ently extending around the periprioct. Comparable structures are not
apparent on RGM T.4439[1] and ST. 32842[1] (Figure 5(c) and (g)).
Discussion: The coprophagic/parasitic relationship between the platy-
ceratid gastropods, and Palaeozoic crinoids and blastoids has been recognized
for over 140 years. It persisted from at least the Middle Ordovician to
Permian (Bowsher, 1955; Conway Morris, 1981, p. 499; Meyer and Ausich,
1983, pp. 401e403, Figure 5; Baumiller, 1990, 1993; Boucot, 1990;
Donovan, 1991b, pp. 251e252; Gahn and Baumiller, 2003, 2006; amongst
others). The mollusc is commonly preserved surmounting the anal opening
(Figure 6(def)) where it produces no stereom overgrowth (but sometimes a
‘scar’), presumably because the platyceratid is vagile. Bowsher (1955) consid-
ered the platyceratidepelmatozoan association to be obligate commensalism,
the gastropod being a coprophage, feeding on the echinoderm’s excrement.
310 Stephen K. Donovan

Yet Rollins and Brezinski (1988) demonstrated that crinoids with platycera-
tid infestations tend to be smaller than those without associated gastropods,
thus indicating that the association was detrimental to the host, that is, para-
sitic. The platyceratid may have utilized the detritus concentrated by the
filtration fan of the crinoid. Further, Baumiller (1990) demonstrated that
some platyceratids drilled into the crinoid tegmen, which further suggests a
parasitic association (see also Baumiller, 1993; Baumiller et al., 1999).
One of the common associations by platyceratids in the Late Palaeozoic
was with the platycrinitid monobathrid camerates. The Permian of West
Timor has provided some of the youngest specimens to demonstrate this
interaction (e.g. Wanner, 1937, pl. 11, Figures 7 and 8; Baumiller et al.,
2004, pp. 393e395, Figure 1; Donovan and Webster, 2013). Specimens
from Timor discussed by Webster and Donovan (2012) either preserved
(or, at least, preserved evidence of) platyceratids on the tegmen or along
the radial summit of the platycrinitids Platycrinites Miller and Neoplatycrinus
Wanner (Figure 6(def)). The specimens described above do not retain
any platyceratid shells, but instead show evidence of different patterns of
infestation in the CD interray of the theca first discussed by Donovan and
Webster (2013). In these associations, the crinoid thecae show distinct modi-
fied morphologies, presumed to have been formed in response to the
gastropod infestation; whether these changes also altered the feeding capa-
bilities of the crinoids is equivocal. Part of the fascination of this association
is its age; ‘Hard substrate communities are poorly known in the Permian’
(Taylor and Wilson, 2003, p. 48).
The circular structures (Figure 5(a)) and depressed areas (Figures 5(c, g)
and 6(a)) of abutting C and D radials on the Neoplatycrinus thecae described
above are interpreted to be the results of the crinoids reacting to infestations
by platyceratid gastropods. Unlike other examples known from the Permian
of West Timor (Wanner, 1937; Baumiller et al., 2004; Webster and Dono-
van, 2012; Figure 6(def) herein), these snails were attached subapically in
the CD interray rather than apically on the tegmen or adjacent to it. This
was undoubtedly, at least in part, due to the posterior position of the peri-
proct of Neoplatycrinus. Note that these patterns of infestation are differently
positioned, either below the periproct, or extending between the C and D
arms to include the periproct.
The two broad morphologies of CD interray described are thus indica-
tive of differing positions chosen by the infesting gastropods and reactions
engendered by their hosts. Most obviously, the circular scars indicate the
homing positions of the producing gastropods (compare with Boucot,
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 311

1990, Figure 8). These scars are prominent external modifications of the cri-
noid’s endoskeleton. They may be the result of abrasion by the edge of the
aperture of the gastropod shell or chemical etching (Bromley, 2004, p. 463),
irregular growth of stereom as a reaction to this by the crinoid or, most prob-
ably, a combination of all of these. For example, RGM B9 (Figure 5(a)) has
both a deeply depressed, circular groove (situated below the periproct;
contrast with Figure 6(a)), but also concentric raised ridges outside it. The
former is presumably produced (mainly) by the gastropod, the latter by
the crinoid, raised above the level of the adjacent theca. It is considered
unlikely that these concentric structures represent occupancy by successively
larger platyceratids through time, although they could reflect the growth of
a single individual gastropod.
Despite these modifications, the gross, slightly inflated morphology of
the crinoid theca is retained, except posteriorly. The gastropod must have
stood proud of the surface of the theca, presumably giving the CD interray,
an inflated appearance in life. Whether the gastropod remained in this
position at least semipermanently is unknown, but these modifications to
the crinoid suggest that it was at least likely. Figure 6(def) shows a platycer-
atid gastropod from this locality so that its size and shape e low, moderately
broad and cap-like e may be compared to the other illustrations herein; its
position on the theca is in sharp contrast to where snails must have infested in
Figures 5 and 6(aec).
The round groove in the radials in the CD interray is a trace fossil that is
close in morphology to Lacrimichnus Santos et al., 2003, based on Neogene
gastropod attachment scars. Jagt (2007) identified an older Lacrimichnus? isp.
on a barnacle plate from the upper Maastrichtian of the type area. Lacrimich-
nus? isp. is also applied as an ichnotaxonomic name to the round scars pro-
duced by Permian capuliform platyceratid gastropods, such as the specimens
illustrated herein and by Boucot (1990, Figure 8). This would rescue such
trace fossils from the limbo that referring to them as circular grooves engen-
ders (compare with discussion by Donovan and Pickerill, 2004, p. 483).
Although produced by acorn barnacles rather than gastropods, the ichnoge-
nus Anellusichnus Santos et al., 2005, is broadly similar to Lacrimichnus, but
typically has an undulating outer furrow.
The second morphology described above involves the CD interray of
the crinoid being depressed, more so than in RGM B9, but without
obvious homing scars except, possibly, faintly in RGM T.3851[1]
(Figure 6(a)). Commonly, the test of uninfested Neoplatycrinus is bulbous,
conical or bowl-shaped. Infested tests retain this geometry in anterior
312 Stephen K. Donovan

view (Figures 5(e) and 6(b)), but, in other views, there is a marked flattening
posteriorly (Figures 5(d, f) and 6(c)), giving the test a prominent anteriore
posterior asymmetry, although remaining bilaterally symmetrical along a
mirror plane AeCD (that is, anterioreposterior). Such flattening is only
developed in the CD interray, that is, posteriorly. This is interpreted as a
different expression of infestation by a platyceratid.
The depressed CD interray produces a marked change in thecal gross
morphology, but, in life, the shell of the coprophagic platyceratid would
have been positioned here, presumably filling the depression. The shells
of these gastropods are low and cap-like (Bowsher, 1955; Boucot, 1990,
Figures 10e13; Webster and Donovan, 2012, Figures 1(aec) and 2(deh);
Figure 6(def) herein). It may have been that the gastropod in part ‘restored’
the gross morphology of the theca with its shell by filling the depressed area.
That is, with a shell in place in the depressed CD interray, the combined
morphology of crinoid þ gastropod may have approximated that of an
uninfested theca. The same may be true for the (less depressed) RGM B9.
This interpretation is speculative, but could be tested by a carefully defined
experiment in a flume tank.
If this supposition is accepted as possible, the question must be asked why
was it advantageous to host, coprophage or both? The gross morphology of
the crown presumably played a part in ensuring water currents for feeding
reached the arms (Jefferies, 1989). If, as I suggest, the platyceratid ‘filled
the gap’ in the theca, made as a response by the crinoid to the snail’s presence
in the CD interray, then the more or less restored hydrodynamics of the
crown would have engendered more efficient suspension feeding by the
crinoid. This would have resulted in greater production of faeces which,
in turn, would have benefitted the gastropod. But, overall, the relationship
was probably more parasitic than mutualistic.
Such a relationship was more than casual, and the modifications of the
theca that favoured this crinoidegastropod association were probably devel-
oped early and retained throughout life. The two morphologies of CD
interray seen in Neoplatycrinuseplatyceratid associations are indicative of
two different patterns of infestation. The depressed CD interrays (particu-
larly Figures 5(c, g) and 6(a)) suggest a mode of infestation in which the
gastropod extended between the C and D arms, perhaps covering the
periproct. This arrangement suggests a long-term relationship between
crinoid and gastropod, the former showing large thecal modifications to
accommodate the snail.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 313

In contrast, other Neoplatycrinus grew to produce a typically radially sym-


metrical theca with no depressed regions in any ray. If, in maturity, such a
theca became infested, the thecal outline could not be modified as much
and the CD interray developed a homing scar morphology, positioned below
the periproct and not between the arms, that is, somewhat different in gross
form (for example, compare Figure 5(a) with Figures 5(c, g) and 6(a)).
It is an axiom within the life sciences that, when a host goes extinct, so
does its biota of parasites and commensals specific to that host (Vickery and
Poulin, 1998; Dobson et al., 2008; Dunn et al., 2009). When the camerate
crinoids went extinct at the end of the Palaeozoic, presumably so did the
associated platyceratids. However, this may not have brought to an end
this biotic association that spanned over 200 million years. Although the
camerates became extinct, platyceratids persisted at least until the Late
Triassic (Bandel, 1992, 2007; Bandel and Frýda, 1999) or even Mid Jurassic
(Carmel Limestone of Utah; P.D. Taylor, written comm.), although their
associations with Mesozoic crinoids, if any, are equivocal.
Reliability: ‘Category 1 for those taxa found in place on a host pelma-
tozoan. Despite the fact that the platyceratids and their host pelmatozoan
taxa are extinct, there is enough information provided by a wealth of spec-
imens preserved in situ to leave little doubt about the reliability of the basic
coprophagous relation. Platyceratids found unassociated with pelmatozoans
pose the serious question concerning whether or not all platyceratids were
coprophages. It appears most reasonable to conclude that only those platy-
ceratid taxa closely associated with pelmatozoans were potentially coproph-
ages, and that only those platyceratid taxa actually found in situ on a host
pelmatozoan may be safely considered to have been coprophages’ (Boucot,
1990, p. 26). Specimens described herein with gastropod shell scars in the
CD interray might be considered Category 3 based on the above arguments.

3.5 Site selectivity of pits in echinoid tests, Upper


Cretaceous (Figure 6)
Material, locality and horizon: Many specimens of the holasteroid
echinoid Hemipneustes striatoradiatus (Leske, 1778) infested by the non-
penetrative pit Sedilichnus excavatus (Donovan and Jagt, 2002) in the
NHMM collections, including those discussed and illustrated by Donovan
and Jagt (2002, 2005; Figure 7 herein). The two specimens discussed in
detail below are bored and encrusted tests from the uppermost Maastrichtian
(Upper Cretaceous) of southern Limburg, the Netherlands. NHMM RZ
314 Stephen K. Donovan

Figure 7 Holasteroid echinoid Hemipneustes striatoradiatus (Leske, 1778) and the pits
Sedilichnus excavatus (Donovan and Jagt, 2002), all Upper Cretaceous (Maastrichtian)
of the Netherlands and Belgium. (After Donovan and Jagt (2002, Figures 2(a, c), 3(a, b)
and 5(aee).) (a, c, d) NHMM MK 4689. (a) Apical view, showing holotype (arrowed)
and paratype borings. (c) Holotype (arrowed) with two paratypes. Scale bar represents
5 mm. (d) Paratype with outline partly controlled by adjacent ambulacral column. Scale
bar represents 5 mm. (b) NHMM 699, right lateral view, paratypes; note the linear ar-
rangements of some groups of borings. (e, i) NHMM RN 452b, profile (test external sur-
face to left) and internal surface. (f, h) NHMM RN 452c, profile (test external surface to
left) and internal surface. (g) NHMM RN 452a, internal surface. Scale bars represent
10 mm unless stated otherwise.

00162 (Figure 8(aee)) is from the base of subunit IVf-6 of the Meersen
Member, Maastricht Formation, at the former Blom quarry, Berg en Terblijt
(for locality map, see Donovan et al., 2011a, Figure 1). NHMM MA 0234-1
(Figure 8(feh)) is from the same subunit (IVf-6), but from the nearby
Ankerpoort-Curfs quarry, now defunct, near Geulhem.
Description: (Based on Donovan and Jagt, 2013b, pp. 113e114) The
test of H. striatoradiatus, NHMM RZ 00162, was used as a substrate by a va-
riety of encrusting and boring invertebrates. The oral surface bears three
attached pycnodonteine oyster valves in close association; a cheilostome
bryozoan colony in the depressed region anterior of the peristome; more
than three poorly preserved spirorbid worm tubes; and a slot-like, figure-
of-eight boring aperture just posterior of the peristome, probably Caulostrep-
sis isp. (Figure 8(b)). Spirorbid worms occur at or just above the ambitus;
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 315

Figure 8 Holasteroid echinoid Hemipneustes striatoradiatus (Leske, 1778) and the pits
Sedilichnus excavatus (Donovan and Jagt, 2002), all Upper Cretaceous (Maastrichtian)
of the Netherlands and Belgium. (aee) NHMM RZ 00162. (After Donovan and Jagt
(2013b), Figure 1.) (a) Apical surface with the positions of four pits, S. excavatus, marked
by asterisks (*). The numerical designation of ambulacra (Roman numerals, IeV) and
interambulacra (Arabic numerals, 1e5) is provided as an explanation of the notation
used in the text. (b) Oral surface showing episkeletozoans (oysters, bryozoan colony,
spirorbids) and annelid boring, Caulostrepsis isp. (c, e) Two views of S. excavatus in
ambulacrum 4. (d) Sedilichnus excavatus in ambulacrum V; compare rounded outline
to (c, e). (feh) NHMM MA 0234-1. (After Donovan and Jagt (2013b), Figure 2.) All oblique
views of the apical surface. (f) Ambulacrum II (right anterior) central with closely group-
ed pair of S. excavatus pits. (g) Ambulacrum IV (left anterior) right of centre, with one
poorly developed S. excavatus in the upper part of the ambulacrum and two further
pits, in close association, in interambulacrum 3 (left anterior). (h) Ambulacrum V (left
posterior) right of centre with S. excavatus between the columns of pore pairs; other
specimens are in interambulacrum 4 (left lateral). An encrusting oyster, situated poste-
riorly, partly overgrows a S. excavatus (far right). Specimens uncoated. Scale bars repre-
sent 10 mm.
316 Stephen K. Donovan

a small circular hole in the posterior column of plates in ambulacrum V (left


posterior) is Sedilichnus simplex (Bromley, 1981). On the apical surface (that
is, supra-ambitally) are small pits and borings (S. simplex, probable Caulostrep-
sis isp.), four specimens of S. excavatus (Figure 8(a, cee)), spirorbids and a
fragment of a shell that may or may not be encrusting. Only S. excavatus oc-
curs on the upper half of the test.
Each S. excavatus is in close association with an ambulacrum of the echi-
noid. Sedilichnus excavatus occurs within ambulacrum II (right anterior), strad-
dling the perradial suture between the two plate columns (Figure 8(a)).
Anterior pores or pore pairs appear weakly developed adapically of the pit.
In ambulacrum I (right posterior), S. excavatus is anterior to the anterior
pore pairs, straddling the adradial suture with interambulacrum 1
(Figure 8(a)). Again, pores or pore pairs are only weakly developed adapically.
In ambulacrum V (left posterior), S. excavatus is lower on the test than
either pit on the right side and is anterior to the anterior column of pore
pairs, situated adjacent to the adradial suture with interambulacrum 4
(Figure 8(a) and (d)). This is the furthest away from a pore column of any
of the four pits; the adjacent pores appear to be normally developed. Each
of these first three pits is shallow; they are rounded to subangular rounded
in outline.
Sedilichnus excavatus in ambulacrum IV (left anterior) is deeper and has
straighter walls, the central boss being elongated; it occupies a central posi-
tion in the ambulacrum, straddling the perradial suture, but abutting the
anterior pore column with a straight side (Figure 8(a, c, e)). Pore pairs in
the anterior ambulacral column are only strongly developed abapically of
this pit. The posterior column of pores is strongly deflected towards this pit.
The second specimen, NHMM MA 0234-1 (Figure 8(feh)) is less mark-
edly encrusted, but more heavily pitted by S. excavatus. Other traces include
some small pits and perforations, situated apically, and referred to S. simplex
and Caulostrepsis? isp. Other scratch marks and excavations of (mainly)
ambulacra in this region are probably the spoor of grazing regular echinoids,
although typical Gnathichnus pentax Bromley, 1975, is not apparent. Only
S. excavatus is considered a pre-mortem infestation.
Sedilichnus excavatus is moderately common on this test with 11 pits, 4 of
which are closely associated with ambulacra (Figure 8(feh)). The ambulacral
infestations are high on the test, as in NHM RZ 00162. However, unlike
that specimen, ambulacrum I (right posterior) is not infested, whereas ambu-
lacrum II (right anterior) has two S. excavatus in close association
(Figure 8(f)). Ambulacra IV and V (left anterior and posterior, respectively)
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 317

each have one associated pit (Figure 8(g) and (h), respectively). In two inter-
ambulacra, pits are vertically one above the other (Figure 8(g) and (h)); in the
posterior interambulacrum (5; Figure 8(h), far right, oblique view) a pit sits
on the interradial suture at about mid-height of the test, about as far as it is
possible to be situated away from an ambulacrum, and partly overgrown by
an oyster.
Discussion: Donovan and Jagt (2002) erected the ichnospecies Oichnus
(now Sedilichnus) excavatus on the basis of locally common infestations of pits
in the late Maastrichtian echinoid H. striatoradiatus (Leske). Ample evidence
was presented to indicate that this association occurred while the echinoid,
which probably lived more or less epifaunally, was alive, most particularly by
the growth of internal blisters within the test and beneath the pits (Donovan
and Jagt, 2005, pl. 1, Figure 2; Figure 7(eei) herein). Sedilichnus excavatus was
diagnosed as ‘Circular to elliptical, non-penetrative Oichnus, almost invari-
ably with a broad, high, raised central boss. Aperture of boring overhanging,
and walls concave’ (Donovan and Jagt, 2002, p. 69).
The type series and related material of S. excavatus included pits in
numerous tests of H. striatoradiatus, some of which were densely infested;
most notably, Donovan and Jagt (2002, p. 69, Figure 2(c)) counted ‘at least
61 individual borings’ on NHMM JJ 699. Despite this wealth of material,
there was not any strong indication of site selectivity, although some pits
did occur in linear associations (Figure 7(b)) and most infested the test
supra-ambitally.
The two analogous specimens of H. striatoradiatus described above,
although relatively sparsely infested by S. excavatus, provide stronger indica-
tions that the settling behaviour of the infesting organism was site selective.
All encrustations on these tests are interpreted to have occurred after the
death of H. striatoradiatus (¼ episkeletozoans sensu Taylor and Wilson,
2002). In life, the tests would have been protected by both densely packed
spines and pedicellariae which would have prevented such varied infesta-
tions. A holasteroid such as H. striatoradiatus would not have been a deep
burrower; rather, it would have lived epifaunally, possibly furrowing
through the sediment at the surface. The occurrence of a partial marginal
(subanal) fasciole under the periproct points to such a mode of life. Thus,
the oral surface would have been in intimate contact with the sediment, pre-
cluding infestation of the oral surface of either specimen by oysters, bryo-
zoans, or spirorbid or serpulid annelids (Figure 8(b)). Further, with the
exception of S. excavatus, the tests preserve no evidence of the growth of
the echinoids being adversely affected by any boring or attachment. Thus,
318 Stephen K. Donovan

the only borings interpreted to be pre-mortem are those of S. excavatus, a


deduction which is in agreement with the observations on other specimens
infesting this echinoid species by Donovan and Jagt (2002, 2005, 2013b).
The four examples of S. excavatus on NHMM RZ 00162 (Figure 8(aee))
show a distinct pattern of infestation not otherwise noted on H. striatoradia-
tus. This may be because other specimens are either too densely (Figure 7(a)
and (b)) or too sparsely infested, or it may be that the present example
preserves a rare behaviour (but see Jagt, 2000, pl. 24, Figures 4 and 5).
Whatever, in this example at least, each S. excavatus is associated with a
different ambulacrum (Figure 8(a)). It is notable that the only ambulacrum
(anterior, III) on the apical surface that was not infested is also functionally
different from all the others (Smith, 1984, pp. 93e95; Jagt, 2000, p. 281).
That is, in NHMM RZ 00162, S. excavatus is only associated with ambu-
lacra bearing well-developed respiratory tube feet, although, as noted
above, infestations seem to be associated with reactions by the test. Most
notably, the posterior plate column in ambulacrum IV is deflected towards
the pit (best seen in Figure 8(e)). There is also a suspicion that in some ante-
rior plate columns of ambulacra (I, II, IV) the pores/pore pairs are weakly
developed, but, as these are invariably less prominent than those of the
posterior pore pairs, this may equally be a preservational artifact or an inac-
curate observation.
These observations are supported by the second specimen, NHMM MA
0234-1 (Figure 8(feh)). Although more densely infested by S. excavatus, this
test supports four pits in close association with three respiratory regions of
ambulacra (II  2, IV, V). These are on the upper half of the test and in po-
sitions analogous to those of NHMM RZ 00162 (compare Figure 8(aee)
and (feh)). But if a close association with a respiratory ambulacrum was
advantageous, it is unknown why seven other O. excavatus infested interam-
bulacra. Two pits are on the interradial suture (Figure 8(f), lower right;
Figure 8(h), far right) and all but two (Figure 8(g), centre of image) are
located towards the middle of an interambulacrum. Figure 7(a) shows an
analogous specimen to NHMM MA 0234-1. However, more densely bored
specimens (such as Figure 7(b)) distracted them from recognizing an associ-
ation between ambulacra and S. excavatus.
This association was presumably advantageous to the organisms forming
certain S. excavatus, which selected positions to settle on live H. striatoradiatus
with precision, either between columns of pore pairs (II, IV, V) or anterior
to the anterior column of pore pairs (I, II, V). This pattern begs a palaeoe-
cological explanation, but this must be speculative, particularly as S. excavatus
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 319

also successfully infested other areas of the test. All S. excavatus in Figure 8 are
above the mid-height of the test and, thus, were presumably above the sedi-
ment surface in this shallowly furrowing echinoid. None of the pits shows a
particularly strong association with the well-developed respiratory tube feet
situated posteriorly on each of the ambulacra of interest. We suggest, tenta-
tively, that this pattern of infestation may have placed the producers of
O. excavatus in advantageous positions to feed on plankton, anterior of the
tube feet on an echinoid moving slowly across the seafloor. The function
that the respiratory tube feet performed for the producer of O. excavatus is
uncertain, but may have been protective in part.
Reliability: Category 4. ‘For some types of behavioral evidence there is
an even larger degree of uncertainty about the maker of this evidence. This is
particularly true for many trace fossils’ (Boucot, 1990, p. 9).

4. DISCUSSION
The echinoderms are a promising group of fossils with which to inves-
tigate the nature of biotic reactions because the endoskeleton may produce
growth deformities of various types if infested in life. Other reviews of para-
sitism on echinoderms have tried to do too much in a restricted space (such
as Kowaleski and Nebelsick, 2003, pp. 294e295, on echinoids); I have
aimed to be expansive, but selective. There are some good examples that
I have chosen to ignore because they are well covered elsewhere (such as
various entries for parasitism on echinoids in Boucot, 1990). Five specific
examples are given in this essay, but none belong to the typical swollen
deformity for which the echinoderms are so well known (Figure 9) and
which have been determined as parasitic infestations by many authors
(Moodie, 1918). Although such structures are locally common, I suggest
that they are more difficult to interpret than some of the examples discussed
in detail herein. Certainly, some modern parasites produce identifiable galls
and cysts in echinoid spines, asteroid arms, ophiuroids discs, and crinoid pin-
nules and cirri (see, for example, Grygier, 1988). But similar pits and swell-
ings may be the result of embedment by an organism that is merely seeking
elevation above the substrate. It will gain an advantage, but is it parasitism
(Zapalski, 2011)?
Pits in the endoskeleton in crinoids and blastoids (such as Figure 9(a)) are
difficult to interpret. A crinoid column would provide very little nourish-
ment, so a bored fossil pluricolumnal should best be interpreted as an
320 Stephen K. Donovan

(a) (b)

Figure 9 (a) Pentagonocyclicus (col.) sp., RGM 544 426 (After Donovan and Lewis (2010),
Figure 2.), perforated by three pits. Wenlock Edge, Shropshire [NGR SO 588 980], England;
Much Wenlock Limestone Formation, Lower Silurian (Wenlock, Homerian). Note the
irregular outline of the pluricolumnal, due to swelling, and the large size and irregular
outline of the pit in the centre; the other two pits, including one at the top, are more
circular and assigned to Sedilichnus paraboloides (Bromley). Specimen whitened with
ammonium chloride for photography. (b) A strongly swollen pluricolumnal, RGM 791
727, in which the living crinoid grew over an encrusting coral. (After Donovan et al.
(2014b), Figure 4(g).) Salthill Quarry, Clitheroe, Lancashire [SD 7550 4265], England; lower
Viséan (¼ upper Chadian) Salthill Cap Beds, Bellman Limestone Member. Scale bars
represent 10 mm.

interaction with an unmineralized invertebrate seeking elevation above the


substrate and protection from the environment. (The latter could also be
provided by the dead parts of a dead crinoid.) An embedding organism
that elicited a reaction by the host to induce stereom growth and produce
a more robust, ankylosed structure would be further protected. This surely
is not what is commonly interpreted as parasitism. Reasonably, a boring/
embedding parasite sensu stricto would be more likely to infest a crinoid
through the more nutritious dorsal cup, tegmen (including the anal tube)
and arms, and would have perforated the endoskeleton (Baumiller, 1990)
rather than just embedding in it. Even though I have specialized in the
systematics and palaeobiology of the crinoid stem for 35 years, I trust it is
now obvious why I have included no bloated and bored crinoid stems in
the examples that I have given above. If you want to read further on such
ancient cysts and galls (reviewed in Conway Morris, 1981; Meyer and
Ausich, 1983, pp. 412e414), their literature is both broad and diffuse, and
may be overly speculative.
Instead, I have concentrated herein on examples of ectoparasites in
which there are obvious interactions with the crinoid theca by organisms
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 321

that apparently were exploiting part of the crinoid’s feeding capability sensu
lato and one example of a bored echinoid where site selectivity suggests an
interaction in life with the ambulacra. In two examples we actually know (or
at least have a very strong suspicion of) the producing organism. The inter-
action between a solitary stony coral and a crinoid theca (Section 3.1) in the
Mississippian is an unexpected one. There are strong indications that the
coral was interfering with the crinoid’s feeding, including its situation at
the level of the base of the crinoid arms and the unusual width of this inter-
ray, suggesting that growth of the crinoid was minimalizing the interference
by the parasite. What this interference was is the subject of informed spec-
ulation and was most probably as simple as robbing food from the adoral
grooves of the arms.
The coprophagic/parasitic association of platyceratid gastropods and
Palaeozoic crinoids is one of the best known organism/organism associations
in the fossil record, being based on many hundreds of specimens worldwide.
The circular scars on or below the anal pyramid in a posterior position on
Neoplatycrinus from Timor, Lacrimichnus? isp. (Section 3.4), provide data
on the nature of these associations additional to than commonly available,
that is, with the snail astride the anal pyramid on the apical surface.
It is apparent from the above examples that sparse infestations provide a
more cogent tale of organic interactions than dense infestations, which are
more likely to confuse the issue. The crinoid theca with multiple pits
through 360 (Section 3.3) preserves excellent evidence of biotic interac-
tions e the pit-forming organisms could not settle on those parts of the theca
that could be swept clean by the tube feet of the arms and the theca itself
shows a growth deformity (Figure 4) e but was any of it parasitism? To
use this example as an excuse for speculation, as in the title of this paper,
perhaps the multiple pits in this specimen represent more than one spatfall.
All pits are of similar size and spacing, but they could have formed in a pro-
gression of multiple reproductive seasons if the producing organism had a
much shorter life history than an adult crinoid. The pits of dead producers
may even have been recycled by juvenile settlers. But if the earliest pit(s)
was solitary or sparse and concentrated on one side of the cup, it may
have produced the sloping base as a growth reaction analogous to that in
Section 3.2 (Figures 2(def) and 3). Later, perhaps larger spatfalls were
excluded from making new pits where the originals were now inhabited
by larval ‘squatters’ because of biological constraints determined by the
size of the adult organism (¼ regular spacing of pits) and avoidance of sunken
plate sutures (Figure 4(b) and (c)). If we think in terms of a progressive
322 Stephen K. Donovan

invasion by pit-forming organisms, the single pit in Synbathocrinus represents


the start of an infestation (Figures 2(def) and 3) whereas the densely bored
Amphoracrinus is an end point (Figure 4). The earliest invaders can be selec-
tive and gain a feeding advantage, indicated by the uneven growth of the
cup, as well as elevation above the substrate; one or two spatfalls later, the
only advantage available is elevation. Thus, early spatfalls gain a feeding
advantage in an optimum position, which produces a growth deformity in
the crinoid theca and quite rightly may be regarded as a parasitic reaction.
Later spatfalls (this may all have happened over a few weeks or months)
did not produce an obvious growth deformity in the cup; was this parasitism
or not?
The Maastrichtian echinoid H. striatoradiatus (Leske) allows a similar anal-
ysis to be made on one species. When originally described by Donovan and
Jagt (2002), it was plainly apparent that the pits of S. excavatus were formed
while the host echinoid was alive by the evidence of the thickened ‘blisters’
of test calcite beneath them (Figure 7(eei)). The pits themselves are of more
complex morphology (Figure 7(c) and (d)) than those infesting either of the
pitted crinoids (Sections 3.2 and 3.3) and it can be more confidently
proposed that they were produced by members of a single species of unmin-
eralized invertebrate showing a particular boring behaviour. Individual echi-
noid tests are more or less well infested on the apical surface, particularly the
upper two thirds (Figure 7(a) and (b)), the ‘record’ being 61 individual pits,
including 2, unusually, on the oral surface. It was readily apparent that the
producers of S. excavatus were ‘hitching a ride’ on H. striatoradiatus, a hard
substrate that elevated them above the seafloor (in most examples) and
permitted them to feed by whatever means. Whether this was parasitism
and how so is difficult to determine. Perhaps comparing infested and nonin-
fested tests from a single horizon may indicate a size variance (compare with
Rollins and Brezinski, 1988), but no qualitative differences are obvious.
Certainly, when first described, there was no suspicion of a host/parasite
relationship between H. striatoradiatus and the producer of S. excavatus.
Contrast these early deliberations with the sparsely infested tests of
H. striatoradiatus described later by Donovan and Jagt (2013b; Figure 8 here-
in). Take away the ‘noise’ of dense infestation and a pattern appears, partic-
ularly in NHMM RZ 00162 (Figure 8(aee)). The four individual pits of
S. excavatus are each closely associated with an ambulacrum; only the ante-
rior ambulacrum, of different gross morphology and function, remained
without an associated pit (Figure 8(a)). Donovan and Jagt (2013b) deduced
that the close association of the pit-forming organism and echinoid
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 323

ambulacra provided some feeding or protective advantage to the borer;


either interpretation might be regarded as parasitic, particularly where the
ambulacral morphology was deformed as a response (Figure 8(cee)).

5. CONCLUSIONS
‘The fossil record contains many examples of the interactions that
occurred between organisms . evidence of interactions may be . cryptic
or difficult to interpret, none more so than the evidence of ancient para-
sitism’ (Littlewood and Donovan, 2003, p. 136). The evidence provided
by the endoskeletons of fossil echinoderms may be particularly rich, but
difficult to interpret. Too rarely is there a good answer to basic questions;
what was the parasite; what was being parasitized exactly and why infest
an innutrious structure like a crinoid stem when the real tasty guts of the an-
imal was in the cup? Some structures that give the superficial impression that
they were formed by parasites, such as deformities in the crinoid column, are
difficult to explain and probably do not represent parasitism per se. Most, at
least, probably represent dwelling traces, the pit or gall produced looking
like parasitism, but being little removed behaviourally from an encruster
attached to the outside of the test which would rarely, if ever, be thought
to be truly parasitic. Dense infestations may mask patterns that are apparent
in weakly infested specimens. And, in the absence of the producing organ-
ism in many examples, which is represented by its trace fossil, a sedimentary
structure, we are forced to speculate not just on whodunit, but also what it
was that they were doing.

ACKNOWLEDGEMENTS
I thank the editors, Tim Littlewood (The Natural History Museum, London) and Kenneth
De Baets (Friedrich-Alexander Universit€at, Erlangen), for inviting me to contribute to this
volume. I also thank my co-authors of the papers that form the backbone of this contribution
for their collaboration and collegial companionship, namely John Jagt (Natuurhistorisch
Museum Maastricht), David Lewis (the Natural History Museum, London), Paul Kabrna
(Barnoldswick, Lancashire) and Gary Webster (Washington State University, Pullman).

REFERENCES
Bandel, K., 1992. Platyceratidae from the Triassic St. Cassian Formation and the evolutionary
history of the Neritomorpha (Gastropoda). Pal€aontol. Z. 66, 231e240.
Bandel, K., 2007. Description and classification of Late Triassic Neritimorpha (Gastropoda,
Mollusca) from the St Cassian formation, Italian Alps. Bull. Geosci. 82, 215e274.
324 Stephen K. Donovan

Bandel, K., Frýda, J., 1999. Notes on the evolution and higher classification of the subclass
Neritimorpha (Gastropoda) with the description of some new taxa. Geol. Palaeontol.
33, 219e235.
Baumiller, T.K., 1990. Non-predatory drilling on Mississippian crinoids by platyceratid
gastropods. Palaeontology 33, 743e748.
Baumiller, T.K., 1993. Boreholes in Devonian blastoids and their implications for boring by
platyceratids. Lethaia 26, 41e47.
Baumiller, T.K., Gahn, F.J., 2003. Predation on crinoids. In: Kelley, P.H., Kowalewski, M.,
Hansen, T.A. (Eds.), Predator-prey Interactions in the Fossil Record. Kluwer Academic/
Plenum, New York, pp. 263e278.
Baumiller, T.K., Gahn, F.J., 2004. Testing predator-driven evolution with Paleozoic crinoid
arm regulation. Science 305, 1453e1455.
Baumiller, T.K., Gahn, F.J., Savill, J., 2004. New data and interpretations of the crinoid-
platyceratid interaction. In: Heinzeller, T., Nebelsick, J.H. (Eds.), Echinoderms. Taylor
and Francis, London, pp. 393e398.
Baumiller, T.K., Leighton, L.R., Thompson, D.L., 1999. Boreholes in Mississippian spirifer-
ide brachiopods and their implications for Paleozoic gastropod drilling. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 147, 283e289.
Boucot, A.J., 1990. Evolutionary Paleobiology of Behavior and Coevolution. Elsevier,
Amsterdam xxiii þ 725 pp.
Bowsher, A.L., 1955. Origin and adaptation in platyceratid gastropods. University of Kansas
Paleontological Contributions. Mollusca Article 5, 1e11.
Breimer, A., 1978. Recent crinoids. In: Moore, R.C., Teichert, C. (Eds.), Treatise on Inver-
tebrate Paleontiology, Part T, Echinodermata 2, vol. 1. Geological Society of America
and University of Kansas press, Boulder and Lawrence, pp. T9eT58.
Brett, C.E., 1978. Host-specific pit forming epizoans on Silurian crinoids. Lethaia 11,
217e232.
Brett, C.E., 1985. Tremichnus: a new ichnogenus of circular-parabolic pits in fossil
echinoderms. J. Paleontol. 59, 625e635.
Bromley, R.G., 1975. Comparative analysis of fossil and recent echinoid bioerosion. Palae-
ontology 18, 725e739.
Bromley, R.G., 1981. Concepts in ichnotaxonomy illustrated by small round holes in shells.
Acta Geol. Hisp. 16, 55e64.
Bromley, R.G., 2004. A stratigraphy of marine bioerosion. In: McIlroy, D. (Ed.), The Appli-
cation of Ichnology to Palaeoenvironmental and Stratigraphic Analysis, vol. 228.
Geological Society Special Publication, pp. 455e479.
Charlton, T.R., Barber, A.J., Harris, R.A., Barkham, S.T., Bird, P.R., Archbold, N.W.,
Morris, N.J., Nicoll, R.S., Owen, H.G., Owens, R.M., Sorauf, J.E., Taylor, P.D.,
Webster, G.D., Whittaker, J.E., 2002. The Permian of Timor: stratigraphy, palaeontol-
ogy and palaeogeography. J. Asian Earth Sci. 20, 719e774.
Conway Morris, S., 1981. Parasites and the fossil record. Parasitology 82, 489e509.
Dobson, A., Lafferty, K.D., Kuris, A.M., Hechinger, R.F., Jetz, W., 2008. Homage to Lin-
naeus: how many parasites? How many hosts? Proc. Natl. Acad. Sci. U.S.A. 105 (Suppl. 1),
11482e11489.
Donovan, S.K., 1991a. Site selectivity of a lower carboniferous boring organism infesting a
crinoid. Geol. J. 26, 1e5.
Donovan, S.K., 1991b. The taphonomy of echinoderms: calcareous multi-element skeletons
in the marine environment. In: Donovan, S.K. (Ed.), The Processes of Fossilization.
Belhaven Press, London, pp. 241e269.
Donovan, S.K., 2010. Cruziana and Rusophycus: trace fossils produced by trilobites. in some
cases? Lethaia 43, 283e284.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 325

Donovan, S.K., 2013. Curiouser and curiouser: more on reworked Echinocorys (Echinoidea;
Late Cretaceous) on the beaches of north Norfolk, eastern England. Swiss J. Palaeontol.
132, 1e4.
Donovan, S.K., 2014. Enigmatic branching structures within Upper Devonian crinoids,
north Devon, UK. Lethaia 47, 151e152.
Donovan, S.K., 2015. When is a fossil not a fossil? When it is a trace fossil. Lethaia 48, 145e146.
Donovan, S.K., Jagt, J.W.M., 2002. Oichnus Bromley borings in the irregular echinoid Hemi-
pneustes Agassiz from the type Maastrichtian (Upper Cretaceous, The Netherlands and
Belgium). Ichnos 9, 67e74.
Donovan, S.K., Jagt, J.W.M., 2005. An additional record of Oichnus excavatus Donovan & Jagt
from the Maastrichtian (Upper Cretaceous) of southern Limburg, The Netherlands. Scr.
Geol. 129, 147e150.
Donovan, S.K., Jagt, J.W.M., 2013a. Aspects of clavate borings in the type Maastrichtian
(Upper Cretaceous) of the Netherlands and Belgium. Neth. J. Geosci. 92, 133e143.
Donovan, S.K., Jagt, J.W.M., 2013b. Site selectivity of the pit Oichnus excavatus Donovan and
Jagt infesting Hemipneustes striatoradiatus (Leske) (Echinoidea) in the type Maastrichtian
(Upper Cretaceous, The Netherlands). Ichnos 20, 112e115.
Donovan, S.K., Jagt, J.W.M., Goffings, L., 2014a. Bored and burrowed: an unusual echi-
noid steinkern from the Type Maastrichtian (Upper Cretaceous, Belgium). Ichnos
21, 261e265.
Donovan, S.K., Jagt, J.W.M., Lewis, D.N., 2011a. Notes on some trace fossils and other para-
taxa from the Maastrichtian type area, southeast Netherlands and northeast Belgium. In:
Jagt, J.W.M., Jagt-Yazykova, E.A., Schins, W.J.H. (Eds.), A Tribute to the Late Felder
Brothers e Pioneers of Limburg Geology and Prehistoric Archaeology, Netherlands
Journal of Geosciences, vol. 90, pp. 99e109.
Donovan, S.K., Kabrna, P., Donovan, P.H., 2014b. Salthill Quarry: a resource being
revitalized. Deposits 40, 32e33.
Donovan, S.K., Lewis, D.N., 1999. An epibiont and the functional morphology of the col-
umn of a platycrinitid crinoid. Proc. Yorks. Geol. Soc. 53, 321e323.
Donovan, S.K., Lewis, D.N., 2010. Aspects of crinoid palaeontology, Much Wenlock
Limestone Formation, Wenlock Edge, Shropshire (Silurian). Proc. Yorks. Geol. Soc.
58, 9e14.
Donovan, S.K., Lewis, D.N., 2011. Strange taphonomy: Late Cretaceous Echinocorys (Echi-
noidea) as a hard substrate in a modern shallow marine environment. Swiss J. Palaeontol.
130, 43e51.
Donovan, S.K., Lewis, D.N., Crabb, P., 2003. Lower Carboniferous Echinoderms of North-
west England, vol. 1. Palaeontological Association Fold-Out Fossils, 12 pp.
Donovan, S.K., Lewis, D.N., Kabrna, P., 2005. An unusual crinoid-coral association from the
Lower Carboniferous of Clitheroe, Lancashire. Proc. Yorks. Geol. Soc. 55, 301e304.
Donovan, S.K., Lewis, D.N., Kabrna, P., 2006. A dense epizoobiontic infestation of a Lower
Carboniferous crinoid (Amphoracrinus gilbertsoni (Phillips)) by Oichnus paraboloides
Bromley. Ichnos 13, 43e45.
Donovan, S.K., Pickerill, R.K., 2002. Pattern versus process or informative versus uninfor-
mative ichnotaxonomy: reply to Todd and Palmer. Ichnos 9, 85e87.
Donovan, S.K., Pickerill, R.K., 2004. Traces of cassid snails predation upon the echinoids
from the Middle Miocene of Poland: comments on Ceranka and Z1otnik (2003). Acta
Palaeontol. Pol. 49, 483e484.
Donovan, S.K., Schmidt, D.A., 2001. Survival of crinoid stems following decapitation:
evidence from the Ordovician and palaeobiological implications. Lethaia 34, 263e270.
Donovan, S.K., Sutton, M.D., Sigwart, J.D., 2010. Crinoids for lunch? An unexpected biotic
interaction from the Upper Ordovician of Scotland. Geology 38, 935e938.
326 Stephen K. Donovan

Donovan, S.K., Sutton, M.D., Sigwart, J.D., 2011b. The last meal of the late Ordovician
mollusc ‘Helminthochiton’ thraivensis reed, 1911, from the Lady Burn Starfish Beds, south-
west Scotland. Geol. J. 46, 451e463.
Donovan, S.K., Webster, G.D., 2013. Platyceratid gastropod infestations of Neoplatycrinus
Wanner (Crinoidea) from the Permian of West Timor: speculations on thecal
modifications. Proc. Geol. Assoc. 124, 988e993.
Dunn, R.R., Harris, N.C., Colwell, R.K., Koh, L.P., Sodhi, N.S., 2009. The sixth mass
coextinction: are most endangered species parasites and mutualists? Proc. R. Soc.
B276, 3037e3045.
Eckert, J.D., 1988. The ichnogenus Tremichnus in the Lower Silurian of western New York.
Lethaia 21, 281e283.
Erickson, G.M., 2014. Breathing life into T. rex. Sci. Am. Spec. 23 (2), 38e45.
Fulton, F.T., 1975. The diet of the chiton Mopalia lignosa (Gould, 1846). Veliger 18, 38e41.
Gahn, F.J., Baumiller, T.K., 2003. Infestation of Middle Devonian (Givetian) camerate cri-
noids by platyceratid gastropods and its implications for the nature of their biotic
interaction. Lethaia 36, 71e82.
Gahn, F.J., Baumiller, T.K., 2006. Using platyceratid gastropod behaviour to test functional
morphology. Hist. Biol. 18, 397e404.
Grygier, M.J., 1988. Unusual and mainly cysticolous crustacean, molluscan, and myzostomidan
associated of echinoderms. In: Burke, R.D., Mladenov, P.V., Lambert, P., Parsley, R.L.
(Eds.), Echinoderm Biology: Proceedings of the Sixth International Echinoderm Confer-
ence, Victoria, British Columbia, 23e28 August, 1987. Balkema, Rotterdam, pp. 775e784.
Grygier, M.J., Nomura, K., 1998. Cysticolous Myzostomida, Notopharyngoides platypus from
Comanthina nobilis (Echinodermata: Crinoidea), at Kushimoto, Honshu, Japan. Species
Divers. 3, 17e24.
Hubbard, J.A.E.B., 1970. Sedimentological factors affecting the distribution and growth of
Viséan caninioid corals in north-west Ireland. Palaeontology 13, 191e209.
Jagt, J.W.M., 2000. Late Cretaceous-Early Palaeogene echinoderms and the K/T boundary
in the southeast Netherlands and northeast Belgium e Part 4: Echinoids. Scr. Geol. 121,
181e375.
Jagt, J.W.M., 2007. A late Cretaceous gastropod homing scar (possibly ichnogenus Lacrimich-
nus) from southern Limburg, the Netherlands. Scr. Geol. 134, 19e25.
Jangoux, M., 1987a. Diseases of Echinodermata. I. Agents microorganisms and protistans.
Dis. Aquat. Org. 2, 147e162.
Jangoux, M., 1987b. Diseases of Echinodermata. II. Agents metazoans (Mesozoa to Bryozoa).
Dis. Aquat. Org. 2, 205e234.
Jangoux, M., 1987c. Diseases of Echinodermata. III. Agents metazoans (Annelida to Pisces).
Dis. Aquat. Org. 3, 59e83.
Jangoux, M., 1987d. Diseases of Echinodermata. IV. Structural abnormalities and general
considerations on biotic diseases. Dis. Aquat. Org. 3, 221e229.
Jefferies, R.P.S., 1989. The arm structure and mode of feeding of the Triassic crinoid Encrinus
liliiformis. Palaeontology 32, 483e497.
Kabrna, P. (Ed.), 2011. Carboniferous Geology: Bowland Fells to Pendle Hill. Craven and
Pendle Geological Society, UK, 204 pp.
Kowalewski, M., Nebelsick, J.H., 2003. Predation on Recent and fossil echinoids. In:
Kelley, P.H., Kowalewski, M., Hansen, T.A. (Eds.), Predator-prey Interactions in the
Fossil Record. Kluwer Academic/Plenum, New York, pp. 279e302.
Lane, N.G., 1984. Predation and survival among inadunate crinoids. Paleobiology 10,
453e458.
Latyshev, N.A., Khardin, A.S., Kasyanov, S.P., Ivanova, M.B., 2004. A study on the feeding
ecology of chitons using analysis of gut contents and fatty acid markers. J. Molluscan
Stud. 70, 225e230.
A Prejudiced Review of Ancient Parasites and Their Host Echinoderms 327

Leske, N.G., 1778. Iacobi Theodori Klein naturalis dispositio echinodermatum, edita et
descriptionibus novisque inventis et synonymis auctorum et aucta a N. G. Leske. G.E.
Beer, Lipsiae xxii þ 278 pp.
Littlewood, D.T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
Macurda Jr., D.B., Meyer, D.L., 1974. Feeding posture of modern stalked crinoids. Nature
247, 394e396.
Meyer, D.L., Ausich, W.I., 1983. Biotic interactions among recent and among fossil crinoids.
In: Tevesz, M.J.S., McCall, P.L. (Eds.), Biotic Interactions in Recent and Fossil Benthic
Communities. Plenum, New York, pp. 377e427.
Miller, J.S., 1821. A Natural History of the Crinoidea or Lily-shaped Animals, with Observa-
tions on the Genera Asteria, Eurayle, Comatula and Marsupites. C. Frost, Bristol, 150 pp.
Mitchell, M., 2003. A Lateral Key for the Identification of the Commoner Lower Carbon-
iferous Coral Genera. Westmorland Geological Society and Manchester Geological
Association, Manchester, 13 pp.
Moodie, R.L., 1918. On the parasitism of Carboniferous crinoids. J. Parasitol. 4, 174e176.
Moore, R.C., Ubaghs, G., Breimer, A., Lane, N.G., 1978. Glossary of crinoid morphological
terms. In: Moore, R.C., Teichert, C. (Eds.), Treatise on Invertebrate Paleontology, Part
T, Echinodermata 2, vol. 1. Geological Society of America and University of Kansas
Press, Boulder and Lawrence. T229, T231, T233eT242.
M€ uller, A.H., 1977. Zur Ichnologie der subherzynen Oberkreide (Campan). Z. Geol. Wiss.
Berlin 5, 881e897.
M€ uller, K.J., Nogami, Y., Lenz, H., 1974. Phosphatische Ringe als Mikrofossilien im
Altpal€aozoikum. Palaeontographica A146, 79e99.
Nichols, D., 1994. Reproductive seasonality in the comatulid crinoid Antedon bifida
(Pennant) from the English Channel. Philos. Trans. R. Soc. Lond. B343, 113e134.
Osborn, H.F., 1905. Tyrannosaurus and other Cretaceous carnivorous dinosaurs. Bull. Am.
Mus. Nat. Hist. 21, 259e265.
Pennant, T., 1777. The British Zoology, fourth ed., vol. 4. B. White, Warrington and London.
Phillips, J., 1836. Illustrations of the Geology of Yorkshire. Part 2. The Mountain Limestone
District. John Murray, London xx þ 253 pp.
Pickerill, R.K., 1994. Nomenclature and taxonomy of invertebrate trace fossils. In:
Donovan, S.K. (Ed.), The Palaeobiology of Trace Fossils. John Wiley and Sons, Chiches-
ter, pp. 3e42.
Pickerill, R.K., Donovan, S.K., 1998. Ichnology of the Pliocene Bowden shell bed, southeast
Jamaica. In: Donovan, S.K. (Ed.), The Pliocene Bowden Shell Bed, Southeast Jamaica,
Contributions to Tertiary and Quaternary Geology, vol. 35, pp. 161e175.
Ramsbottom, W.H.C., 1961. The British Ordovician Crinoidea. Monogr. Palaeontogr. Soc.
114 (492), 1e37.
Reed, F.R.C., 1911. A new fossil from Girvan. Geol. Mag. 48, 113e114.
Rohde, K., 2005. Marine Parasitology. CABI Publishing, Oxford, 592 pp.
Rollins, H.B., Brezinski, D.K., 1988. Reinterpretation of crinoid-platyceratid interaction.
Lethaia 21, 207e217.
Santos, A., Mayoral, E., Mu~ niz, F., 2003. New trace fossils produced by etching molluscs
from the upper Neogene of the southwestern Iberian Peninsula. Acta Geol. Pol. 53,
181e188.
Santos, A., Mayoral, E., Mu~ niz, F., 2005. Bioerosion scars of acorn barnacles from the south-
western Iberian Peninsula, upper Neogene. Riv. Ital. Paleontol. Stratigr. 111, 181e189.
Sirenko, B.I., 2000. Feeding of Polyplacophora and its role in evolution of the class. In: Mol-
luscs: Problems of Taxonomy, Ecology and Phylogeny. Zoological Institute of the
Russian Academy of Sciences, St. Petersburg, pp. 129e134 (in Russian).
Smith, A.B., 1984. Echinoid Palaeobiology. George Allen and Unwin, London xii þ 191 pp.
328 Stephen K. Donovan

Smith, A.B., Kroh, A. (Eds.), 2011. The Echinoid Directory. World Wide Web Electronic
Publication. http://www.nhm.ac.uk/research-curation/projects/echinoid-directory
(accessed 21.08.14.).
Sutton, M.D., Sigwart, J.D., 2012. A chiton without a foot. Palaeontology 55, 401e411.
Taylor, P.D., Wilson, M.A., 2002. A new terminology for marine organisms inhabiting hard
substrates. Palaios 17, 522e525.
Taylor, P.D., Wilson, M.A., 2003. Palaeoecology and evolution of marine hard substrate
communities. Earth Sci. Rev. 62, 1e103.
Ubaghs, G., 1978. Skeletal morphology of fossil crinoids. In: Moore, R.C., Teichert, C. (Eds.),
Treatise on Invertebrate Paleontology, Part T, Echinodermata 2, vol. 1. Geological Society
of America and University of Kansas Press, Boulder and Lawrence, pp. T58eT216.
Vickery, W.L., Poulin, R., 1998. Parasite extinction and colonisation and the evolution of
parasite communities: a simulation study. Int. J. Parasitol. 28, 727e737.
Wanner, J., 1916. Die Permischen Echinodermen von Timor, I. Teil. Palaontologie von
Timor, vol. 11, 329 pp.
Wanner, J., 1937. Neue beitr€age zur kenntnis der Permischen echinodermen von Timor.
VIII-XIII. Palaeontographica IV (Suppl. 4), 212 pp. Abteilungen 2, Abschnitt.
Warén, A., Moolenbeek, R., 1989. A new eulimid gastropod, Trochostilifer eucidaricola, para-
sitic on the pencil urchin Eucidaris tribuloides from the southern Caribbean. Proc. Biol.
Soc. Wash. 102, 169e175.
Webster, G.D., 1998. Distortion in the stratigraphy and biostratigraphy of Timor, a historical
review with an analysis of the crinoid and blastoid faunas. Proc. R. Soc. Victoria 110,
45e72.
Webster, G.D., Donovan, S.K., 2012. Before the extinction e Permian platyceratid gastro-
pods attached to platycrinitid crinoids and an abnormal four-rayed Platycrinities s.s wachs-
muthi (Wanner) from West Timor. Palaeoworld 21, 153e159.
Welch, J.R., 1976. Phosphannulus on Paleozoic crinoid stems. J. Paleontol. 50, 218e225.
Werle, N.G., Frest, T.J., Mapes, R.H., 1984. The epizoan Phosphannulus on a Pennsylvanian
crinoid stem from Texas. J. Paleontol. 58, 1163e1166.
Wright, J., 1952. A monograph of the British Carboniferous Crinoidea. Volume 1, part 4.
Monogr. Palaeontogr. Soc. Lond. 106 (458), 104e148.
Wright, J., 1955. A monograph of the British Carboniferous Crinoidea. Volume 2, part 1.
Monogr. Palaeontogr. Soc. Lond. 108 (467) (for 1954), 191e254.
Zapalski, M.K., 2011. Is absence of proof a proof of absence? Comments on commensalism.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 302, 484e488.
Zonneveld, J.-P., Gingras, M.K., 2014. Sedilichnus, Oichnus, Fossichnus, and Tremichnus: ‘small
round holes in shells’ revisited. J. Paleontol. 88, 895e905.
CHAPTER EIGHT

Differentiating Parasitism and


Other Interactions in Fossilized
Colonial Organisms
Paul D. Taylor
Department of Earth Sciences, Natural History Museum, London, UK
E-mail: p.taylor@nhm.ac.uk

Contents
1. Introduction 330
2. Colonial Animals 331
3. Putative Parasites of Fossil Colonial Animals 333
3.1 Recognition of parasitism in fossils 333
3.2 Symbiotic intergrowths and bioclaustrations 334
3.2.1 Caunopores 335
3.2.2 Rugose corals and stromatoporoids 336
3.2.3 Chaetosalpinx and other bioclaustrations 336
3.2.4 Cornulitids and colonial hosts 338
3.2.5 Celleporaria and Culicia 338
3.2.6 Pyrgomatid barnacles 339
3.3 Galls 340
3.4 Borings 340
3.5 Supposed parasites of graptolites 341
4. Fossil Colonial Animals as Parasites 341
5. Discussion 342
Acknowledgements 343
References 343

Abstract
Colonial species occur in a wide range of aquatic invertebrates, some having excellent
fossil records, notably corals, bryozoans and graptolite hemichordates. In contrast to
unitary animals, colonial animals grow by adding repetitive modules known as zooids.
The ability of colonies to endure partial mortality and the typically plastic growth of
benthic colonial species facilitates the formation of macrosymbiotic associations,
some of which may be parasitic. However, as with unitary fossils, it is notoriously
difficult to identify whether the symbioses are parasitisms (þ/) or mutualisms
(þ/þ). Intergrowths between host colonies of stromatoporoid sponges, corals or bryo-
zoans, and skeletal or soft-bodied symbionts are particularly common in Ordoviciane
Devonian shallow-water deposits. Soft-bodied symbionts in such intergrowths are
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.05.002 All rights reserved. 329
330 Paul D. Taylor

represented by moulds in the host skeletons, a process of preservation termed


bioclaustration. As yet, however, there is a lack of convincing data showing that any
of these symbionts were parasites. By comparison with modern analogues, some fossil
galls provide more convincing examples of parasitism, and the destructive effects of
borings into the skeletons of benthic colonies also argue in favour of parasitism. Pelagic
graptoloid hemichordates from the Early Palaeozoic occasionally contain cysts or tubes
that have been attributed to parasites on the grounds that they would have adversely
affected the hydrodynamics of the floating colonies. Future studies should test for para-
sitism by comparing the sizes of colonies hosting symbionts with those lacking
symbionts.

1. INTRODUCTION
Coloniality has evolved on multiple occasions in invertebrates
(Blackstone and Jasker, 2003). All colonial animals are aquatic and most
are marine, including corals and bryozoans (see Hughes, 1989 for a good
introduction to colonial and other clonal animals). Colonial animals are
characterized by a modular construction, comprising repetitive clonal units
called zooids. New zooids are budded to bring about colony growth,
which is usually indeterminate and plastic: different colonies belonging
to the same species may grow for different lengths of time and in different
patterns, achieving a range of sizes and shapes.
Relatively little is known about the parasites of colonial animals living
today (e.g. Hill and Okamura, 2007; but see Hooper, 2005 for sponges,
and Boero and Bouillon, 2005 for cnidarians). These can range from unicel-
lular microparasites e some of which might be closely related with mutual-
istic endosymbionts (Okamoto and McFadden, 2008) e to multicellular
macroparasites including helminths (Aeby, 2003), molluscs (e.g. Lorenz,
2005) and various arthropods (e.g. Duffy, 2003; Boxshall, 2005; Grygier
and Hoeg, 2005; L€ utzen, 2005). In some cases, sponges and cnidarians can
be important parasites of colonial organisms themselves (e.g. Hooper,
2005; Boero and Bouillon, 2005). Some of the most-studied endoparasites
of extant colonial organisms are myxozoans parasitizing freshwater bryozoans
(Okamura et al., 2015), which have not yet been reported from the fossil
record. Myxozoans are now considered to be cnidarians and impact their
bryozoan hosts by reducing colony growth, causing periodic castration and
zooid gigantism (Hartikainen and Okamura, 2012; Hartikainen et al., 2013).
Even less is known about parasites of ancient animal colonies found in
the fossil record. Direct fossil evidence of microbial parasites in colonial
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 331

animals is rare, and although examples of potential macrosymbioses


involving colonial animals are reasonably common in the fossil record, it
is difficult to find conclusive evidence that these represent parasitisms rather
than mutualisms (see Darrell and Taylor, 1993). An added complication is
that the costs and benefits incurred by interacting species today may vary
in ecological time (Bronstein, 1994).
This paper introduces coloniality, reviews macrosymbioses involving
colonial animals, including possible parasitisms, and makes suggestions for
further research particularly aimed at differentiating between parasitic and
mutualistic symbioses.

2. COLONIAL ANIMALS
The zooids in animal colonies are clonal and primitively have iden-
tical morphologies. However, zooidal polymorphism occurs in some
species, with zooids showing functional and morphological specialization,
e.g. for sexual reproduction or defence. Zooids of colonial species are typi-
cally reduced in size relative to individuals of related solitary animals. Small
zooid size is usually paralleled by a reduction in complexity.
Animal colonies divide into two basic ecological groups e benthic
and pelagic e both represented in the fossil record. Benthic colonial
species are found among the following phyla: Cnidaria, Bryozoa,
Entoprocta, Urochordata and Hemichordata. Anthozoan cnidarians are
often colonial, including many scleractinian and rugose corals and all
tabulate corals, as are the majority of hydrozoan cnidarians. Bryozoans
are an exclusively colonial phylum, while a few species of entoprocts
(¼kamptozoans) are colonial. Many urochordates, such as ascidians, and
pterobranch hemichordates are colonial. In addition, key aspects of the
growth and life history of species belonging to Porifera are closer to colo-
nial than to unitary organisms (see Simpson, 1973; but cf. Ereskovskii,
2003), and for the purpose of this review sponges are included among colo-
nial animals. Benthic colonies are usually sessile, fixed permanently to a
substrate such as a rock, shell or seaweed, although a few have evolved
automobility and others move by virtue of being attached to mobile
animals. A significant proportion of benthic animal colonies have bio-
mineralized skeletons, typically of calcium carbonate, and as a result they
are common components of the fossil record, especially in shallow-water
marine deposits.
332 Paul D. Taylor

Pelagic colonies occur in Cnidaria, Urochordata and Hemichordata.


They include siphonophore hydrozoans among cnidarians, urochordate
salps and the graptoloid hemichordates (see Cooper et al., 2012; Maletz,
2014). These colonies either float passively in the water column, or propel
themselves through the water using self-generated currents. Pelagic colonies
are invariably unmineralized and are therefore poorly represented in the fos-
sil record, with the exception of the graptolites, a group of Palaeozoic hemi-
chordates with resistant collagen-based organic skeletons that become
progressively degraded to carbon during fossilization.
Almost all colonial animals are suspension feeders, consuming small
phyto- or zooplankton and sometimes bacteria. In the case of Cnidaria, these
are captured using the harpoon-like nematocysts or with mucous sheets and
threads, but in the other phyla self-generated currents propel entrained food
particles towards the mouths of the zooids. Symbiotic associations with pho-
tosymbionts, such as the zooxanthellae found in reef corals, may provide a
further source of nutrition.
Colony growth occurs through the asexual addition of new zooids, and
to a variable extent by extension of existing zooids. Multiplication of col-
onies is more often sexual, typically involving fertilization of eggs by sperm
from a different colony and dispersal of the resultant larvae. In some species,
however, alternative modes of colony multiplication exist such as breakage
of colonies into two or more fragments, or the formation of dispersive
asexual propagules.
Partial mortality is an important characteristic of animal colonies. This is
the death of some e but not all e of the zooids making up the colony. Par-
tial mortality occurs routinely when zooids in the colony progressively reach
the terminal stage of their ontogeny. As a result, the oldest zooids at the cen-
tres of colonies in some species constitute a dead ‘necromass’ surrounded by
or supporting a living zone of younger zooids. Physical damage or attacks
and infestation of zooids one or a few at a time by predators or parasites
may also result in partial mortality. These, along with other forms of stress,
sometimes prompt reactions from the colony that are manifested in the skel-
eton, including the sealing-off of zooids (e.g. Nothdurft and Webb, 2009).
The combination of inherent plasticity in growth and form, especially in
sessile benthic species (e.g. Wulff, 1985; McKinney et al., 1990), and the abil-
ity of colonial animals to sustain partial mortality, makes them prone to host-
ing macrosymbionts, including parasites. The shape of the colony can
potentially be modified and specific zooids ‘sacrificed’ in order to accommo-
date symbionts without great disadvantages to the colony as a whole, which is
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 333

the effective unit of natural selection (at least in colonies that are not frag-
mented). In the case of colonial species with biomineralized or resistant
organic skeletons, clear evidence of the presence of macrosymbionts may
be preserved in the fossils. Indeed, Tapanila (2008) stated that the majority
of marine symbioses in the Palaeozoic involved colonial animals. As discussed
below, however, ascertaining the nature of the interaction between the sym-
bionts (i.e. parasitism or mutualism) in fossil material is a major challenge.

3. PUTATIVE PARASITES OF FOSSIL COLONIAL


ANIMALS
3.1 Recognition of parasitism in fossils
As with most putative parasites in the fossil record, those involving
colonial animals are subject to interpretive caveats. It is rarely possible to
quantify the interactions between the symbionts: does one symbiont gain
and the other lose (þ/ parasitism), or do both gain (þ/þ mutualism)
from their association? (Zapalski (2011) has argued persuasively for the aban-
donment of a third category, commensalism (þ/0), while Tapanila (2008)
retained commensalism as a null hypothesis). Ideally, distinguishing between
these kinds of symbiosis demands the availability of relative data on either
reproductive fecundities, or population growth of the two symbionts in
and out of symbiosis. This kind of data is effectively unobtainable in the fossil
record. Two alternative lines of reasoning might be applied to distinguish
mutualism from parasitism in the fossil record. The first is inductive
reasoning: if an association is extant and its parasitic nature has been demon-
strated in modern material, then the most parsimonious interpretation is that
it was also parasitic in the geological past. The second is to look for changes
in the size, growth rate and/or morphology of hosts when they harbour
symbionts: if hosts with symbionts are smaller, grew more slowly (as evident
from closer spacing of growth bands) or are significantly malformed relative
to those without symbionts, then it is reasonable to infer a parasitic associa-
tion. Abnormal growth patterns in fossil colonial animals have also been sug-
gested as being due to parasite infestations (e.g. Bull, 1994) but testing such
hypotheses is difficult or impossible, especially in older fossils belonging to
groups with no close living relatives.
In their general review of supposed parasitisms in the fossil record,
Baumiller and Gahn (2002) evaluated published examples in terms of: (1) ev-
idence for long-term association; (2) benefit to the putative parasite; and
334 Paul D. Taylor

(3) detriment to the host. Among the 33 tabulated associations, they found
very few instances in which all of these criteria were satisfied, and none of
these involved colonial symbionts. Probably the best-supported example
of a fossil parasitism is the Ordovician to Permian symbiosis between crinoids
and nutrient-stealing platyceratid gastropods, for which a detrimental impact
on the host crinoids can be inferred from the fact that infested crinoids are
smaller than noninfested examples. No comparable reduction in host colony
size seems yet to have been demonstrated in any fossil symbioses involving
colonial species. Palmer and Wilson (1988) argued that any interaction pro-
moting a growth response of a symbiont may be energetically costly, and
therefore potentially parasitic, as is any infestation that eliminates zooids of
a colonial animal, but this cost may of course have been balanced or sur-
passed by unknown advantages of living with a symbiont.

3.2 Symbiotic intergrowths and bioclaustrations


Cylindrical- or corkcrew-shaped symbionts (Figure 1) can be found inter-
grown with several groups of Palaeozoic colonial animals that have massive
skeletons, especially stromatoporoid sponges, tabulate corals and bryozoans.
The symbiont tubes are generally oriented more or less perpendicularly to
the surface of the host colony. Sometimes the symbionts had mineralized
skeletons of their own which are fossilized, but in other instances they
were soft-bodied and are preserved only by virtue of becoming progressively
embedded in the host skeleton as it grew, a process of preservation termed
bioclaustration by Palmer and Wilson (1988). In a review of soft-bodied,
bioclaustrated ‘endosymbionts’, Tapanila (2005) noted that they first
appeared in the Ordovician, diversified in the Silurian, and peaked in the

Figure 1 Vertical section of a Palaeozoic stromatoporoid sponge or a coral showing the


morphologies of bioclaustrational trace fossils formed by symbiotic intergrowths,
which are potentially parasites of the host colony. After Tapanila (2005).
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 335

Middle Devonian. A similar pattern seems to pertain for other symbiotic in-
tergrowths too.

3.2.1 Caunopores
The best known intergrowths involve ‘caunopores’ (Figure 2(a)), which are
branching, tubular symbionts found embedded in the skeletons of stromatop-
oroid sponges (Stel and Stoep, 1982; Mistiaen, 1984; Kershaw, 1987; Young
and Noble, 1989; Darrell and Taylor, 1993; Zhen, 1996). Caunopores are
common in the Silurian and Early to Middle Devonian, particularly in reefal
settings. Although the tubular symbionts are usually assumed to be syringop-
oroid tabulate corals, some examples have a different skeletal microstructure
and are of smaller diameter than is typical for such corals. If caunopores
were indeed corals (or other cnidarians), then possible disadvantages to the
host stromatoporoids may have been offset by the protection afforded by
the stinging nematocysts of the corals. However, the net cost or benefit to
the host stromatoporoids is unclear, placing a question mark over whether
this symbiosis is a parasitic or mutualistic association.

(a) (b) (c)

Figure 2 Intergrowths between fossil colonial animals and embedded symbionts,


some possibly parasitic. (a) Caunopores (?tabulate corals) (circular voids) and a larger
septate rugose coral (upper right) visible in a horizontal thin section of a stromatopor-
oid sponge from the Silurian (Hemse Beds) of Gotland, Sweden; NHMUK PEI127D; scale
bar 2 mm. (b) Vertical section through a rugose coral embedded in a stromatoporoid;
Silurian, Gotland; NHMUK PEI132H; scale bar 5 mm. (c) Bryozoan Celleporaria palmata
showing three corallites of the scleractinian coral Culicia parasita embedded in the
colony surface; Pliocene (Coralline Crag Formation) of Suffolk, UK; NHM BZ3428; scale
bar 5 mm.
336 Paul D. Taylor

3.2.2 Rugose corals and stromatoporoids


Solitary (Figure 2(b)) and occasionally ramose colonial rugose corals can also
be found embedded within stromatoporoids in Silurian and Devonian de-
posits (Kershaw, 1987; Stel and Stoep, 1982; Soto and Méndez-Bedia,
1985; Nestor et al., 2010; Vinn and M~ otus, 2014a). Individual stromatopor-
oids sometimes host both rugose corals and caunopores (Figure 2(a)), but it is
unclear whether the rugose coral symbionts were parasites of the host stro-
matoporoids. Vinn and M~ otus (2014a) postulated benefits to the corals in
terms of being able to feed at a higher tier in the water column and having
a more stable substrate than usual, while the stromatoporoid hosts may have
incurred advantages in being protected against predation by the nematocysts
of the corals. However, the latter would have been balanced by loss of some
of their feeding surface area.

3.2.3 Chaetosalpinx and other bioclaustrations


Palaeozoic tabulate corals frequently hosted vermiform symbionts (Tapanila,
2002, 2004, 2005; M~ otus and Vinn, 2009; Zapalski, 2007). The most com-
mon of these is the ichnogenus Chaetosalpinx. Zapalski’s (2009) study of
Chaetosalpinx infesting the Devonian tabulate Favosites showed this symbiont
to be absent during the early growth of the coral, subsequently appearing
and increasing in absolute numbers of individuals and numbers per zooid
as the colony grew. According to Zapalski (2007), perforation of the
skeleton (and soft tissues) of the host and modification of its morphology
are evidence for parasitism.
Stromatoporoids from the Devonian (Givetian) of France described by
Zapalski and Hubert (2011) contain helical tubes that they identified as
Torquaysalpinx. According to these authors, Torquaysalpinx biomineralized
its own calcareous walls, delineating tubes on average 0.33 mm in diameter.
Roughly 10% of studied stromatoporoids contained Torquaysalpinx tubes.
This symbiosis was interpreted as the first evidence of parasitism in a Palae-
ozoic sponge. The reason for supposing it to be a parasitic association was the
downward deflection of the growth lamellae of the stromatoporoid hosts
around the symbionts, suggesting a localized decrease in growth rate.
Caunopores and various other bioclaustrated symbionts commonly produce
similar deflections in the laminae of their hosts (see Tapanila, 2005, Figure 1).
The problem with using this kind of evidence for parasitism is that a local-
ized decrease in relative growth rate has no bearing on the overall absolute
growth rate of the host stromatoporoid, and therefore should not be used to
infer a negative impact on the host.
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 337

Both Chaetosalpinx and the spiral bioclaustration Helicosalpinx infest stro-


matoporoids in a Late Silurian biostrome in Estonia (Vinn and M~ otus,
2014b). Almost 78% of stromatoporoids from this biostrome contain one
or the other of these symbionts, with Chaetosalpinx being commoner than
Helicosalpinx and occurring in three host stromatoporoid species compared
with only one host species for Helicosalpinx.
Examples of the trepostome bryozoan Mesotrypa from the Middle Ordo-
vician (Darriwilian) of Estonia were recently described containing tubes of
0.65e0.85 mm in diameter (Vinn et al., 2014). The tubes, which have no
mineralized skeletons of their own, were named as a new ichnogenus and
ichnospecies, Anoigmaichnus odinsholmensis. They were interpreted as prob-
able annelids that initially bored into their hosts before becoming bio-
claustrated as the host continued upward growth. Interpretation of
Anoigmaichnus as a likely parasite of Mesotrypa was based on the fact that its
recruitment may have resulted in the death of a few zooids of the bryozoan
as well as stimulating ‘excessive secretion of zooecial walls around its body’
(p. 131). A very similar tubular bioclaustration has been reported in speci-
mens of the cystoporate bryozoan Stellatoides from the Devonian of
Germany (Ernst et al., 2014). Named as Chaeotosalpinx tapanilai, this
symbiont also lacks a skeleton of its own and has tubes averaging 0.7 mm
in diameter. The trace fossil resembles tubes made by some modern bryo-
zoan around spionid polychaete worms and sometimes tanaidacean
crustaceans.
Trepostome bryozoans from the Upper Ordovician of Cincinnati, USA
may contain a bioclaustration named Catellocaula vallata by Palmer and
Wilson (1988). This trace fossil, consisting of pits linked by narrower tun-
nels, was formed by the partial overgrowth of a soft-bodied organism that
settled on the surface of the living bryozoan colonies. Regarded as a parasite,
Catellocaula was thought by Palmer and Wilson (1988) to have been made by
colonial hydroid cnidarians or alternatively colonial ascidian tunicates. It dif-
fers from the intergrowths described above in that the Catellocaula-producing
organism grew horizontally across the surface of the host rather than verti-
cally parallel to the growth of the host.
Like Catellocaula, the bioclaustration Caupokeras was described from fossil
bryozoans, in this case Devonian fenestrates in which it forms a modular
network of tubes each expanding distally in diameter from about 0.15 to
0.50 mm (McKinney, 2009; Suarez Andrés, 2014). This trace fossil was
compared with structures produced by modern cheilostome bryozoans
around the stolons of symbiotic zancleid hydroid colonies (see Boero
338 Paul D. Taylor

et al., 2000). In some instances, association with hydroids possibly endows


cheilostome bryozoans with superior ability to compete for substrate space
(Ristedt and Schuhmacher, 1985). However, after considering possible
benefits and costs to both partners in the fossil symbiosis, McKinney
(2009) concluded that it was not possible to resolve whether the symbiosis
was parasitic or mutualistic.
An unnamed symbiont bioclaustration takes the form of funnel-like
tubes found in encrusting cyclostome bryozoans from the mid-Cretaceous
of Europe (Taylor and Voigt, 2006). The tubes, which can be numerous
within individual host colonies, have been found in four unrelated species.
They may occupy a significant proportion of the surface area of the host
bryozoan colonies, suggesting a negative impact on the host and a possible
parasitic relationship.

3.2.4 Cornulitids and colonial hosts


Cornulitid tubeworms, which are of unknown affinity but may have
resembled phoronid lophophorates with a mineralized skeleton, have
been described embedded within both Silurian stromatoporoids (Vinn
and Wilson, 2010) and Ordovician heliolitid tabulate corals (Vinn and
M~otus, 2008, 2012). These tubeworms evidently fouled the living surfaces
of their hosts and were compelled to grow upwards to keep pace with the
continued growth of the host.

3.2.5 Celleporaria and Culicia


Although differing in both age and the taxonomic identities of the two sym-
bionts, there are parallels between the intergrowth symbioses described
above and that between the cheilostome bryozoan Celleporaria palmata host-
ing the scleractinian coral Culicia parasita (Figure 2(c)) in the Neogene of
Europe (Cadée and McKinney, 1994; Chaix and Cahuzac, 2005). The
bryozoan grows as ramose colonies with thick branches onto which small
individuals of Cuclicia settled. Each coral grew in length in concert with
thickening of the bryozoan branches through frontal budding, becoming
cylindrical and progressively more deeply embedded within the branches
of the host bryozoan. Whereas C. parasita may be an obligate symbiont,
C. palmata is not. While Cadée and McKinney (1994) quite reasonably
believed the symbiosis to be a mutualism based on a range of conceivable
benefits to each of the two partners, a less likely scenario is that Culicia, as
the species epithet parasita suggests, was actually a parasite of Celleporaria.
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 339

3.2.6 Pyrgomatid barnacles


Pyrgomatids are a family of balanomorph barnacles that live today as symbi-
onts of scleractinian and hydrozoan corals, and also sometimes of sponges.
This speciose group is characterized by a cup-shaped to tubular morphology
and vertical growth that allows deep embedment in the skeletons of living
coral colonies. Recruitment onto their hosts begins when the barnacle larvae
penetrate the tissue of the host coral, apparently using the spearhead-shaped
third antennular segment (Brickner and Høeg, 2010).
Anderson (1992, p. 337) remarked: “Much remains to be done before
the physiological interactions between coral-inhabiting balanomorphs and
their hosts are adequately understood”. Nevertheless, he speculated on the
possibility of nutrient uptake by some pyrgomatids from their coral hosts.
Occupation of space that would otherwise have been occupied by coral
polyps is another characteristic of pyrgomatids that should impact negatively
on the host. In addition, the cirral teeth of the barnacles are employed to rasp
away coral tissue and prevent overgrowth (Anderson, 1992). One tribe of
pyrgomatids e Hoekiini e feed exclusively on tissues of the coral
Hydnophora and can be reasonably regarded as true parasites (Ross and
Newman, 1995).
A recent molecular phylogenetic study has suggested increased levels of
host specificity during the evolution of pyrgomatids, with basal clades con-
taining more generalist species than derived clades (Tsang et al., 2014).
Another molecular phylogeny (Malay and Michonneau, 2014) found the
parasitic hoekiines to be of relatively modern origin. Taken together, there-
fore, these two studies point to increasing levels of host specificity and a shift
from a relatively benign symbiosis to a more overtly parasitic association, at
least in some clades.
Pyrgomatids have a fossil record extending back to the Oligocene but
seem not to have become common until the Miocene (Ross and Newman,
2000). Unfortunately, there are no known fossils of the parasitic hoekiines,
which have fragile plates that may not be easily fossilized. While pyrgomatids
are normally preserved as body fossils, loss of the plates can leave a deep
funnel-shaped or cylindrical hole in the coral host. This was given the trace
fossil name Imbutichnus by Santos et al. (2012) and compared to other
bioclaustrations (see above).
In a twist to the pyrgomatid/coral symbiosis, Tilbrook (1997) described
Pliocene examples of the pyrgomatid Megatrema anglicum fouling corallites
of Culicia parasita (as Cryptangia woodii) embedded within colonies of the
bryozoan Celleporaria palmata (see above). This three-component system is
340 Paul D. Taylor

particularly interesting for two reasons. Firstly, the presence of the coral
brings the pyrgomatids into symbiosis with bryozoans with which they
are not otherwise associated. Secondly, the barnacles smother the individual
corals on which they settle, prompting lateral budding and quite possibly
entailing a significant cost, suggesting that the pyrgomatids in this instance
were parasites of the corals.

3.3 Galls
Diverse modern parasites induce their hosts to form galls (e.g. Rouse, 2005),
sometimes in colonial hosts. For example, pycnogonid arthropods may
inhabit galls in gorgonian and scleractinian hosts (Staples, 2005), corallio-
philid gastropods can form galls in various cnidarian groups (Lorenz,
2005), and gall-inducing copepods are known to occur in stylasterine
hydrozoans (Zibrowius, 1981). Coralliophilids have a fossil record dating
back to the Eocene, with the oldest example inhabiting galls apparently
being a species of Coralliophila found in the coral Cladocora from the Early
Oligocene of SW France (Lozouet and Renard, 1998). Voigt (1959) intro-
duced the genus Endosacculus for galls in Cretaceous octocorals considered to
have been induced by the presence of ascothoracican cirrepedes (see also
Voigt, 1967).

3.4 Borings
A plethora of macroboring organisms are capable of penetrating into the
hard skeletons of colonial animals (see Bromley, 2004; Wilson, 2007).
Benthic colonies with calcareous skeletons, particularly corals (e.g. Wilson
et al., 2014), have been targets for diverse macroboring organisms (see
Wilson, 2007) since the Cambrian. Glynn (1997) reviewed bioerosion of
coral reefs, in which sponges, annelids, crustaceans, sipunculans and molluscs
are the dominant macroendoliths. While the great majority of these animals
bioerode dead coral skeletons, some attack living tissues and may be agents
of partial mortality and therefore are potentially parasites of the host corals.
When dense or large relative to the host colony, macroborers may weaken
the skeleton, resulting eventually in breakage, detachment from the substrate
and colony mortality (see Highsmith, 1981). However, the impact of
macroborers on still-living colonial hosts is less clear and is potentially com-
plex. While bored colonies may need to divert resources to strengthening
and be otherwise disadvantaged, pointing to a parasitic relationship,
Kleemann (1994) made the interesting suggestion for bivalves boring into
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 341

scleractinian corals that after death the decaying tissues of the bivalve could
provide an extra source of nutrition for the coral host.
Microendoliths are also commonly found in the calcareous skeletons of
benthic fossil colonies. For example, Kolodziej et al. (2012) described pre-
mortem microborings into scleractinian corals of Early Cretaceous age.
These include traces attributed to autotrophic chlorophyte algae as well as
those of probable fungal origin. Like the macroborings described above,
both of these groups of microendoliths are capable of weakening the skele-
tons of the host corals. By analogy with extant analogues (see Golubic et al.,
2005), the fungal microendoliths may additionally have attacked the coral
polyps, raising the possibility of a parasitic relationship.

3.5 Supposed parasites of graptolites


Some extant groups of parasites specifically target zooplanktonic colonial
organisms such as hyperiid amphipods (L€ utzen, 2005). In this context, it is
interesting to mention that a large number of papers describe supposed
examples of parasitism in pelagic graptoloid graptolites, as reviewed by Bates
and Loydell (2000). These involve outgrowths of the collagenous skeleton
(periderm) in the form of blisters or cysts, or less often, small tubes. Both
types of outgrowth can be inferred to have had a detrimental effect on
the hydrodynamics of the floating graptoloid colony (Underwood, 1993)
and also possibly on its buoyancy. Unfortunately, the taxonomic affiliation
of the culprits remains unknown.
Benthic tuboid and dendroid graptolites sometimes contain larger
diameter tubes e originally termed tubothecae e that were described by
Kozlowski (1970) and inferred to be tubes formed by commensal annelids.
However, Urbanek and Mierzejewski (1982) used SEM to restudy tubothe-
cae and found them to be secretions of the graptolites themselves but around
symbionts.

4. FOSSIL COLONIAL ANIMALS AS PARASITES


Benthic colonies often foul or bore into living hosts (e.g. Taylor,
1994; Key et al., 1999; Taylor and Wilson, 2003; Bromley, 2004; Puce
et al., 2005). Both at the present day and in the fossil record, however,
the nature of the interactions between these colonies and their living sub-
strates hosts are poorly understood (e.g. David et al., 2009): few can be
recognized with certainty as parasites. An exception may be some recent
342 Paul D. Taylor

clionid sponges that are capable of undermining the skeleton beneath the
polyps of living corals as they compete for space with their hosts (L opez-
Victoria et al., 2006). There is a rich fossil record of sponge borings (ichno-
genus Entobia) into corals and other biotic substrates (see Bromley, 2004),
which warrant further study in this context.
Berkowski and Zapalski (2014) recently described examples of the tabu-
late coral Hamarilopora minima that encrusted the stems of living crinoids in
the Early Devonian of Morocco. Loss of stem flexibility as a result of encrus-
tation would have impacted the host crinoid negatively, suggesting that the
association was parasitic. Similar associations have been reported involving a
wide range of epibionts, both colonial and unitary, on crinoid stems else-
where in the fossil record (see references in Berkowski and Zapalski, 2014).

5. DISCUSSION
Benthic colonial invertebrates today are hosts for numerous microen-
dosymbionts. These include multifarious bacteria in sponges (Vogel, 2008),
the dinoflagellate Symbiodinium (zooxanthellae) in reef-building corals (e.g.
Baker, 2003), and the bryostatin-producing bacteria of the bryozoan Bugula
neritina (e.g. Davidson et al., 2001; Linneman et al., 2014). Direct fossil
evidence for such microsymbionts is scant and determining which are para-
sitic can be difficult. Exceptions are the boring microsymbionts mentioned
above, which leave traces (ichnofossils) in the skeletons of their hosts.
Whereas microsymbionts are typically cryptic and difficult to study in
fossil animal colonies, macrosymbionts offer opportunities for future
research, especially when both partners have fossilizable mineralized skele-
tons that are intergrown. In cases where the symbiosis is not obligatory, it
should be possible to make a comparative study of colonial hosts with and
without symbionts. The comparative sizes of hosts could be compared
and a parasitic association inferred if hosts with symbionts were smaller
than those without. In addition, relative growth rates in hosts with and
without symbionts could be compared in cases where the hosts have clear
morphological or isotopic growth banding, again permitting inference of
parasitisms if the presence of symbionts lowered the growth rate of the
host colonies. Given the abundance of symbiotic intergrowths in fossil ani-
mal colonies in some geological sections (e.g. Silurian of Gotland), there is
clearly scope for studying the temporal dynamics of the associations if a suf-
ficient amount of material can be collected for rigorous analysis.
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 343

A greater understanding of parasitisms involving colonial hosts at the


present day may reveal new opportunities for studying ancient macrosym-
bioses. This will be possible if studies on modern parasitisms pay particular
attention to their effects on the skeletal characters that are routinely fossil-
ized. In addition, isotopic signatures of symbiosis (cf. Dreier et al., 2012)
offer opportunities for identifying and studying the geological histories of
ancient symbioses in animal colonies.

ACKNOWLEDGEMENTS
The final version of this paper benefitted from the comments of the reviewers e Beth
Okamura, Leif Tapanila, Olev Vinn and Miko1aj Zapalski e to whom I am extremely
grateful.

REFERENCES
Aeby, G.S., 2003. Corals in the genus Porites are susceptible to infection by a larval trematode.
Coral Reefs 22, 216.
Anderson, D.T., 1992. Structure, function and phylogeny of coral-inhabiting barnacles
(Cirripedia, Balanoidea). Zool. J. Linn. Soc. 106, 277e339.
Baker, A.C., 2003. Flexibility and specificity in coral-algal symbiosis: diversity, ecology, and
biogeography. Ann. Rev. Ecol. Evol. Syst. 34, 661e689.
Bates, D.E.B., Loydell, D.K., 2000. Parasitism on graptoloid graptolites. Palaeontology 43,
1143e1151.
Baumiller, T.K., Gahn, F.J., 2002. Fossil record of parasitism on marine invertebrates
with special emphasis on the platyceratid-crinoid interaction. Paleontol. Soc. Pap. 8,
195e209.
Berkowski, B., Zapalski, M.K., 2014. Unusual tabulate-crinoid biocoenosis from the Lower
Devonian of Morocco. Lethaia 47, 176e186.
Blackstone, N.W., Jasker, B.D., 2003. Phylogenetic considerations of clonality, coloniality,
and mode of germline development in animals. J. Exp. Zool. Mol. Dev. Evol. 297B,
35e47.
Boero, F., Bouillon, J., 2005. Cnidaria and Ctenophora (cnidarians and comb jellies). In:
Rohde, K. (Ed.), Marine Parasitology. CSIRO Publishing, Collingwood, pp. 177e182.
Boero, F., Bouillon, J., Gravili, C., 2000. A survey of Zanclea, Halocoryne and Zanclella (Cni-
daria, Hydrozoa, Anthomedusae, Zancleidae) with description of new species. Ital. J.
Zool. 7, 93e124.
Boxshall, G., 2005. Copepoda (copepods). In: Rohde, K. (Ed.), Marine Parasitology.
CSIRO Publishing, Collingwood, pp. 123e138.
Brickner, I., Høeg, J.T., 2010. Antennular specialization in cyprids of coral-associated
barnacles. J. Exp. Mar. Biol. Ecol. 392, 115e124.
Bromley, R.G., 2004. A stratigraphy of marine bioerosion. Spec. Pub. Geol. Soc. Lond. 228,
455e479.
Bronstein, J.L., 1994. Conditional outcomes in mutualistic interactions. Trends Ecol. Evol. 9,
214e217.
Bull, E.E., 1994. Implications of normal and abnormal growth structures in a Scottish Silurian
dendroid graptolite. Palaeontology 39, 219e240.
Cadée, G.C., McKinney, F.K., 1994. A coral-bryozoan association from the Neogene of
northwestern Europe. Lethaia 27, 59e66.
344 Paul D. Taylor

Chaix, C., Cahuzac, B., 2005. Le genre Culicia (Scléractiniaire): systématique, écologie et
biogéographie au Cénozoïque. Eclogae Geol. Helv. 98, 169e187.
Cooper, R.A., Rigby, S., Loydell, D.K., Bates, D.E.B., 2012. Palaeoecology of
Graptoloidea. Earth Sci. Rev. 112, 23e41.
Darrell, J.G., Taylor, P.D., 1993. Macrosymbiosis in corals: a review of fossil and potentially
fossilizable examples. Cour. Forschungsinst. Senckenberg 164, 185e198.
David, B., Stock, S.R., De Carlo, F., Hétérier, V., De Ridder, C., 2009. Microstructures of
Antarctic cidaroid spines: diversity of shapes and ectosymbiont attachments. Mar. Biol.
156, 1559e1572.
Davidson, S.K., Allen, S.W., Lim, G.E., Anderson, C.M., Haygood, M.G., 2001. Evidence
for the biosynthesis of bryostatins by the bacterial symbiont ‘‘Candidatus Endobugula
sertula’’ of the bryozoan Bugula neritina. Appl. Environ. Microbiol. 67, 4531e4537.
Dreier, A., Stannek, L., Blumenberg, M., Taviani, M., Sigovini, M., Wrede, C., Thiel, V.,
Hoppert, M., 2012. The fingerprint of chemosymbiosis: origin and preservation of
isotopic biosignatures in the nonseep bivalve Loripes lacteus compared with Venerupis
aurea. FEMS Microbiol. Ecol. 81, 480e493.
Duffy, J.E., 2003. The ecology and evolution of eusociality in sponge-dwelling shrimp. In:
Kikuchi, T., Higashi, S., Azuma, N. (Eds.), Genes, Behaviors, and Evolution in Social
Insects. University of Hokkaido Press, Sapporo, pp. 217e252.
Ereskovskii, A.V., 2003. Problems of coloniality, modularity, and individuality in sponges
and special features of their morphogeneses [sic] during growth and asexual
reproduction. Russ. J. Mar. Biol. 29 (Suppl. 1), S46eS56.
Ernst, A., Taylor, P.D., Bohatý, J., 2014. A new Middle Devonian cystoporate bryozoan
from Germany containing a new symbiont bioclaustration. Acta Palaeontol. Pol. 59,
173e183.
Glynn, P.W., 1997. Bioerosion and coral reef growth: a dynamic balance. In: Birkeland, C.
(Ed.), Life and Death of Coral Reefs. Chapman & Hall, New York, pp. 69e98.
Golubic, S., Radtke, G., Le Campion-Alsumard, T., 2005. Endolithic fungi in marine
ecosystems. Trends Microbiol. 13, 229e235.
Grygier, M.J., Høeg, J.T., 2005. Ascothoracida (acothoracids). In: Rohde, K. (Ed.), Marine
Parasitology. CSIRO Publishing, Collingwood, pp. 149e154.
Hartikainen, H., Okamura, B., 2012. Castrating parasites and colonial hosts. Parasitology
139, 547e556.
Hartikainen, H., Fontes, I., Okamura, B., 2013. Parasitism and phenotypic change in colonial
hosts. Parasitology 140, 1403e1412.
Highsmith, R.C., 1981. Coral bioerosion: damage relative to skeletal strength. Am. Nat. 117,
193e198.
Hill, S.L.L., Okamura, B., 2007. Endoparasitism in colonial hosts: patterns and processes.
Parasitology 134, 841e852.
Hooper, J.N.A., 2005. Porifera (sponges). In: Rohde, K. (Ed.), Marine Parasitology. CSIRO
Publishing, Collingwood, pp. 174e177.
Hughes, R.N., 1989. A Functional Biology of Clonal Animals. Chapman & Hall, London,
331 pp.
Kershaw, S., 1987. Stromatoporoid-coral intergrowths in a Silurian biostrome. Lethaia 20,
371e380.
Key Jr., M.M., Winston, J.E., Volpe, J.W., Jeffries, W.B., Voris, H.K., 1999. Bryozoan
fouling of the blue crab Callinectes sapidus at Beaufort, North Carolina. Bull. Mar. Sci.
64, 513e533.
Kleemann, K., 1994. Associations of corals and boring bivalves since the Late Cretaceous.
Facies 31, 131e140.
Ko1odziej, B., Golubic, S., Radtke, G., Tribollet, A., 2012. Early Cretaceous record of
microboring organisms in skeletons of growing corals. Lethaia 45, 34e45.
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 345

Kozlowski, R., 1970. Tubotheca e a peculiar morphological element in some graptolites.


Acta Palaeontol. Pol. 15, 393e410.
Linneman, J., Paulus, D., Lim-Fong, G., Lopanik, N.B., 2014. Latitudinal variation of a
defensive symbiosis in the Bugula neritina (Bryozoa) sibling species complex. PLoS One
e108783.
L
opez-Victoria, M., Zea, S., Weil, E., 2006. Competition for space between encrusting
excavating Caribbean sponges and other coral reef organisms. Mar. Ecol. Prog. Ser.
312, 113e121.
Lorenz, F., 2005. Mollusca (molluscs). In: Rohde, K. (Ed.), Marine Parasitology. CSIRO
Publishing, Collingwood, pp. 240e245.
Lozouet, P., Renard, P., 1998. Les Coralliophilidae, Gastropoda de l’Oligocene et du
Miocene inférieur d’Aquitaine (Sud-Ouest de la France): Systématique et coraux
h^otes. Geobios 31, 171e185.
L€
utzen, J., 2005. Amphipoda (amphipods). In: Rohde, K. (Ed.), Marine Parasitology.
CSIRO Publishing, Collingwood, pp. 165e169.
McKinney, F.K., 2009. Bryozoan-hydroid symbiosis and a new ichnogenus, Caupokeras.
Ichnos 16, 193e201.
McKinney, F.K., Broadhead, T.W., Gibson, M.A., 1990. Coral-bryozoan mutualism:
structural; innovation and greater resource exploitation. Science 248, 466e468.
Malay, C.D.M., Michonneau, F., 2014. Phylogenetics and morphological evolution of coral-
dwelling barnacles (Balanomorpha: Pyrgomatidae). Biol. J. Linn. Soc. 113, 162e179.
Maletz, J., 2014. Graptolite reconstructions and interpretations. Pal€aontol. Z. http://
dx.doi.org/10.1007/s12542-014-0234-4.
Mistaen, B., 1984. Comments on caunopore tubes: stratigraphic distribution and
microstructure. Palaeontogr. Am. 54, 501e508.
M~otus, M.A., Vinn, O., 2009. The worm endosymbionts in tabulate corals from the Silurian
of Podolia, Ukraine. Est. J. Earth Sci. 58, 185e192.
Nestor, H., Copper, P., Stock, C.W., 2010. Late Ordovician and Early Silurian Stromatop-
oroid Sponges from Anticosti Island, Eastern Canada: Crossing the O/S Mass Extinction.
National Research Council of Canada, Ottawa, 173 pp.
Nothdurft, L.D., Webb, G.E., 2009. Clypeotheca, a new skeletal structure in scleractinian
corals: a potential stress indicator. Coral Reefs 28, 143e153.
Okamoto, N., McFadden, G.I., 2008. The mother of all parasites. Future Microbiol. 3,
391e395.
Okamura, B., Gruhl, A., Bartholomew, J.L., 2015. Myxozoan Evolution, Ecology and
Development. Springer, Dordrecht.
Palmer, T.J., Wilson, M.A., 1988. Parasitism of Ordovician bryozoans and the origin of
pseudoborings. Palaeontology 31, 939e949.
Puce, S., Calcinai, B., Bavestrello, G., Cerrano, C., Gravili, C., Boero, F., 2005. Hydrozoa
(Cnidaria) symbiotic with Porifera: a review. Mar. Ecol. 26, 73e81.
Ristedt, H., Schuhmacher, H., 1985. The bryozoan Rhynchozoon larreyi (Audouin, 1826)da
successful competitor in Coral Reef Communities of the Red Sea. P.S.Z.N.I. Mar. Ecol.
6, 167e179.
Ross, A., Newman, W.A., 1995. A coral-eating barnacle, revisited (Cirripedia:
Pyrgomatidae). Contrib. Zool. 65, 129e175.
Ross, A., Newman, W.A., 2000. Coral barnacles: Cenozoic decline and extinction in the
Atlantic/East Pacific versus diversification in the Indo-West Pacific. In: Proc. 9th Int.
Coral Reef Symp, 1, pp. 179e184.
Rouse, G.W., 2005. Myzostomida (myzostomids). In: Rohde, K. (Ed.), Marine Parasitology.
CSIRO Publishing, Collingwood, pp. 189e193.
Santos, A., Mayoral, E., Baarli, B.G., Da Silva, C.M., Cachao, M., Johnson, M.E., 2012.
Symbiotic association of a pyrgomatid barnacle with a coral from a volcanic Middle
346 Paul D. Taylor

Miocene shoreline (Porto Santo, Madeira Archipelago, Portugal). Palaeontology 55,


173e182.
Simpson, T.L., 1973. Coloniality among the Porifera. In: Boardman, R.S., Cheetham, A.H.,
Oliver Jr., W.J. (Eds.), Animal Colonies. Dowden, Hutchinson & Ross, Stroudsburg,
pp. 549e565.
Soto, F., Méndez-Bedia, I., 1985. Estudio de una asociaci on coral Rugoso-Estromatop orido
en el arrecife de Arnao (Fm. Moniello, Asturias, NO de Espa~ na). Trab. Geol. Fac. Cienc.
Univ. Oviedo 15, 203e209.
Staples, D., 2005. Pycnogonida (pycnogonids). In: Rohde, K. (Ed.), Marine Parasitology.
CSIRO Publishing, Collingwood, pp. 222e226.
Stel, J.H., Stoep, E.v.d, 1982. Interspecifieke relaties en boringen in einige Silurische
stromatoporen. Grondboor Hamer 1, 11e23.
Suarez Andrés, J.L., 2014. Bioclaustration in Devonian fenestrate bryozoans. The ichnogenus
Caupokeras McKinney, 2009. Span. J. Palaeontol. 29, 5e14.
Tapanila, L., 2002. A new endosymbiont in Late Ordovician tabulate corals from Anticosti
Island, eastern Canada. Ichnos 9, 109e116.
Tapanila, L., 2004. The earliest Helicosalpinx from Canada and the global expansion of
commensalism in Late Ordovician sarcinulid corals (Tabulata). Palaeogeogr. Palaeocli-
matol. Palaeoecol. 215, 99e110.
Tapanila, L., 2005. Palaeoecology and diversity of endosymbionts in Palaeozoic marine
invertebrates: trace fossil evidence. Lethaia 38, 89e99.
Tapanila, L., 2008. Direct evidence of ancient symbiosis using trace fossils. Palaeontol. Soc.
Pap. 14, 19e35.
Taylor, P.D., 1994. Evolutionary palaeoecology of symbioses between bryozoans and hermit
crabs. Hist. Biol. 9, 157e205.
Taylor, P.D., Voigt, E., 2006. Symbiont bioclaustrations in Cretaceous cyclostome
bryozoans. Cour. Forschungsinst. Senckenberg 257, 131e136.
Taylor, P.D., Wilson, M.A., 2003. Palaeoecology and evolution of marine hard substrate
communities. Earth Sci. Rev. 62, 1e103.
Tilbrook, K.J., 1997. Barnacle and bivalve associates of a bryozoan-coral symbiosis from the
Coralline Crag (Pliocene) of England. Tert. Res. 18, 7e22.
Tsang, L.M., Chu, K.H., Nozawa, Y., Chan, B.K.K., 2014. Morphological and host speci-
ficity evolution in coral symbiont barnacles (Balanomorpha: Pyrgomatidae) inferred from
a multi-locus phylogeny. Mol. Phylogenet. Evol. 77, 11e22.
Underwood, C.J., 1993. The position of graptolites within Lower Palaeozoic planktic
ecosystems. Lethaia 26, 189e202.
Urbanek, A., Mierzejewski, P., 1982. Ultrastructure of the tuboid graptolite tubotheca.
Pal€aontol. Z. 56, 87e93.
Vinn, O., M~ otus, M.-A., 2008. The earliest endosymbiotic mineralized tubeworms from the
Silurian of Podolia, Ukraine. J. Paleontol 82, 409e414.
Vinn, O., M~ otus, M.-A., 2012. New endobiotic cornulitid and Cornulites sp. aff. Cornulites
celatus (Cornulitida, Tentaculita) from the Katian of Vormsi Island, Estonia. GFF 134,
3e6.
Vinn, O., M~ otus, M.-A., 2014a. Endobiotic rugosan symbionts in stromatoporoids from the
Sheinwoodian (Silurian) of Baltica. PLoS One 9 (2), e90197.
Vinn, O., M~ otus, M.-A., 2014b. Symbiotic worms in biostromal stromatoporoids from the
Ludfordian (Late Silurian) of Saaremaa, Estonia. GFF 136, 503e506.
Vinn, O., Wilson, M.A., 2010. Endosymbiotic Cornulites in the Sheinwoodian (Early
silurian) stromatoporoids of Saaremaa, Estonia. N. Jb. Geol. Pal€aontol. Abh 257, 13e22.
Vinn, O., Wilson, M.A., M~ otus, M.-A., Toom, U., 2014. The earliest bryozoan parasite:
middle Ordovician (Darriwilian) of Osmussar Island, Estonia. Paleogeogr. Palaeoclima-
tol. Palaeoecol. 414, 129e132.
Differentiating Parasitism and Other Interactions in Fossilized Colonial Organisms 347

Vogel, G., 2008. The inner lives of sponges. Science 320, 1028e1030.
Voigt, E., 1959. Endosacculus moltkiae n. g. n. sp., ein vermutlicher fossiler Ascothoracide
(Entomostr.) als Cystenbildner bei der Oktokoralle Moltkia minuta. Pal€aontol. Z. 33,
211e223.
Voigt, E., 1967. Ein vermutlicher Ascothoracide (Endosacculus (?) najdini n. sp.) als Bewohner
einer kretazischen Isis aus der UdSSR. Pal€aontol. Z. 41, 86e90.
Wilson, M.A., 2007. Macroborings and the evolution of marine bioerosion. In: Miller III, W.
(Ed.), Trace Fossils. Concepts, Problems, Prospects. Elsevier, Amsterdam, pp. 356e367.
Wilson, M.A., Vinn, O., Palmer, T.J., 2014. Bivalve borings, bioclaustrations and symbiosis
in corals from the Upper Cretaceous (Cenomanian) of southern Israel. Paleogeogr.
Palaeoclimatol. Palaeoecol. 414, 243e245.
Wulff, J.L., 1985. Clonal organisms and the evolution of mutualism. In: Jackson, J.B.C.,
Buss, L.W., Cook, R.E. (Eds.), Population Biology and Evolution of Clonal
Organisms. Yale University Press, New Haven, pp. 437e466.
Young, G.A., Noble, J.P.A., 1989. Variation and growth of a syringoporid species in
stromatoporoids from the Silurian of eastern Canada. Mem. Assoc. Australas. Palaeontol.
8, 91e98.
Zapalski, M.K., 2007. Parasitism versus commensalism: the case of tabulate endobionts.
Palaeontology 50, 1375e1380.
Zapalski, M.K., 2009. Parasites in Emsian-Eifelian favosites (Anthozoa, Tabulata) from the
Holy Cross Mountains (Poland): changes of distribution within colony. Spec. Pub.
Geol. Soc. Lond. 314, 125e129.
Zapalski, M.K., 2011. Is absence of proof a proof of absence? Comments on commensalism.
Palaeogeogr. Paleoeclimatol. Palaeoecol. 302, 484e488.
Zapalski, M.K., Hubert, B.L.M., 2011. First fossil record of parasitism in Devonian calcareous
sponges (stromatoporoids). Parasitology 138, 132e138.
Zhen, Y.Y., 1996. Succession of coral associations during a Givetian transgressive-regressive
cycle in Queensland. Acta Palaeontol. Pol. 41, 59e88.
Zibrowius, H., 1981. Associations of Hydrocorallia Stylasterina with gall-inhabiting
Copepoda Siphonostomatoidea from the south-west Pacific. Part I. On the stylasterine
hosts, including two new species, Stylaster papuensis and Crypthelia cryptotrema. Bijdr.
Dierkd. 51, 268e286.
CHAPTER NINE

Palaeoparasitology e Human
Parasites in Ancient Material
 jo*, 1, Karl Reinhardx, Luiz Fernando Ferreira*
Adauto Arau
*Fundaç~ao Oswaldo Cruz, Laborat orio de Paleoparasitologia, Rio de Janeiro, RJ, Brazil
x
School of Natural Resources, University of Nebraska, Lincoln, NE, USA
1
Corresponding author: E-mail: adauto@ensp.fiocruz.br

Contents
1. Introduction e Parasitism 350
2. Humans and Parasites 352
3. Palaeoparasitology 353
4. Recommended Material and Techniques for Microscopic Examination in 363
Palaeoparasitology
4.1 Light microscopy techniques 363
4.2 Counting remains under the microscope 365
4.2.1 Analysis of sediments 365
4.2.2 Coprolites in mummies 367
4.3 Molecular techniques applied to palaeoparasitology 368
4.3.1 Molecular diagnosis 368
5. Parasite Finds in Human Archaeological Remains 369
5.1 Ascaris lumbricoides and Trichuris trichiura 369
5.2 Hookworms 371
5.3 Enterobius vermicularis 372
5.4 Diphyllobothrium sp. 373
6. Other Parasites: Parasites of Animals Found in Human Coprolites; Parasites in 374
Prehistoric Asia
7. Origin and Evolution of Trypanosomatids in Humans and the Paradigm Shift 375
from Results in Palaeoparasitology
8. Ectoparasites 376
9. Conclusions 377
Acknowledgements 378
References 378

Abstract
Parasite finds in ancient material launched a new field of science: palaeoparasitology.
Ever since the pioneering studies, parasites were identified in archaeological and palae-
ontological remains, some preserved for millions of years by fossilization. However, the
palaeoparasitological record consists mainly of parasites found specifically in human
archaeological material, preserved in ancient occupation sites, from prehistory until
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.03.003 All rights reserved. 349
350 Adauto Ara
ujo et al.

closer to 2015. The results include some helminth intestinal parasites still commonly
found in 2015, such as Ascaris lumbricoides, Trichuris trichiura and hookworms, besides
others such as Amoebidae and Giardia intestinalis, as well as viruses, bacteria, fungi and
arthropods. These parasites as a whole provide important data on health, diet, climate
and living conditions among ancient populations. This chapter describes the principal
findings and their importance for knowledge on the origin and dispersal of infectious
diseases.

1. INTRODUCTION e PARASITISM
This chapter uses the broad definition of parasitism. The concept
ranges all the way from mobile fragments of genetic material, both transmis-
sible and non-transmissible, up to higher plants and vertebrate animals
(Trager, 1988; Ara ujo et al., 2003; Araujo and Ferreira, 2014; Schmid-
Hempel, 2011; Ewald and Swain Ewald, 2014). The definition considers
the complex parasiteehosteenvironment system and its interdependent
relations, where changes or alterations in one component can influence
the others. This is an ecological focus to parasitism, since both the macro-
environment and microenvironment induce changes and influence rela-
tions. The study of parasitism should be based on ecology. These
relations consider the classifications between symbionts, commensals,
mutualists and parasites, among others, as a single concept called parasitism.
This unified approach allows a better understanding of parasitic diseases in
their evolutionary process.
The broad definition of parasitism is not new. Brazilian parasitologist
Samuel B. Pessoa, in the first edition of his book Parasitologia Médica in
1945, already reminded parasitologists that ‘parasitism, commensalism, and
symbiosis are categories created by our spirit, all displaying characteristics
of the same general laws’ (Pessoa, 1951, p. 5).
When German scientists Karl Georg Friedrich Rudolf Leuckart and
Heinrich Anton de Bary independently created the concepts of symbiosis
and parasitism, respectively (both in 1879), they made no mention of dis-
eases, benefits or harm that could result from the relationship between spe-
cies. The same connotation was already revisited by various other authors,
including French parasitologist Emile Brumpt (Brumpt, 1913). Beginning
with the first editions of his book, Brumpt highlighted the fact that para-
sitism and symbiosis converged in a single concept, since beings that were
considered parasites and those viewed as symbionts or commensals could
all cause harm or benefit to their hosts.
Palaeoparasitology e Human Parasites in Ancient Material 351

We thus propose the concept of parasitism as an organism that finds its


ecological niche in another, namely, its host (Ara ujo et al., 2003; Ara
ujo
and Ferreira, 2014). The disease or harm does not result inexorably from
the parasite’s presence in the host, but can also result from conditions of
the parasite, the host or the environment.
This article mainly features protozoa, helminths and arthropods, always
highlighting the ecological characteristics of parasitism. Other parasites
will be mentioned in passing.
The article assumes Aidan Cockburn’s ideas on parasitic diseases: ‘Infec-
tious disease is composed of three variables, the host, the pathogen and the
environment. It is a constant state of flux, capable of changing in step with
any variation in any one of its components. New diseases appear, old ones
alter, and some may disappear completely’ (Cockburn, 1963). The process
is dynamic. Parasitic infection means a parasite’s presence in a given host,
but parasitic (or infectious) disease is the set of signs and symptoms that char-
acterize an alteration resulting from changes in one or more of the compo-
nents in the parasiteehosteenvironment system. Infection and disease are
thus distinct phenomena in the host, both dependent on the parasite’s pres-
ence. Parasitic disease can occur in an individual or become a population-
based phenomenon, characterized as an endemic or epidemic, with different
overall impacts on the hosts. On the one hand, infection does not necessarily
lead to disease, and many parasite species can be present in the host, often
throughout life, without any symptoms. Infection is not synonymous
with disease, and parasitic disease is the same as infectious disease (Ara ujo
et al., 2003).
We use the concept of virulence as the parasite’s capacity to multiply and
spread in a given host population, while pathogenicity refers to the capacity
to induce a lesion or disease (Ara ujo et al., 2003). Other authors use the
terms synonymously (Poulin and Combes, 1999). However, the term path-
ogen becomes inapplicable, since the development of signs and symptoms in
the host depends on the three components (parasite, host and environment),
and the parasite’s presence is thus not the exclusive determinant of the dis-
ease. We will always use the term ‘parasite’, both for infections and for their
capacity to cause lesions in the host.
There are various examples of the above-mentioned issues, and they can
be exemplified by the classical experiment on infection with Entamoeba inva-
dens in snakes (Barrow and Stockton, 1960). When the amoebae remain in
the host at an adequate temperature, they behave without causing any clin-
ical manifestations, feeding on bacteria and other elements from the
352 Adauto Ara
ujo et al.

intestinal content. However, when raising the temperature alters the envi-
ronment, the parasites invade the tissues and the snake can die. Barrow
and Stockton (1960) conducted experiments in eight species of snakes
infected with this parasite under different temperatures, from 13  C to
25  C. No lesions were observed at lower temperatures, although the
amoebae continued to be isolated in culture. However, at a temperature
25  C, the snakes presented lesions, and even those that did not show
lesions at low temperatures became sick at higher temperatures.
When Carlos Chagas described the signs and symptoms of American
trypanosomiasis, later named Chagas disease in his honour, the first case of
an infected patient was a child with an acute manifestation of the disease,
in whom trypanosomes were found in the bloodstream (Coura, 1997).
The child was Berenice, and she died at 79 years of age from a cause other
than Chagas disease. For years she attended congresses during which she was
examined without presenting severe symptoms, but it was always possible to
isolate Trypanosoma cruzi from her peripheral blood during her >70 years of
life (Salgado, 1980).

2. HUMANS AND PARASITES


Throughout human evolution, from prehominids to present-day
Homo sapiens, changes occurred in the relationship between humans and
their parasites due to territorial occupations and changes made to environ-
ments by a host capable of creating technologies, migrating and altering
newly occupied territories in a complex sociobiological relationship with
nature. In this context, palaeoparasitology consists of methods and tech-
niques capable of recovering, in space and time, diverse moments in the tra-
jectory of human hosts and their parasites, both those inherited from
ancestors and those acquired from the environment while humans have
roamed the planet.
Like all biological species, parasites have a single origin. When they
become species-specific in a given host species, they accompany it wher-
ever the environment allows the maintenance of their life cycle. Parasites
are excellent markers of prehistoric migrations, since they are found in
occupation sites in events and landscapes traversed by ancestral human
groups (Ara ujo et al., 2008). These are called phylogenetically inherited
parasite species, when they are shared by phylogenetically proximate hosts,
as exemplified by Enterobius vermicularis infection in humans and
Palaeoparasitology e Human Parasites in Ancient Material 353

phylogenetically closely related primates. There are also species of parasites


that originated from phylogenetically remote hosts and that adapted to
humans over the course of their evolutionary history and the conquest of
new territories. An example is infection with Trypanosoma cruzi, a proto-
zoan whose original life cycle included wild mammals and the insect vector.
However, T. cruzi adapted perfectly to the human host, including the
domiciliation of triatomine species, ever since the first human migrations
to the Americas (Ferreira et al., 2011). These are called ecologically
acquired parasites, having adapted to (or infected) humans during part of
their cycle by mechanisms in which humans approached their natural niche
of infection (Pavlovsky, 1964) or pathocenosis (Grmek, 1983). Pathoecol-
ogy has been defined as reconstruction of the ecology of a disease based on
evidence recovered from archaeological sites (Reinhard and Pucu, 2014),
which can also explain some parasites acquired during the social and biolog-
ical evolution of the human species.

3. PALAEOPARASITOLOGY
Palaeoparasitology is the study of parasites found in archaeological or
palaeontological material (Ferreira, 2014). This branch of palaeopathology
and parasitology opened new avenues for studying the evolution of the
health-disease process in the human species, as well as in other hosts
(Mitchell, 2013). Palaeoparasitological findings provide a consistent source
for palaeoepidemiology by revealing the presence of infection in a given
archaeological context and the possible consequences for the emergence
or disappearance of infectious diseases in prehistoric populations.
Franco-British physician and microbiologist Sir Marc Armand Ruffer
was the pioneer of palaeoparasitology. Ruffer studied Egyptian mummies
in the early twentieth century, developing rehydration techniques for
mummified tissues, allowing their visual examination in histological sections
with staining used in routine histopathology. He found calcified and well-
preserved Schistosoma haematobium eggs in the kidney tissue of a mummy
dated 3200 BC (Ruffer, 1910). This was the first parasite find in ancient
human material, shedding light on haematuria in Egypt, which had been
described in ancient texts but not previously proven. Although Ruffer’s
article does not show an image of the egg, his diagnosis was confirmed by
German parasitologist Arthur Looss and Scottish pathologist Dr Alexander
R. Ferguson, Ruffer’s faculty colleagues at the Cairo School of Medicine.
354 Adauto Ara
ujo et al.

For years, little was published in the field of palaeoparasitology, although


some studies were done with coprolites from extinct animals and with
human coprolites, aimed at describing ancient diet. The first results that
demonstrated the presence of intestinal helminths in preserved human
remains came from cooperation between archaeologists and parasitologists.
Lothar Szidat described Ascaris lumbricoides and Trichuris trichiura eggs in a
naturally mummified human body discovered during excavations of peat
fields in Prussia (Szidat, 1944), in a swampy region with low pH that facili-
tated the preservation of bodies sacrificed hundreds of years ago.
Other collaborations between archaeologists and parasitologists occurred
during this initial period, as reviewed by Dittmar (2009) and Reinhard and
Araujo (2012), but it was the partnership between botanist Eric Callen and
parasitologist Thomas Cameron that laid the ground for the technique
used to study parasites in coprolites. Callen and Cameron (1960) developed
the rehydration of desiccated materials with trisodium phosphate aqueous
solution. With this technique, desiccated materials like unlithified coprolites
(and others potentially containing parasites) recover their original mor-
phology, allowing parasitological testing using routine techniques. Rehydra-
tion does not alter the parasite structure measurements, and the standard
measurements of current parasites can be used (Fugassa, 2014). Since then,
parasitological tests in coprolites have multiplied, with growing collaboration
between parasitologists and archaeologists. Coprolites are found in archaeo-
logical layers (Figures 1e3) or in the intestinal contents of preserved bodies.
In 1978, research in palaeoparasitology in Brazil began at the Oswaldo
Cruz Foundation, and the term palaeoparasitology was coined to refer to

Figure 1 Archaeological layers at the Site of Toca do Boqueir~ao da Pedra Furada, 1986
(Piauí state, Brazilian northeast).
Palaeoparasitology e Human Parasites in Ancient Material 355

Figure 2 The Italian archaeologist Fabio Parenti excavating a coprolite in the Site of
Toca do Boqueir~ao da Pedra Furada, 1986 (Piauí state, Brazilian northeast).

this field of knowledge (Ferreira et al., 1979). At the time, there was little
research production in the area, with just a few articles on parasite finds pub-
lished by North American authors in important journals or even mimeo-
graphed copies (Samuels, 1965; Heizer and Napton, 1969; Fry, 1970) and
reviews on analyses in human coprolites in the United States, especially
focussing on diet and food remains (Wilke and Hall, 1975), while England
and a few other European countries featured studies on parasites in ancient
material (Gooch, 1972; Jones, 1982). As with the first studies in these coun-
tries, Brazilian researchers developed collaborations with archaeology teams,
since the entire palaeoparasitological analysis hinged on painstaking

Figure 3 Archaeological layer showing the coprolite evidenced in the Site of Toca do
Boqueir~ao da Pedra Furada, 1986 (Piauí state, Brazilian northeast).
356 Adauto Ara
ujo et al.

archaeological excavations. It is indispensable to determine the sample’s age,


either through radiocarbon dating or another applicable technique, as well as
a detailed description of the archaeological context.
Brazil’s first samples from coprolites and mummified bodies came from
the Institute of Brazilian Archaeology in Rio de Janeiro and soon afterwards
from sites in Minas Gerais State and the archaeological region of S~ao Rai-
mundo Nonato, Piauí State, in Northeast Brazil, where we found hook-
worm and Trichuris trichiura eggs (Ferreira et al., 1980, 1983; Ara ujo et al.,
1981). Our research group began collaboration with archaeologists from
other countries early on: in 1983, we received samples of coprolites from
Dr Lautaro Nu~ nez at Universidad del Norte in Chile. These were human
coprolites containing Diphyllobothrium pacificum eggs, according to morpho-
metric diagnosis. This was the first diagnosis of D. pacificum in prehistoric
groups from the Andean coast, followed by a series of other finds, confirm-
ing the ingestion of raw sea fish by the local population (Ara ujo et al., 1983;
Patrucco et al., 1983). These eggs differ only in size from those of another
species, Diphyllobothrium latum, that has infected the human host since the
Neolithic and whose cycle takes place in freshwater fish and is common
in Europe.
Collaboration between parasitologists and archaeologists led to new
progress in the field (Reinhard, 1992a; Ferreira et al., 1984; Dittmar,
2009), concentrating on the study of parasites in ancient material and their
interpretations. Palaeoparasitology is a broad field, not necessarily limited
to the human host or to animals that may transmit their parasites to humans
but also including studies of parasites in extinct or current animals, whose
remains can be found and which contribute to the study of the palaeoclimate
and the origins of parasitism in current hosts (Poinar and Boucot, 2006;
Dentzien-Dias et al., 2013; Poinar, 2014; Silva et al., 2014).
Palaeoparasitology now intersects with various other fields of science,
based on the extensive current literature, featuring special sections in main-
stream scientific journals such as The Journal of Parasitology (Faulkner and
Reinhard, 2014) and the International Journal of Paleopathology (Buikstra,
2013; Dittmar et al., 2012; Dittmar, 2013). A major portion of the research
focuses on the reconstitution of eating habits and health in the past, inferred
by finding parasites that are specific to certain animals (Le Bailly and Bouchet,
2013; Arriaza et al., 2013a; Sianto et al., 2012; Jimenez et al., 2012; Reinhard
et al., 2013); diet, parasites and health of populations (Reinhard et al., 2007);
description of specific human parasites, the finding of which allowed new
approaches to theories on migrations and prehistoric contacts between Asian
Palaeoparasitology e Human Parasites in Ancient Material 357

and American populations by transpacific navigations (Ara ujo et al., 1981,


1988, 2008) and other great waves of human groups moving around the
medieval world (Anastasiou and Mitchell, 2013), and studies in palaeoepi-
demiology and the origin of infectious diseases (Hugot et al., 1999; Ara ujo
et al., 2009; Ferreira et al., 2011; Baum and Bar-Gal, 2003), among others.
In addition to these advances in knowledge obtained by microscopic
analysis of ancient material, new techniques, particularly in molecular
biology, have allowed diagnoses that were previously imprecise or false-
negative, like the detection of infections by intestinal and systemic protozoa.
The diagnosis of protozoa can now be done with specific enzyme-linked
immunosorbent assays kits, as with Giardia intestinalis, Cryptosporidium sp.
and Entamoeba histolytica (Gonçalves et al., 2004; Le Bailly et al., 2008; Frías
et al., 2013), and molecular biology techniques for systemic protozoa
(Fernandes et al., 2008; Lima et al., 2008; Guhl et al., 2014).
There is debate on the origin and dispersal of some diseases such as syph-
ilis, tuberculosis and the plague in Greece during the siege of Athens around
400 BC. The debate on the Athens plague involves speculation, that is,
whether it was an epidemic of smallpox, epidemic typhus, typhoid fever
or bubonic plague itself. Researchers from the University of Athens pub-
lished the results of the analysis of dental pulp from skeletons dated to
that time and showed the presence of DNA fragments from Salmonella enter-
ica serovar Typhi, the etiological agent of typhoid fever, but their conclu-
sions were challenged by Shapiro et al. (2006). As for syphilis and
tuberculosis, the discussion focuses on their origin or absence in the Amer-
icas before the arrival of Columbus and his sailors. Syphilis purportedly orig-
inated in the Americas, but tuberculosis was believed to have been
introduced during colonial times. Data from palaeoparasitology now leave
no doubt that the tuberculosis bacillus already circulated in the prehistoric
population of the Americas (G omez i Prati et al., 2003; Wilbur and Buikstra,
2006; Wilbur et al., 2008; Stone et al., 2009; Jaeger et al., 2013), but the jury
is still out on syphilis, although the origin of Treponema species in the African
continent is well established (Harper et al., 2008).
Archaeologists, with the exception of excavations performed by palaeon-
tologists, sponsored all palaeoparasitological studies. Each archaeologist had his
or her own research goal to uncover ancient cultural practices and to under-
stand the vitality of those cultures. Recognizing this, Reinhard (1992a,b)
called attention to parasitology as a tool for archaeologists (Reinhard et al.,
1985). As reviewed by Bryant and Reinhard, several archaeologists adjusted
their research designs specifically to recover coprolites (Bryant and Reinhard,
358 Adauto Ara
ujo et al.

2012; Reinhard and Bryant, 2008). Parasitologists recover data that have been
unique in fulfiling archaeological research goals (Reinhard, 1992a,b).
The vitality of archaeological populations has been an inherent theme in
archaeology that was codified in ‘bioarchaeology’, which focuses on the
analysis of skeletal human remains (Buikstra and Beck, 2008). By compara-
tive analysis of bone pathology, bioarchaeologists assess the relative adaptive
success of past populations. Because skeletons were the source of data, evi-
dence of infectious disease was often limited to nonspecific indicators of
stress, periosteal bone reaction and porotic hyperostosis. The evidence for
actual infectious organisms comes from analysis of sediments associated
with skeletons (Fugassa, 2014) or from coprolites (Jiménez et al., 2012).
A long-standing archaeological hypothesis was that ancient hunter-
gatherers, relative to later agricultural peoples, were free of infections
(Diamond, 1987). Testing this hypothesis was a research design of many
coprolite recovery excavations in the 1960s and 1970s. To a large degree,
archaeology and ethnography support this hypothesis (Reinhard, 1988;
London and Hruschka, 2014). The low prevalence of parasitism led to a
desire to explore the basis for hunter-gatherer avoidance of parasitism.
One aspect is related to the presence of medicinal plants in the diet
(Reinhard, 1985). Beyond the use of anthelmintics among hunter-gatherers,
small band size, diffuse population and frequent movement of camps helped
reduce parasite infection (Reinhard, 1988). The agricultural revolution was
recognized by archaeologists as an abrupt change in the human condition
that gave rise to a dramatic increase in infection (Diamond, 1987). The
bone pathology, porotic hyperostosis, generally exhibits a higher prevalence
among agricultural skeletal series. Porotic hyperostosis is exhibited as ‘spongy’
expansions of the cranial diploe coinciding with an erosion of the outer cra-
nial table of the parietals and occipital of effected skulls. The aetiology of the
pathology was long debated and finally resolved by Walker et al. (2009) who
concluded that porotic hyperostosis lesions ‘are a result of the megaloblastic
anemia acquired by nursing infants through the synergistic effects of depleted
maternal vitamin B12 reserves and unsanitary living conditions that are
conducive to additional nutrient losses from gastrointestinal infections around
the time of weaning’ (Walker et al., 2009, p. 119). The fluctuations of
porotic hyperostosis have been a key issue in archaeology, especially in the
Americas, that now can be asserted to have an origin in large part to parasitism
(Reinhard, 1992a,b). Jiménez et al. (2012) underscored Walker et al. (2009)
model in their characterization of extreme helminth prevalence at a site in
Durango, Mexico, the skeletons from which exhibited porotic hyperostosis.
Palaeoparasitology e Human Parasites in Ancient Material 359

The period of agriculture is itself nuanced with the emergence of com-


plex societies and empire civilizations. These nuances have been explored
in the Andes for the Inca Empire and the earlier Chiribaya culture (Martinson
et al., 2003; Santoro et al., 2003; Reinhard and Buikstra, 2003). The differ-
ences in louse and helminth infections were compared between three villages
in the Ilo area of southern Peru based on the analysis of coprolites from
burials, mummies and latrines. The sites included a coastal fishing village,
an inland farming village and an administrative town. Social differentiation
was best defined at the town where a wide variety of social status was evi-
denced by burial offerings. Fish tapeworm, Diphyllobothrium pacificum was
the most common helminth found. This tapeworm infection was most com-
mon among the farmers (69%), less common among the fishers (25%) and
lowest at the administrative centre (20%). The presence of tapeworm at all
sites indicates that fish was traded to all habitations, but control of the infec-
tion was best accomplished for the ruling elite and among the source fisher
population itself (Martinson et al., 2003). The Chiribaya population was
joined by refugees from the collapse of the Tiwanaku Empire of the Lake
Titicaca region. Analysis of head lice from the refugee population mummies
compared to the Chiribaya mummies showed that the refugees were more
infested than even the most infested Chiribaya demographic group. This sug-
gests that the lifestyle of the refugees was cramped and stressful such that
grooming was reduced and reinfestation was unavailable (Reinhard and
Buikstra, 2003). The Chiribaya culture farmers of the Atacama Desert of
northern Chile were sometimes subsumed in the Inca Empire. This was
the fate of Mitas Chiribaya farmers in the Lluta Valley of Chile. Because of
the high corn production of the valley, it was absorbed into the Empire. Pre-
viously, farmers had lived in dispersed, small communities. Under the Incas,
the farmers were aggregated into large concentrated towns. The corn pro-
duction of the Inca farmers was largely taken away in taxes so that the farmers
had to diversify their subsistence with fishing. Parasitologically, both Inca and
Chiribaya farmers were parasitized by whipworm (Trichuris trichiura). How-
ever, differences emerged as the farmers became incorporated into the
Empire. Hymenolepidid infection, present in the Chiribaya farmers, disap-
peared among the Inca farmers. Pinworm infection (Enterobius vermicularis)
was absent among Chiribaya farmers but subsequently reached a prevalence
of 21% among the Inca farmers. Thus, the parasite record signals crowding
and subsistence diversification as the Empire expanded.
Catastrophic climate events can have palaeoparasitological implications.
Arriaza et al. (2010) found this to be the case for preagriculture Chinchorro
360 Adauto Ara
ujo et al.

peoples of the coast of northern Chile. The data suggest that El Ni~ no events
caused fluctuation in coastal fish resources that in turn led to varying suscep-
tibility of fishers to fish tapeworm infection. But drought in another part of
the hemisphere is reflected in the archaeoparasite record. The excavation of
Antelope House in Canyon de Chelly, Arizona, revealed the hundreds of
years of prehistory of a Puebloan village that terminated with the Great
Drought that lasted from 1276 and continued through 1299 (Morris,
1988). Parasite prevalence and diversity increased at the end of the occupa-
tion as people aggregated at the last villages with persistent water sources
(Reinhard, 2008). This was mirrored at Elden Pueblo near Flagstaff, Ari-
zona, that showed the highest concentration of helminth eggs in the final
occupation strata of the site. Parasitism peaked around the time of
abandonment.
The emergence and control of parasitism in historic times, and relation
of ethnicity to infection, is a topic among historic archaeologists. Charles
Fisher and his colleagues directed excavations of latrine sediments, as well
as sediments from yards, streets, drains and other contexts to reconstruct
the emergence and control of parasitism in Albany, New York. The analysis
of hundreds of samples showed the introduction of geohelminths (soil-
transmitted helminths) with the original Dutch colony. Infection peaked
in the late 1700s and early 1800s but began to abate in the late 1800s as
drainages were covered and as water projects brought clean water into
the city. Interestingly, the water projects were designed first to provide
water for fighting fires. By the turn of the twentieth century, faecal-borne
parasitism was controlled in both the elite and poor neighbourhoods
(Fischer et al., 2007).
The longer-term emergence and control of faecal-borne geohelminths
have been a source of research for many analysts working in Europe (Reinhard
and Pucu, 2014). This research falls into the archaeological construct of
palaeoepidemiological transitions. There were two palaeoepidemiological
transitions. The first was associated with the Neolithic Revolution, which
separated the ancient hunter-gatherer prehistory with the agricultural period.
The Neolithic Revolution is thought to have resulted in a rise in infectious
diseases. The Industrial Revolution is thought to represent reduction of
infectious disease and a rise of occupational disease. This scheme was tested
by Reinhard and Pucu (2014) for both Europe and the Americas. Interest-
ingly, the palaeoparasitology record of Europe bears out the hypothesized
changes. The Neolithic Revolution is accompanied by the ubiquitous pres-
ence of whipworm and roundworm (Ascaris lumbricoides) infections. By
Palaeoparasitology e Human Parasites in Ancient Material 361

Roman times, these parasites are found wherever archaeologists sample lat-
rines, burials or mummies. This transition, strangely, does not exist in the
Americas where geohelminths, although introduced long in prehistory,
never take hold as ubiquitous parasites among agriculturalists. The reasons
for a relatively geohelminth free American prehistory have yet to be eluci-
dated. Reinhard and Pucu (2014) suggest that Native American medicinal
treatments, settlement patterns, faecal avoidance and lower population den-
sity may have helped maintain post-agricultural revolution parasitism at low
levels. In addition, the European use of human faeces as fertilizer and depen-
dence on the humoral theory of disease may have contributed to spread of
infections and failure to effectively treat infections (Reinhard and Pucu,
2014).
When ideal archaeological preservation can be combined with historical
documents, very detailed reconstruction of infection patterns can be recon-
structed. Kim et al. (2013) and Seo et al. (2014b) accomplished this for the
city of Hansung, the capital area of the Joseon Dynasty, which is now
located within Seoul, South Korea. By examining legal documents relating
to the distribution of night soil (latrine contents) from the city to farmers,
and the economic interdependence between major cities and nearby farm-
lands, researchers discovered that the night soils produced in a major city
were recycled by farmers as fertilizers for fields. With that night soil, soil-
transmitted parasites were spread on the fields to contaminate vegetables
that were then marketed back to the city. High levels of geohelminth infec-
tion were an unavoidable result of this legally defined recycling process.
Flood-deposited sediments in the city that were washed in from farm fields
during periodic torrential rains in ancient times show that parasite eggs were
indeed contaminants of the fields.
Remarkable preservation of Joseon Dynasty mummies provides fasci-
nating evidence of food-borne fluke infection. Eighteen Joseon mummies
have been studied for parasites as reviewed by Seo et al. (2014a,b). Four spe-
cies of trematodes were found. Two species, Clonorchis sinensis (Chinese liver
fluke) and Metagonimus yokogawai, are similar in that they are transferred by
the consumption of fish. A total of six mummies were infected with one or
both of these flukes, five with C. sinensis and three with M. yokogawai. Para-
gonimus westermani, a lung fluke, was found in four of the mummies. Its final
intermediate host includes species of freshwater crabs. Finally, oyster con-
sumption was a source of infection with Gymnophalloides seoi, an intestinal
fluke. Two mummies were found to be positive for eggs of this parasite
(Shin et al., 2012). In total, 11 of 18 Joseon mummies, representing the elite
362 Adauto Ara
ujo et al.

of this society, were infected with one or more species of fluke and under-
scores the importance of uncooked meat in Ancient Korean diet.
The studies summarized above present a sampling of the ‘big picture’
issues that have been addressed by cooperation of parasitologists and archae-
ologists. Other studies provide ‘vignettes’ of ancient parasitism. Such studies
focus on brief archaeological moments or give a detailed glimpse of humane
parasite interactions. Kristjansd
ottir and Collins (2011) excavated cemeteries
from Skriðuklaustur, a medieval monastic site and hospital used between AD
1493 and 1554 in Eastern Iceland. During this brief period, eight individuals
died with hydatid cysts from eight of 160 burials recovered. The eight burials
were clustered together in one part of the cemetery, which suggests that they
were recognized as a specific symptom complex and were treated as such.
A cyst in one of the skeletons was 17 cm in diameter. The cysts are a result
of infection with Echinococcus granulosus. It was probably introduced in Ice-
land sometime after the late ninth century and became endemic by AD
1200. Although some of infected individuals lived with other diseases,
including syphilis, the authors note that ‘they seem to have first been buried
according to this shared condition and not relative to any of the other iden-
tified illnesses or age. The indications are that this ailment may have been
recognised in medieval Iceland as having its own classification and perhaps
requiring distinct treatment if not in life, at least in death’ (Kristjansdottir
and Collins, 2011, p. 485). The implication is that hydatid cyst disease
was recognized early in Icelandic medical history. Thus, this excavation pro-
vides a ‘vignette’ of a symptom complex recognized by monastic scholars.
Latrines provide other ‘vignettes’ of human parasitism. The diet and
parasitism of a single Roman Centurion were presented by Kuijper and
Turner (1992). Along with a great diversity of pollen and seeds of spices
and food, the eggs of intestinal parasites were found. The authors report
that ‘many thousands’ of eggs were present per cubic centimetre. Whip-
worm and Ascaris lumbricoides were most common, but taeniid eggs, consis-
tent with the genera Taenia or Echinococcus, were also present. Thus, the
dietary habits and intestinal infections of a single officer were evidenced in
the latrine sediments. Another vignette of parasites carried by pilgrims comes
from a thirteenth century latrine in the city of Acre, Israel. At that time, Acre
was part of the Frankish Kingdom of Jerusalem. The fish tapeworm (eggs
consistent with Diphyllobothrium latum) was identified. The authors note
that fish tapeworm was also found in the latrine block of the Hospital of
St John. The Order of St John was a religious order that cared for crusaders
and pilgrims. The presence of fish tapeworm eggs in a crusader period latrine
Palaeoparasitology e Human Parasites in Ancient Material 363

in the Levant indicates that infected pilgrims from northern Europe travelled
to Acre with active fish tapeworm infections (Mitchell et al., 2011).
This sampling of studies reflects the popularity of parasitology as a tool
for archaeologists (Reinhard, 1992a). The productive collaboration of
archaeology and parasitology will continue to grow in regional distribution
and in the development of new research questions beneficial to both fields.

4. RECOMMENDED MATERIAL AND TECHNIQUES FOR


MICROSCOPIC EXAMINATION IN
PALAEOPARASITOLOGY
4.1 Light microscopy techniques
It is always important to highlight the use of trisodium phosphate
aqueous solution (Na3PO4), introduced by Callen and Cameron (1960)
and based on the experiments by Van Cleave and Ross (1947) for the rehy-
dration of desiccated samples. Based on the rehydration time recommended
by the authors (72 h of immersion of the coprolite), bacterial or fungal
contamination can occur and jeopardize the test results. Thus, in our labo-
ratory, we began to use several drops of acetic formalin to prevent the
growth of contaminant microorganisms (Ara ujo et al., 1998). Since the
advent of molecular biology techniques, formalin is no longer recommen-
ded, and samples submitted to rehydration should preferably be refrigerated
to avoid the development of current microorganisms (Fugassa, 2014).
Samuels (1965) conducted the first tests in various solutions and
concluded by recommending trisodium phosphate solution. Experiments
in our laboratories that used other rehydration solutions such as distilled
water, potassium or sodium hydroxide, physiological saline, and others
resulted in significant losses of parasite elements previously counted in fresh
faeces after experimental desiccation and rehydration. The results have not
been published, but serve as recommendations for other authors. For
example, Giardia intestinalis cysts practically disappeared after important
warping of the outer membrane, leading to breakage. The same occurred
with hookworm eggs, which are known to be fragile and prone to
breaking. We thus reemphasize using the rehydration solution proposed
by Callen and Cameron (1960).
After rehydration of the material, any parasitological technique for
microscopic examination can be used. However, the Kato-Katz technique
(Siqueira et al., 2011) should be avoided, since the reagents destroy some
eggs after about 2 h, such as those of hookworms, or alter the outer shell
364 Adauto Ara
ujo et al.

in the case of Ascaris lumbricoides. Other concentration techniques that use


strong acid or alkaline reagents can also alter the results. Our laboratories
have successfully used two techniques for parasitological examination. Con-
centration by spontaneous sedimentation is a simple, efficient and inexpen-
sive technique developed by Adolpho Lutz in 1919 (Camacho et al., 2013)
for the diagnosis of schistosomiasis mansoni by finding the parasite’s eggs. It is
still used in clinical test laboratories for both protozoa and intestinal hel-
minths and was introduced in palaeoparasitology because it is efficient,
does not use any reagent except for the rehydration solution, and allows
saving all the material used in the test. Nothing is lost.
The technique consists of taking 2e3 g of rehydrated material, crushing
it with a glass pestle, filtering the suspension through gauze folded four
times, and pouring it through a glass funnel into a 300- to 400-mL conical
glass jar; the suspension is then left to settle for at least 2 h with the rehyd-
rating liquid touching the lower part of the gauze (Figure 4). Next, the sedi-
ment is collected from the glass jar using a disposable Pasteur pipette to
prepare the microscope slides. An additional recommendation is to add a
drop of glycerol to improve the transparency under microscopy. The slides
can be sealed with modified Noyer seal, a 1:1 mixture of resin and beeswax,

(a) (c) (e)

(b)
(d)

Figure 4 Lutz’ sedimentation technique steps e (a) mixing the sediment, after rehyd-
rating the coprolite or the sediment; (b) pouring the rehydrated sediment through the
gauze to the conical glass jar; (c) after 2e4 h, the sedimentation process is completed;
(d) preserved macroresidues; (e) collecting the sediment for microscopic analysis.
Palaeoparasitology e Human Parasites in Ancient Material 365

boiled until the mixture become homogeneous, without emitting any


smoke. Extreme care is needed in preparing the mixture, which is highly
flammable. The slide covers are sealed using a metal loop heated over a Bun-
sen burner, spreading the melted mixture over the edges of the slide cover
with the hot metal loop. Before applying the mixture, it is crucial to dry the
edges of the slide cover in order to avoid spilling liquid over it and prevent-
ing the proper sealing. Under microscopic examination, all the parasite
structures are measured and photographed and the other microfossils are
observed. We recommend examining 20 slides in all the fields for each sam-
ple preserved by desiccation and at least 500 slides for permineralized (fossil)
material, but in the latter case using acids or ultrasound to dissociate the
material (Silva et al., 2014).
All the unused material settled in the glass jar should be collected with a
pipette and stored separately, refrigerated or even frozen, for future use. The
same is true for the material trapped in the gauze, which can be examined for
food remains and occasionally fragments of helminths. In this case to pre-
serve the material, we recommend drying the material in an oven at
45  C or even at room temperature under low relative humidity, on absor-
bent paper, until there is no longer any change in weight.
For samples collected from the pelvic cavity of skeletons or sediment
from archaeological sites, we recommend the sugar solution flotation tech-
niques (Fugassa, 2014), which offers good recovery of parasite structures.
The following is a description of details for counting remains found under
the microscope.

4.2 Counting remains under the microscope


4.2.1 Analysis of sediments
Shaft features are remains of wells, latrines and other vertical pits. Much of
the earliest work in England and Germany focused on the recovery of eggs
from shaft latrines. The challenge in analysing sediments from such features is
separating the parasite evidence from inert sand and other organic remains.
A similar challenge is encountered when recovering eggs from skeletonized
burials. In many sites, the sand and organics are held together in a calcium
carbonate matrix. Two methods from that period are still used today. Jones
(1985) modified the Stoll’s dilution method from clinical parasitology to
archaeological samples. There has been a renewed interest in this simple,
rapid and efficient method (Fugassa et al., 2006, 2008). Another older
method is derived from palynology. Hevly et al. (1979) discovered five spe-
cies of parasites represented in sediment samples from an archaeological site
366 Adauto Ara
ujo et al.

in northern Arizona. This was a significant discovery showing that sediment


digestion is successful in recovering eggs from calcareous soils. In turn this
was a breakthrough because parasite egg distribution could be traced strati-
graphically using long-established pollen analysis techniques. The inert min-
eral components are removed by hydrochloric acid, sedimentation and
finally hydrofluoric acid baths. Finally, the cellulose component is removed
by acetylation. The different steps of this method were tried on eggs from
five parasite species. It was shown that nematode cestode, and trematode
eggs survived the first stages of the processing until acetolysis. There was
some destruction of cestode and trematode eggs in acetolysis solution
(Fischer et al., 2007). From calcareous sediments, the eggs were more readily
identifiable after palynological processing than established flotation methods.
This became a quantitative method when combined with another palyno-
logical method called pollen concentration. This method is based on adding
known numbers of exotic spores to known amounts of sample and was pre-
sented by Warnock and Reinhard (1992). The counted ratio of parasite eggs
to spores can then be used to quantify the numbers of eggs per gram and
millilitre of sediment. Most recently, Florenzano et al. (2012) showed that
careful application of acetolysis, called ‘light acetolysis’, results in remarkably
cleansed eggs with clear surface morphology. This palynologicaleparasito-
logical hybrid method is advantageous for several reasons. It is a rapid
method designed for processing dozens of samples simultaneously. It results
in the recovery of pollen from diet, medicines and environmental sources.
The data can be used to generate graphs of parasite egg distribution vertically
and horizontally. The disadvantage of this method is that specialized equip-
ment, such as an acid-rated fume hood, is needed for the hydrofluoric acid
treatments.
The basic method for the analysis of coprolites has been rehydration in
trisodium phosphate. After rehydration, the methods varied. By 1986, two
main methods had emerged. The Lutz spontaneous sedimentation method,
described above, uses gravitation-based, passive filtration of microscopic
remains through gauze mesh (Camacho et al., 2013; Ferreira et al., 1980;
Reinhard et al., 1986; Sianto et al., 2005). The second method is based on
disaggregating the samples on standard stir plates and screening the disaggre-
gated residues through metal screens (Fugassa et al., 2011; Reinhard et al.,
1986; Reinhard, 1988). The disaggregation/screening method was derived
from previous dietary methods (Bryant, 1974; Reinhard and Bryant,
1992). Both methods had the goal of separating remains larger than about
one-third of a millimetre from microscopic remains. When the Lycopodium
Palaeoparasitology e Human Parasites in Ancient Material 367

parasite egg concentration method was introduced (Sianto et al., 2009),


researchers rapidly incorporated egg concentration with either Lutz or disag-
gregation/screening. Recently, Jiménez et al. (2012) compared these two
methods and found them comparable with regard to parasite egg recovery.
Currently, the combination of parasite quantification using exotic spores is
combined with either the Lutz spontaneous sedimentation or the disaggrega-
tion screen processing methods by most researchers (Camacho et al., 2013;
Fugassa et al., 2011; Jiménez et al., 2012; Kumm et al., 2010; Martinson
et al., 2003; Reinhard and Bryant, 2008; Reinhard and Urban, 2003; Santoro
et al., 2003; Sianto et al., 2005; Dufour and Le Bailly, 2013).
Bouchet et al. (1999) use ultrasound for disaggregation of rehydrated
material, which must be done carefully to avoid breaking the parasite struc-
tures. This technique is not often recommended, except in particular cases
when there is a lot of mineral aggregate with the coprolite or latrine sedi-
ments. It can also be used for permineralized (fossil) coprolites.
These techniques have been described by Reinhard et al. (1986), Ara ujo
et al. (1998), Bouchet et al. (2003), Silva et al. (2014), Hugot et al. (2014)
and Fugassa (2014), with comments and indications for each of them.

4.2.2 Coprolites in mummies


Some mummies contain coprolites. These can be analysed as described
above for coprolites. However, many mummy coprolites are saturated
with breakdown products that inhibit rehydration. If rehydration does not
occur with trisodium phosphate, we have found that a 4.0% solution of
KOH is effective in dispersing the breakdown products allowing for the
disaggregation of the sample.
It has been our experience that less than half of mummies contain cop-
rolites. Aufderheide (2003) reports that even fewer sub-adult mummies
contain coprolites. The intestinal tract is sometimes completely decomposed
in spontaneously mummified remains. In other mummies, only fragments of
the large intestine remain. When the large intestine is preserved, but is
empty of coprolites, microfossils and macrofossils can still be recovered for
the mummy. The challenge relates to finding a control sample. In our lab-
oratory, we sometimes need to resolve this problem by deriving two samples
from the same intestine: intestine interior and intestine exterior washes
(Searcey et al., 2013). After rehydration in 0.5% trisodium phosphate solu-
tion, we wash the exterior of the intestine and maintain the recovered
microfossils as a ‘control’. We then open the section and wash the interior
to recover parasite remains as well as dietary and medicinal residue. We
368 Adauto Ara
ujo et al.

call this method ‘intestinal wash’. Quantification of microfossils is done by


adding a known number of exotic Lycopodium spores to a known volume
or weight of sediment (Reinhard et al., 2006).
The quantification of the structures found in the process adds important
information to the results, but caution is recommended in the interpreta-
tions. Intestinal helminths and protozoa vary in the deposition or elimina-
tion of parasite structures in the faeces. There are intervals in the
deposition, with enormous variation in the number of forms eliminated
per day. Importantly, the variation in preservation in archaeological sites
or mummified bodies can result in either paucity or abundance of parasite
structures (Piombino-Mascali et al., 2013; Kumm et al., 2010; Shin et al.,
2009). Even in tests in modern patients, helminth egg count is no longer
used to estimate parasite burden, except in specific cases such as schistosomiasis
mansoni (Caldeira et al., 2012).

4.3 Molecular techniques applied to palaeoparasitology


4.3.1 Molecular diagnosis
Firstly, an experimental test is done to determine the extraction standard and
amplification methodology. The most appropriate techniques are thus veri-
fied before applying them to the material from the collection. After testing,
the technique is applied to the selected samples, depending on the parasite to
be investigated.
Preparation of experimental material under different conditions. These
require simulating the state of preservation of the samples to be studied
(desiccation, formalin, alcohol and other procedures).
1. DNA extraction in experimental samples
2. Definition of the appropriate technique and protocols
3. DNA extraction from different samples in the collections
4. Data analysis and comparisons
5. Palaeoparasitological identification and interpretation
6. Selection of protocols and recommended techniques
Depending on the type of preservation of the material, the extraction can
proceed according to the recommendations by Campos and Gilbert (2012).
Multicopy targets, <300 bp, can be prioritized (Poinar et al., 2006). There
are some in-house targets for parasites and hosts, and others have been
described in the literature and used in previous studies by different authors.
For purification, we use the GFX DNA and Gel Band GE Healthcare kit
according to the manufacturer’s protocol.
Palaeoparasitology e Human Parasites in Ancient Material 369

Nucleotide sequencing: amplification products are sequenced directly in


both directions using the automatic sequencer Applied BiosystemsÒ 3130
Genetic Analyser (FIOCRUZ Platform). The products can also be cloned
prior to direct sequencing. To analyse and edit the sequences, we use the
programs Chromas Lite version 2.1 (Technelysium Pty Ltd 2007), BioEdit
v7.1.11.
As exemplified by Dittmar (2009, 2014), novel techniques are able to
recover and analyse ancient DNA (aDNA). As is explained, multiplex
DNA techniques allow one to obtain aDNA sequences of different parasite
species after a single polymerase chain reaction. Other techniques, such as
metagenomics, amplify all DNA of the sample, including contaminants.
Metagenomics has been applied with success by authors (Khairat et al.,
2013; Kay et al., 2014) for recovering DNAs of different parasites from
mummies, and Wood et al. (2013) in coprolites of extinct birds of
New Zealand.

5. PARASITE FINDS IN HUMAN ARCHAEOLOGICAL


REMAINS
5.1 Ascaris lumbricoides and Trichuris trichiura
Two intestinal helminths stand out as the most common in human
archaeological material, whether in coprolites, intestinal content of mummi-
fied bodies or sediment from skeletons (a comprehensive table was published
by Leles et al., 2010). They are Ascaris lumbricoides and Trichuris trichiura, both
perhaps inherited from ancestral humans in Africa (Mitchell, 2013; Ara ujo
and Ferreira, 2014). However, their distribution differs, with the first found
more frequently in Europe, especially during the medieval period, but very
rarely in the New World. Meanwhile, T. trichiura is common in both the
Old and New Worlds (Leles et al., 2010). This created a paradox for the
palaeoepidemiology of Ascaris infection. Why are A. lumbricoides eggs so
rare in New World material when the mechanism of transmission, by the
ingestion of infective eggs, is similar to that of T. trichiura? Although this
question has not been totally elucidated, it was discussed by Leles et al.
(2010), showing that it is possible to detect the presence of the infection
by molecular biology techniques and confirming its presence in prehistoric
New World groups. Even so, it is still rare to find this parasite’s eggs or
genetic material in archaeological sites of the Americas, perhaps due to the
conditions of preservation, since most of the sites in which its presence
370 Adauto Ara
ujo et al.

was diagnosed by recovering genetic material are located in semiarid regions


(Leles et al., 2010).
The discussion on the origin of parasitism by Ascaris lumbricoides in
humans has sparked debate as to whether it originated after the domestica-
tion of swine, thus resulting from the transfer of A. suum and its subsequent
speciation in the human host, or the opposite, that is, humans transmitted
the infection to swine. Recent studies in current and archaeological mate-
rials showed that it is really the same species in both hosts, while A. lumbri-
coides has assumed priority in the scientific nomenclature (Leles et al., 2012).
Some evidence points to an ancient origin of the parasite in the human
host, with findings suggestive of the parasite’s eggs in caves inhabited by
Pleistocene groups (Bouchet et al., 1996) and finds in Africa in coprolites
dated 12,000e9000 years BP (Evans et al., 1996). There is growing evidence
that the two purported Ascaris species circulate between humans and swine,
with minor variations in the haplotypes and indiscriminate transmission
between the two hosts (Betson et al., 2014; Cavallero et al., 2013; Dutto
and Petrosillo, 2013; Vlaminck et al., 2014).
Infection by Ascaris lumbricoides and in certain regions by Trichuris trichiura
appears to have been very abundant in Europe during the medieval period.
In some cases, the eggs from these parasites reach incalculable levels, as in the
case of a Sicilian mummy, in which T. trichiura eggs exceeded the number of
food remains (Kumm et al., 2010; Piombino-Mascali et al., 2013). The same
occurred with A. lumbricoides, although with lower numbers (but still signif-
icant), in the analysis of latrines in medieval Belgium (Fernandes et al., 2005).
Infection by these two intestinal helminths was part of life in medieval
Europe (Reinhard et al., 2013; Reinhard and Pucu, 2014).
A. lumbricoides and Trichuris trichiura dispersed not only in Europe, but
wherever the human hosts migrated, possibly from Africa (Evans et al.,
1996), having been found in various regions of medieval Asia (Matsui
et al., 2003; Han et al., 2003; Shin et al., 2009; Oh et al., 2010; Kim et al.,
2013; Seo et al., 2014a). Both infections reached the Americas during the pre-
historic period (Ara ujo et al., 1981; Ferreira et al., 1980, 1983; Gonçalves
et al., 2003; Leles et al., 2010; Reinhard et al., 1986). The evidence is
more consistent for Trichuris, but we can state that both became cosmopolitan
infections since ancient times.
Ascaris infection can cause severe harm in malnourished children,
including intestinal obstruction and even death from acute septicaemia
due to rupture of the intestine by the formation of a bolus of Ascaris lumbri-
coides in the digestive tract lumen. The Wirsung canal in the pancreas can
Palaeoparasitology e Human Parasites in Ancient Material 371

also become obstructed by erratic penetration of an adult worm, the length


of which can exceed 30e40 cm in the case of females and with a low worm
burden. Adult worms consume a considerable amount of sugar ingested by
the host, which can lead to low blood glucose when the parasite burden is
high, especially in children.
As for Trichuris infection, clinical manifestations are very rare, but can be
complicated in the case of a high parasite burden. In such circumstances, the
helminths extend beyond the ascending colon (where they are normally
located), reaching the final segments of the large intestine, where they cause
a loss of muscle elasticity and prolapsed rectum. This occurs mainly in
malnourished children and the elderly. The individuals themselves often
reduce the prolapse until this becomes too difficult or a bacterial infection
occurs, often with concomitant myiasis. The prolapse is quite peculiar, prob-
ably related to a condition called maculo, described among indigenous people
and even among the European colonizers and African slaves (Rezende,
2003; Bianucci et al., 2015).

5.2 Hookworms
The hookworms that most frequently parasitize humans, Ancylostoma duode-
nale and Necator americanus, contributed to the debate on transpacific contacts
or migrations of Asian populations to the Americas (Ara ujo et al., 1981,
1988, 2008; Ferreira and Ara ujo, 1996; Reinhard et al., 2001). The two spe-
cies have probably accompanied the human host since the first human
migrations from Africa (Ferreira et al., 2014). Like Ascaris lumbricoides and
Trichuris trichiura, the hookworms that parasitize humans dispersed
throughout the globe in regions that allowed maintenance of their life cycle.
However, hookworm egg finds are rare in archaeological material from
Europe and Africa, but are more frequent in both South America and North
America (Gonçalves et al., 2003). These parasites require passage through the
soil for approximately two weeks until their soil cycle is complete, from the
elimination of the eggs in the feces until the release of infective larvae. These
changes require temperatures >20  C and moist soil. Only then are the
larvae capable of penetrating the host and continuing their life cycle. Under
microscopy, the eggs are nearly indistinguishable, but A. duodenale eggs are
slightly smaller (46.4e77.5  23.5e61 mm) than those of N. americanus
(54.2e85.3  30e54 mm), based on statistical analyses (Rep, 1963).
Due to these biological characteristics, the humans that crossed the
Bering land-and-ice bridge could not have introduced hookworm infection
by this route. Prehistoric transpacific migrations or contacts by seafaring
372 Adauto Ara
ujo et al.

Asian populations to the Americas were thus hypothesized (Ara ujo et al.,
1981, 2008; Montenegro et al., 2006).
Infections with high parasite burden, associated with deficiencies in the
human host’s health and nutrition, cause severe cases of ancylostomiasis,
with intense anaemia due to intestinal blood loss through faeces, oedema
of the lower limbs, paleness, weakness, and other signs and symptoms. Texts
by naturalists and chroniclers from the colonial period in the Americas
include descriptions consistent with ancylostomiasis, with such details as
generalized oedema and the habit of eating dirt or clay among indigenous
people (Sousa, n.d.; Ara
ujo et al., 1981). Geophagy is one of the symptoms
of hypochromic microcytic anaemia, associated with hookworm infection
and deficient iron intake, especially in malnourished children and young
people and pregnant women with iron deficiency.

5.3 Enterobius vermicularis


Another nematode, specific to humans and their closest primate relatives,
Enterobius vermicularis (pinworm) has a peculiar life cycle with no need for
passage through the soil for the eggs to be able to infect new hosts. The
mechanism of transmission can occur by ingestion of eggs eliminated in
the faeces and contaminating food or water, by the host’s own hands,
through the air or by autoinfection. Thus, the pinworm cycle is not altered
by high or very low temperatures, but the infection rates are facilitated by
the agglomeration of hosts.
Infection by Enterobius vermicularis originated from African hosts and dates
to prehominids, since it is shared by primates close to humans, such as the
chimpanzee (Pan troglodytes). From Africa, pinworm infection spread to the
rest of the globe, reaching the Americas with the first migrants. E. vermicularis
eggs were found in an Egyptian mummy, proving ancient infection in Africa
(Horne, 2002), with some rare finds in Europe (Herrmann, 1985) and Asia
(Shin et al., 2011), but with most finds concentrated in the Americas (Ferreira
et al., 1997; Hugot et al., 1999).
Based on the biological characteristics of its life cycle, Enterobius vermicu-
laris infection was introduced in the Americas both by prehistoric migrations
across the Bering route (Ferreira et al., 1997) and transpacific contacts of
Asian groups with the Americas (I~ niguez et al., 2003).
Enterobiasis, or oxyuriasis, rarely produces major clinical manifestations,
but there are characteristic symptoms such as anal itching, resulting from the
presence of gravid females that migrate close to the area of oviposition. The
females cause local irritation, leading to pruritus, the intensity of which varies
Palaeoparasitology e Human Parasites in Ancient Material 373

between individuals. More serious clinical cases only occur rarely, caused by
bacterial invasion of the lesions provoked by the itching. Transmission of this
helminthic infection is facilitated by the agglomeration of hosts and is easily
diagnosed in day care centres, nursing homes, and other locations that can
facilitate transmission directly from host to host or by dispersal of eggs in
circulating air or contaminated food, in addition to autoinfection by eggs
carried directly from hand to mouth or the simple penetration of larvae
hatched from the eggs deposited in the perianal region. Prehistoric infection
in North America was very well documented by Hugot et al. (1999), study-
ing this parasite among the Ancestral Puebloans that inhabited the dwellings
built of clay and wood in the rock shelters and caves in the canyons, which
were clustered and with intercommunications. There were also collective
rooms with directed air circulation, which facilitated aerosol transmission
of the parasite’s eggs.

5.4 Diphyllobothrium sp.


Diphyllobothriasis was also an emblematic parasitic infection in certain
regions of the prehistoric Americas, as well as in Europe during the Neolithic
(Le Bailly et al., 2005). There are various finds of eggs from Diphyllobothrium
latum and related species of this cestode in Europe, showing infection asso-
ciated with the ingestion of raw or poorly cooked lacustrine fish. This
zoonosis became rare after the Middle Ages, certainly due to changes in
dietary customs and sanitary control developed in the countries (Bouchet
and Le Bailly, 2014).
Along the Pacific Coast of South America, especially among peoples in
the Atacama Desert, diphyllobothriasis has persisted since prehistory. There
are various records of Diphyllobothrium pacificum eggs in the prehistoric coastal
population, especially in Chile and Peru. This parasite’s cycle occurs in crus-
taceans and saltwater fish, showing that marine resources were used
frequently, even among groups located far from the sea (Ara ujo et al.,
1983). The habit of eating raw fish has persisted from prehistory till 2015.
Parasitologist Jean Baer recorded this species in Peruvian population in
2015 when he was called to investigate an outbreak of diarrhoea, vomiting
and abdominal pain in the 1960s (Baer et al., 1967). Importantly, he referred
to the antiquity of this parasite infection after visiting the Museum of
Archaeology in Lima, where the collection includes pottery similar to the
current ceramic ware for serving ceviche, a typical dish made of raw fish
and spices. Proof of the observation by Baer et al. (1967) came when the
parasite eggs were found in human coprolites from the Atacama Desert
374 Adauto Ara
ujo et al.

(Ara
ujo et al., 1983). It was later shown that the infection was quite common
in the different human groups that inhabited the region (Reinhard and
Urban, 2003).

6. OTHER PARASITES: PARASITES OF ANIMALS


FOUND IN HUMAN COPROLITES; PARASITES IN
PREHISTORIC ASIA
Prehistoric parasitic infections in Asia displayed a different profile from
that of other continents. Although parasitism by Ascaris lumbricoides and Tri-
churis trichiura was also prevalent, infections with parasites of fish in humans
were quite common (Seo et al., 2008). These findings show a dependency
on (or preference for) eating raw saltwater fish, sometimes leading to impor-
tant zoonotic infections in the population. There are reports of significant
changes between countries, for example, when Ascaris and Trichuris
infections were introduced from China into Japan along with rice farming
(Matsui et al., 2003).
There are interesting finds of parasites of animals in human coprolites.
Parasites of animals found in human faeces identify the consumption of
foods of animal origin, ingested raw or poorly cooked, where the parasite
eggs can cross the digestive tract without infecting the human host. Mean-
while, some parasites of animals infect humans and can cause diseases. The
most common infection of this type in Europe and in certain regions of
the Americas is diphyllobothriasis, but other parasites of fish consumed tradi-
tionally in Asia are also frequently found in archaeological remains. They
reveal the persistence of a diet based on raw or poorly cooked fish from pre-
history to the present.
False parasitism, or the presence of eggs or other parasite structures in
human faeces and pertaining to parasites that are not capable of infecting
the human host, can indicate the consumption of raw or poorly cooked
animals, or even plants, as in the case of eggs from Meloidogyne sp., a parasite
of plants found in human coprolites (unpublished). In some cases, it is
possible to identify the species or group of animals that served as food.
This was the case described by Sianto et al. (2012), showing that the inges-
tion of small lizards or geckos was common in prehistoric Brazil. Eggs of
Pharyngodonidae, a parasite of lizards, were found in human coprolites,
which allowed a discussion on the ingestion, in prehistory, of live or recently
killed geckos. This is still a habit among isolated human groups exposed to
food shortages in semiarid regions of Brazil’s hinterlands (Sianto et al., 2012).
Palaeoparasitology e Human Parasites in Ancient Material 375

7. ORIGIN AND EVOLUTION OF TRYPANOSOMATIDS


IN HUMANS AND THE PARADIGM SHIFT FROM
RESULTS IN PALAEOPARASITOLOGY
Trypanosoma cruzi infection occurs among insect vectors, wild and
domestic animals and humans. It is considered a zooanthroponosis (Coura
and Dias, 2009; Coura and Borges-Pereira, 2010, 2012), since the parasite
circulated primarily only among wild animals. Humans encroached on the
parasite’s cycle when they reached the Americas and occupied natural foci
of the T. cruzi life cycle (Ara
ujo et al., 2009).
According to the classical theory on the origin of Chagas disease, it
began after the domestication of small rodents from genus Cavia (or ‘cuyes’)
in the Bolivian altiplano, spreading from there southward and reaching
what is now Brazil much later (Pinto-Dias, 2013). Based on recent
data from palaeoparasitology, we now know that the parasite already in-
fected humans 7000 years ago in central South America, in the Brazilian
cerrado, including the form of the disease characterized by digestive tract
lesions (Lima et al., 2008), and in the Cohuailla Desert along the border bet-
ween the southern United States and northern Mexico (Reinhard et al.,
2003).
The human groups that used the same caves where infected vectors and
animals were found were exposed to the vectors and to maintenance of the
wild cycle. American trypanosomiasis appears to have a very ancient origin,
and according to Poinar (2014) its life cycle has occurred in triatomines and
mammals since the Cenozoic. The occurrence of Trypanosoma antiquae in tri-
atomines found in Dominican amber (Poinar, 2005), now dated to the
Miocene (Iturralde-Vinent and MacPhee, 1996; Iturralde-Vinent, 2001),
and associated with bat hair indicates the antiquity of the infection and prob-
ably its origin in these mammals (Poinar, 2014). Other Trypanosomatidae
may potentially be traced back to the Cretaceous Burmese Amber (Poinar
and Poinar, 2004a,b; Poinar, 2007).
Infection in humans probably dates to soon after their arrival in the
Americas (Ferreira et al., 2011), wherever they occupied existing transmis-
sion foci among wild mammals (Ara ujo et al., 2009). With the increasing
transformation of spaces by humans and their domestic animals, new adap-
tations of the parasite to new hosts occurred, as well as to new habitats, in a
process called domiciliation (Lardeux, 2013). T. cruzi is known as a single
species, but with different lineages that were characterized over the course
of this long evolutionary process (Guhl et al., 2014).
376 Adauto Ara
ujo et al.

Findings show that the infection circulated among prehistoric groups


that inhabited the rock shelters and caves, or wherever they created favour-
able conditions for colonization by triatomines (Ara ujo et al., 2009), in both
North and South America, in various biomes where the vector and other
animal hosts were found. The occurrence of lesions in the digestive tract
(Lima et al., 2008; Reinhard et al., 2003) and heart (Fornaciari et al.,
1992; Rothhammer et al., 1985) has raised speculation on the repercussions
or impact of the disease in prehistoric groups, which included small bands of
hunter-gatherers. Sudden death from cardiac arrhythmia, especially among
young and apparently healthy individuals, would have had a significant
impact on nomadic hunter-gatherer groups. Meanwhile, the chronic man-
ifestations of Chagas disease, such as heart failure and the development of
megas (extensive dilations) in the digestive tract, would have caused limita-
tions in the individuals, with the need for special care, as in the case of the
mummy from Coahuaila (Reinhard et al., 2003).
Further in relation to kinetoplastids, little is known about the origins of
infections by species from genus Leishmania in humans. There are various
gaps in the knowledge, especially in relation to asymptomatic cases, often
more numerous than symptomatic ones (Marzochi et al., 1985). The origins
of Leishmania species purportedly resulted from a process of speciation in the
Old World (Tuon et al., 2008). As with genus Trypanosoma, a putative Leish-
mania species has been found in vectors preserved in amber (Poinar and
Poinar, 2004a,b). An international research project is under way, coordi-
nated in Italy, on the origins and evolution of Leishmania species (Nerlich
et al., 2012).

8. ECTOPARASITES
This group of parasites includes different classes of arthropods and
some other invertebrates. The most common classes in humans are Insecta
and Arachnida (mites and ticks). Some of these ectoparasites, such as head
lice Pediculus humanus, are also shared with nonhuman primates, and thus
likely infested humans since the prehominids (Araujo et al., 2003; Mitchell,
2013). According to Reed et al. (2004) and Raoult et al. (2008), lice have
parasitized primates for some 25 million years, and when chimpanzees
diverged from humans 6 million years ago, genus Pediculus remained in
both species. This parasite has been found in Egyptian mummies and in
archaeological sites in North America and South America. Infestation
became as widespread in the Andean region (Arriaza et al., 2012; Dutra
Palaeoparasitology e Human Parasites in Ancient Material 377

et al., 2014) as in medieval Europe, where the hairdos of the nobility were
veritable breeding grounds for lice (Fornaciari et al., 2009). The same was
true of the refined hairdos of the Tiwanaku: when the Incas arrived as con-
querors of other Andean empires, they levied ‘lice taxes’, now interpreted
as a way to control infestation or simply as a form of human head count
(Souffez, 2001).
There is extensive research on the origin and evolution of pediculosis.
Reed et al. (2004) and Raoult et al. (2008) provided important evidence
on the antiquity and evolution of this parasite infection in humans, with
the oldest finding thus far in the Americas (Ara ujo et al., 2000). Studies
on the meaning and impact of mass lice infestations were conducted in
Andean populations (Dutra et al., 2014; Arriaza et al., 2012, 2013b), where
exceedingly well-preserved mummified bodies are found, including well-
preserved hair.
The preservation of eggs or lice is so extraordinary that they were found
on human shrunken heads from the Jívaro, an indigenous group on the
border of Brazil with Ecuador, dated to the colonial period and the Second
Empire in Brazil, in the nineteenth century (Ara ujo et al., 2005). Eggs of
ectoparasites have in rare cases even been found on dinosaur feathers or
mammal hair preserved in amber (e.g. Martill and Davis, 1998; Poinar,
2014). Both lice (Order Phthiraptera) and fleas (Order Siphonaptera), which
currently parasitize bird and mammals, can be traced back to the Cretaceous
at least (see the review in Nagler and Haug, 2015).

9. CONCLUSIONS
Palaeoparasitology contributes new data to knowledge on the origin
and evolution of parasitic diseases. It situates e in time and space e the pres-
ence of infection in human groups that have already disappeared, situating
their distribution and tracing prehistoric migrations across the continents
through parasites preserved in the occupation sites.
Palaeoparasitology has also contributed greatly to studies on the evolu-
tion of parasites and their hosts. Despite previous claims that parasites do
not leave fossils (e.g. Dorris et al., 1999), parasite finds have become frequent
in remains from extinct animals dating back millions of years (Littlewood
and Donovan, 2003; Poinar, 2014), particularly those preserved in amber
and in coprolites found in palaeontological sites (Poinar and Boucot,
2006; Dentzien-Dias et al., 2013). It is thus possible to study the presence
378 Adauto Ara
ujo et al.

of parasites millions of years old and their relations with current species, as in
the case of Oxyurid and Ascarid eggs found in cynodonts, dated 240 million
years BP (Silva et al., 2014; Hugot et al., 2014). This group of parasites has
persisted to 2015, sharing a long evolutionary history with its hosts.
The development and growing interest of research groups in palaeopar-
asitology have stimulated progress in knowledge on parasiteehosteenviron-
ment relations, reinforcing the ecological approach that should be
encouraged in studies on parasites and hosts.

ACKNOWLEDGEMENTS
The research received funding from the Brazilian National Council on Scientific and Tech-
nological Development (CNPq), Coordinating Division for the Advancement of Graduate
Studies (CAPES) in the Science without Borders programme (Ciencia sem Fonteiras), Carlos
Chagas Filho Rio de Janeiro State Research Foundation (FAPERJ) and InovaENSP (Sergio
Arouca National School of Public Health).

REFERENCES
Anastasiou, E., Mitchell, P.D., 2013. Human intestinal parasites from a latrine in
the 12th century Frankish castle of Saranda Kolones in Cyprus. Int. J. Paleopathol.
3, 218e223.
Ara
ujo, A., Ferreira, L.F., Confalonieri, U., 1981. A contribution to the study of helminth
findings in archaeological material in Brazil. Rev. Bras. Biol. 41, 873e881.
ujo, A., Ferreira, L.F., Confalonieri, U., Nunez, L., 1983. Eggs of Diphyllobothrium pacif-
Ara
icum in pre-Columbian human coprolites. Paleopathol. News 41, 11e13.
Ara
ujo, A., Ferreira, L.F., Confalonieri, U., Chame, M., 1988. Hookworm and the peopling
of America. Cad. Sa ude Publica 2, 226e233.
Ara
ujo, A., Reinhard, K., Bastos, O.M., Costa, L.C., Pirmez, C., Iniguez, A.M., Vicente, A.C.,
Morel, C.M., Ferreira, L.F., 1998. Paleoparasitology: perspectives with new techniques.
Rev. Inst. Med. Trop. S~ao Paulo 40, 371e376.
Araujo, A., Ferreira, L.F., Guidon, N., Freire, N.M.S., Reinhard, K., Dittmar, K., 2000. Ten
thousand years of head lice infection. Parasitol. Today 16, 269e270.
Ara
ujo, A., Jansen, A.M., Bouchet, F., Reinhard, K., Ferreira, L.F., 2003. Parasites,
the diversity of life, and paleoparasitology. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1),
5e11.
Ara
ujo, A., Mendonça de Souza, S.M., Nascimento, F., Reinhard, K., 2005. The Jivaro
shrunken heads from the National Museum, Rio de Janeiro, Brazil: authentic or
counterfeits. J. Biol. Res. 80, 152e157.
Ara
ujo, A., Reinhard, K., Ferreira, L.F., Gardner, S., 2008. Parasites as probes for prehistoric
migrations? Trends Parasitol. 24, 112e115.
Ara
ujo, A., Jansen, A.M., Reinhard, K., Ferreira, L.F., 2009. Paleoparasitology of Chagas dis-
ease: a review. Mem. Inst. Oswaldo Cruz 104, 9e16.
Ara
ujo, A., Ferreira, L.F., 2014. Parasitism. In: Ferreira, L.F., Reinhard, K., Ara
ujo, A. (Eds.),
Foundations of Paleoparasitology. Editora Fiocruz/International Federation of Tropical
Medicine, pp. 111e119.
Arriaza, B.T., Reinhard, K.J., Ara ujo, A., Orellana, N.C., Standen, V.G., 2010. Possible
influence of the ENSO phenomenon on the pathoecology of diphyllobothriasis and ani-
sakiasis in ancient Chinchorro populations. Mem. Inst. Oswaldo Cruz 105, 66e72.
Palaeoparasitology e Human Parasites in Ancient Material 379

Arriaza, B., Orellana, N., Barbosa, H.S., Menna Barreto, R., Ara ujo, A., Standen, V., 2012.
Severe head lice infestation in an Andean mummy of Arica, Chile. J. Parasitol. 98,
433e436.
Arriaza, B., Standen, V., Reinhard, Karl, Ara ujo, A., Heukelbach, J., Dittmar, K., 2013a. On
head lice and social interaction in archaic Andean coastal populations. Int. J. Paleopathol.
10, 1e15.
Arriaza, B., Standen, V., N ~ez, H., Reinhard, K., 2013b. Study of archaeological nits/eggs
un
of Pediculus humanus capitis by scanning electron microscopy. Micron 45, 145e149.
Aufderheide, A., 2003. The Scientific Study of Mummies. Cambridge University Press, p. 320.
Baer, J.G., Miranda, H., Fernandez, W., Medina, J., 1967. Human diphyllobothriasis in Peru.
Z. Parasitenkd 28, 277e289.
Barrow Jr., J.H., Stockton, J.J., 1960. The influences of temperature on the hosteparasite
relationships of several species of snakes infected with Entamoeba invadens. J. Protozool.
7, 377e383.
Baum, J., Bat-Gal, G.K., 2003. The emergence and co-evolution of human pathogens. In:
Greenblatt, C., Spigelman, M. (Eds.), Emerging Pathogens. Oxford University Press,
Oxford, pp. 67e78.
Betson, M., Nejsum, P., Bendall, R.P., Deb, R.M., Stothard, J.R., 2014. Molecular epide-
miology of ascariasis: a global perspective on the transmission dynamics of Ascaris in peo-
ple and pigs. J. Infect. Dis. 15, 932e941.
Bianucci, R., Torres, E.J.L., Dutra, J.M.F.D., Ferreira, L.F., Nerlich, A.G., Souza, S.M.M.,
Giuffra, V., Chieffi, P.P., Bastos, O.M., Travassos, R., Souza, W., Ara ujo, A., 2015.
Trichuris trichiura in a post-Colonial Brazilian mummy. Mem. Inst. Oswaldo Cruz 110,
145e147.
Bouchet, F., Bafier, D., Girard, M., Morel, P., Paicheler, J.C., David, F., 1996. Paléopara-
sitologie en contexte Pleistocene: premiere observations a la Grande Grotte d’Arcy-sur-
Cure (Yonne), France. C.R. Acad. Sci. III 319, 147e151.
Bouchet, F., Lefréve, C., West, D., Corbert, D., 1999. First paleoparasitological
analysis of a midden in the Aleutian Islands (Alaska): results and limits. J. Parasitol.
85, 369e372.
Bouchet, F., Guidon, N., Dittmar, K., Harter, S., Ferreira, L.F., Chaves, S.M., Reinhard, K.,
Araujo, A., 2003. Parasite remains in archaeological sites. Mem. Inst. Oswaldo Cruz 98
(Suppl.1), 47e52.
Bouchet, F., Le Bailly, M., 2014. The findings in Europe. In: Ferreira, L.F., Reinhard, K.,
Araujo, A. (Eds.), Foundations of Paleoparasitology. Ed. Fiocruz/Int. Fed. Tropical
Medicine, pp. 363e388.
Brumpt, E., 1913. Précis de Parasitologie. Masson, Paris, p. 9.
Bryant Jr., V.M., 1974. Prehistoric diet in southwest Texas: the coprolite evidence. Am.
Antiq. 39, 407e420.
Bryant Jr., V.M., Reinhard, K.J., 2012. Coprolites and archaeology: the missing links in
understanding human health. In: Hunt, A. (Ed.), Vertebrate Coprolites. New Mexico
Museum of Natural History and Science Bulletin 51.
Buikstra, J.E., Beck, J., 2008. In: Buikstra, Jane E., Beck, Lane A. (Eds.), The Contextual
Analysis of Human Remains. Academic Press/Elsevier, 347e258.
Buikstra, J., 2013. EIC editorial: introducing the paleoparasitology issue. Int. J. Paleopathol.
3, 139.
Caldeira, K., Teixeira, C.F., Silveira, M.B., Fries, L.C.C., Romanzini, J., Bittencourt, H.R.,
Graeff-Teixeira, C., 2012. Comparison of the Kato-Katz and Helmintex methods for the
diagnosis of schistosomiasis in a low-intensity transmission focus in Bandeirantes, Parana,
southern Brazil. Mem. Inst. Oswaldo Cruz 107, 690e692.
Callen, E.O., Cameron, T.W.M., 1960. A prehistoric diet revealed in coprolites. New Sci. 8,
35e40.
380 Adauto Ara
ujo et al.

Camacho, M., Pessanha, T., Leles, D., Dutra, J., Silva, R., Mendonça de Souza, S.M.,
Ara ujo, A., 2013. Lutz’ spontaneous sedimentation technique and the paleoparasitological
analysis of sambaqui (shellmounds) sediments. Mem. Inst. Oswaldo Cruz 108, 155e159.
Campos, P.F., Gilbert, T.M., 2012. DNA extraction from formalin-fixed material. Methods
Mol. Biol. 840, 81e85.
Cavallero, S., Snabel, V., Pacella, F., Perrone, V., D’Amelio, S., 2013. Phylogeographical
studies of Ascaris spp. based on ribosomal and mitochondrial DNA sequences. PLoS
Negl. Trop. Dis. 7, e2170. http://dx.doi.org/10.1371/journal.pntd.0002170.
Cockburn, A., 1963. The Evolution and Eradication of Infectious Diseases. The Johns Hop-
kins Press, Baltimore, pp. 255.
Coura, J.R., 1997. Síntese historica e evoluç~ao dos conhecimentos sobre a doença de Chagas.
In: Coura, J.R., Dias, J.C.P. (Eds.), Clínica e Terapênutica da Doença de Chagas. Editora
Fiocruz, pp. 469e486.
Coura, J.R., Dias, J.C., 2009. Epidemiology, control and surveillance of Chagas disease:
100 years after its discovery. Mem. Inst. Oswaldo Cruz 104 (Suppl. 1), 31e40.
Coura, J.R., Borges-Pereira, J., 2010. Chagas disease: 100 years after its discovery. A systemic
review. Acta. Trop. 115, 5e13.
Coura, J.R., Borges-Pereira, J., 2012. Chagas disease. What is known and what should be
improved: a systemic review. Rev. Soc. Bras. Med. Trop. 45, 286e296.
Dentzien-Dias, P.C., Poinar Jr., G., de Figueiredo, A.E., Pacheco, A.C., Horn, B.L.,
Schultz, C.L., 2013. Tapeworm eggs in a 270 million-year-old shark coprolite. PLoS
One 8, e55007. http://dx.doi.org/10.1371/journal.pone.0055007.
Diamond, J., 1987. The worst mistake in the history of the human race. Discov. Mag. 64e66.
Dittmar, K., 2009. Old parasites for a new world: the future of paleoparasitological research.
A review. J. Parasitol. 95, 365e371.
Dittmar, K., Ara ujo, A., Reinhard, K., 2012. The study of parasites through time: archaeo-
parasitology and paleoparasitology. In: Grauer, A.L. (Ed.), A Companion to
Paleopathology. Wiley-Blackwell, New Jersey, pp. 170e190.
Dittmar, K., 2013. Guest editorial. Int. J. Paleopathol. 3, 140e141.
Dittmar, K., 2014. Paleoparasitology and ancient DNA. In: Ferreira, L.F., Reinhard, K.,
Ara ujo, A. (Eds.), Foundations of Paleoparasitology. Editora Fiocruz/International
Federation of Tropical Medicine, pp. 277e288.
Dorris, M., De Ley, P., Blaxter, M.L., 1999. Molecular analysis of nematode diversity and the
evolution of parasitism. Parasitol. Today 15, 188e193.
Dufour, B., Le Bailly, M., 2013. Testing new parasite egg extraction methods in paleopara-
sitology and an attempt at quantification. Int. J. Paleopathol. 3, 199e203.
Dutra, J.M., Alves, A.D., Pessanha, T., Rachid, R., Souza, W., Linardi, P.M., Ferreira, L.F.,
Mendonça de Souza, S.M., Ara ujo, A., 2014. Prehistorical Pediculus humanus capitis infes-
tation: quantitative data and low vacuum scanning microscopy. Rev. Inst. Med. Trop.
S~ao Paulo 56, 115e119.
Dutto, M., Petrosillo, N., 2013. Hybrid Ascaris suum/lumbricoides (Ascarididae) infestation in a
pig farmer: a rare case of zoonotic ascariasis. Centr. Eur. J. Public Health 21, 224e226.
Evans, A.C., Markus, M.B., Mason, R.J., Steel, R., 1996. Late stone-age coprolite reveals
evidence of prehistoric parasitism. S. Afr. Med. J. 86, 274e275.
Ewald, P.W., Swain Ewald, H.A., 2014. Joint infectious causations of human cancer. Adv.
Parasitol. 84, 1e26.
Faulkner, C.T., Reinhard, K.J., 2014. A retrospective examination of paleoparasitology and
its establishment in the Journal of Parasitology. J. Parasitol. 100, 253e259.
Fernandes, A., Ferreira, L.F., Gonçalves, M.L.C., Bouchet, F., Klein, C.H., Iguchi, T.,
Sianto, L., Ara ujo, A., 2005. Intestinal parasite analysis in organic sediments collected
from a 16th Century Belgium archaeological site. Cad. Sa ude Publica 21, 329e332.
Palaeoparasitology e Human Parasites in Ancient Material 381

Fernandes, A., I~ niguez, A.M., Lima, V.S., Souza, S.M., Ferreira, L.F., Vicente, A.C.,
Jansen, A.M., 2008. Pre-Columbian Chagas disease in Brazil: Trypanosoma cruzi I in the
archaeological remains of a human in Peruaçu Valley, Minas Gerais, Brazil. Mem. Inst.
Oswaldo Cruz 103, 514e516.
Ferreira, L.F., Araujo, A., Confalonieri, U., 1979. Subsidios para a Paleoparasitologia do Bra-
sil I - Parasitos encontrados em copr olitos no município de Unai, Minas Gerais. In:
Congr. Bras. Parasitol., Abstr.,Campinas, SP, p. 56.
Ferreira, L.F., Araujo, A.J.G., Confalonieri, U.E.C., 1980. The finding of eggs and larvae of
parasitic helminths in archaeological material from Unai, Minas Gerais, Brazil. Trans. R.
Soc. Trop. Med. Hyg. 74, 798e800.
Ferreira, L.F., Araujo, A., Confalonieri, U.E.C., 1983. The finding of helminth eggs in a Bra-
zilian mummy. Trans. R. Soc. Trop. Med. Hyg. 77, 65e67.
Ferreira, L.F., Araujo, A., Confalonieri, U., Nu~ nez, L., 1984. The finding of eggs of Diphyl-
lobothrium pacificum in human coprolites (4100-1950 B.C.) from northern Chile. Mem.
Inst. Oswaldo Cruz 79, 175e180.
Ferreira, L., Araujo, A., 1996. On hookworms in the Americas and trans-pacific contact. Par-
asitol. Today 12, 454e455.
Ferreira, L.F., Reinhard, K., Ara ujo, A., Coura, L.C., 1997. Paleoparasitology of oxyuriasis.
Acad. Nac. Med. 157, 20e24.
Ferreira, L.F., Jansen, A.M., Ara ujo, A., 2011. Chagas disease in prehistory. Acad. Bras. Ci.
83, 1041e1044.
Ferreira, L.F., 2014. An introduction to paleoparasitology. In: Ferreira, L.F., Reinhard, K.,
Ara ujo, A. (Eds.), Foundations of Paleoparasitology. Editora Fiocruz/International
Federation of Tropical Medicine, pp. 27e41.
Ferreira, L.F., Reinhard, K.J., Ara ujo, A., 2014. The origin of parasites of humans. In:
Ferreira, L.F., Reinhard, K., Ara ujo, A. (Eds.), Foundations of Paleoparasitology. Editora
Fiocruz/International Federation of Tropical Medicine, pp. 121e139.
Fischer, C.L., Reinhard, K., Kirk, M., DiVirgilio, J., 2007. Privies and parasites: the archae-
ology of health conditions in Albany, New York. Hist. Archaeol. 41, 172e197.
Florenzano, A., Mercuri, A.M., Pederzoli, A., Torri, P., Bosi, G., Olmi, L., Rinaldi, R.,
Mazzanti, M.B., 2012. The significance of intestinal parasite remains in pollen samples
from medieval pits in the Piazza Garibaldi of Parma, Emilia Romagna, Northern Italy.
Geoarchaeology 27, 34e47.
Fornaciari, G., Castagna, M., Viacava, P., Tognetti, A., Bevilacqua, G., Segura, E.L., 1992.
Chagas’ disease in Peruvian Inca mummy. Lancet 339, 128e129.
Fornaciari, G., Giuffra, V., Marinozzi, S., Picchi, M.S., Masetti, M., 2009. ‘Royal’ pediculosis
in Renaissance Italy: lice in the mummy of the King of Naples Ferdinand II of Aragon
(1467e1496). Mem. Inst. Oswaldo Cruz 104, 671e672.
Frías, L., Leles, D., Ara ujo, A., 2013. Studies on protozoa in ancient remains e a review.
Mem. Inst. Oswaldo Cruz 108, 1e12.
Fry, G.F., 1970. Preliminary analysis of the Hogup Cave coprolites. In: Aïkens, M. (Ed.),
Hogup Cave, Appendix III. University of Utah, Salt Lake City, pp. 247e250 (Anthro-
pological Papers) 93.
Fugassa, M.H., Ara ujo, A., Guichon, R.A., 2006. Quantitative paleoparasitology applied to
archaeological sediments. Mem. Inst. Oswaldo Cruz 101 (Suppl. 2), 29e33.
Fugassa, M.H., Taglioretti, V., Gonçalves, M.L., Ara ujo, A., Sardella, N.H., Denegri, G.M.,
2008. Capillaria spp. eggs in Patagonian archaeological sites: statistical analysis of morpho-
metric data. Mem. Inst. Oswaldo Cruz 103, 104e105.
Fugassa, M.H., Reinhard, K.J., Johnson, K.L., Gardner, S.L., Vieira, M., Ara ujo, A., 2011.
Parasitism of prehistoric humans and companion animals from Antelope Cave, Mojave
County, Northwest Arizona. J. Parasitol. 97, 862e867.
382 Adauto Ara
ujo et al.

Fugassa, M., 2014. Paleoparasitological diagnosis. In: Ferreira, L.F., Reinhard, K., Ara ujo, A.
(Eds.), Foundations of Paleoparasitology. Ed. Fiocruz/World Federation of Tropical
Medicine, Rio de Janeiro, pp. 223e254.
Gomez i Prat, J., Mendonça de Souza, S.M.F., 2003. Prehistoric tuberculosis in America:
adding comments to a literature review. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1),
151e159.
Gonçalves, M.L.C., Ara ujo, A., Ferreira, L.F., 2003. Human intestinal parasites in the
past: new findings and a review. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 103e118.
Gonçalves, M.L.C., Silva, V.L., Andrade, C.M., Reinhard, K., Rocha, G.C., Le Bailly, M.,
Bouchet, F., Ferreira, L.F., Araujo, A., 2004. Amoebiasis distribution in the past: first
steps using a immunoassay technique. Trans. R. Soc. Trop. Med. Hyg. 98, 88e91.
Gooch, P.S., 1972. Helminths in Archaeological and Prehistoric Deposits. Commonwealth
Institute of Helminthology, St Albans, UK. Annotated Bibliography, N 9.
Grmek, M.D., 1983. Les Maladies a l’Aube de la Civilisation Occidentale. Payot, Paris.
Guhl, F., Auderheide, A., Ramírez, J.D., 2014. From ancient to contemporary molecular eco-
epidemiology of Chagas disease in the Americas. Int. J. Parasitol. 44, 605e612.
Han, E.T., Guk, S.M., Kim, J.L., Jeong, H.J., Kim, S.N., Chai, J.Y., 2003. Detection of para-
site eggs from archaeological excavations in the Republic of Korea. Mem. Inst. Oswaldo
Cruz 98 (Suppl. 1), 123e126.
Harper, K.N., Ocampo, P.S., Steiner, B.M., George, R.W., Silverman, M.S., Bolotin, S.,
Pillay, A., Saunders, N.J., Armelagos, G.J., 2008. On the origin of the Treponematoses:
a phylogenetic approach. PLoS Negl. Trop. Dis. e148. http://dx.doi.org/10.1371/
journal.pntd.0000148.
Heizer, R.F., Napton, L.K., 1969. Biological and cultural evidence from prehistoric human
coprolites. Science 165, 563e568.
Herrmann, B., 1985. Parasitologisch-epidemiologische auswertungen mit- telalterlicher
kloaken. Z. Archaol. Mittelalt. 13, 131e161.
Hevly, R.H., Kelly, R.E., Anderson, G.A., Olsen, S.J., 1979. Comparative effects of climate
change, cultural impact, and volcanism in the paleoecology of Flagstaff, Arizona, A.D.
900e1300. In: Sheets, E., Grayson, D. (Eds.), Volcanic Activity and Human History.
Academic Press, New York, pp. 487e523.
Horne, P.D., 2002. First evidence of enterobiasis in ancient Egypt. J. Parasitol. 88, 1019e1021.
Hugot, J.P., Reinhard, K., Gardner, S.L., Morand, S., 1999. Human enterobiasis in evolu-
tion: origin, specificity and transmission. Parasite 6, 201e208.
Hugot, J.P., Gardner, S.L., Borba, V.H., Ara ujo, P., Leles, D., Da-Rosa, A., Dutra, J.,
Ferreira, L.F., Ara
ujo, A., 2014. Discovery of a 240 million year old nematode parasite
egg in a cynodont coprolite sheds light on the early origin of pinworms in vertebrates.
Parasit. Vectors 7, 486. http://dx.doi.org/10.1186/s13071-014-0486-6.
I~
niguez, A.M., Reinhard, K., Ara ujo, A., Ferreira, L.F., Vicente, A.C.P., 2003. Enterobius ver-
micularis: ancient DNA from North and South American human coprolites. Mem. Inst.
Oswaldo Cruz 98 (Suppl. 1), 67e69.
Iturralde-Vinent, M.A., MacPhee, R.D.E., 1996. Age and paleogeographical origin of
dominican amber. Science 273, 1850e1852.
Iturralde-Vinent, M.A., 2001. Geology of the amber-bearing deposits of the greater Antilles.
Carribean J. Sci. 37, 141e166.
Jaeger, L.H., Mendonça de Souza, S.M., Dias, O.F., I~ niguez, A.M., 2013. Mycobacterium
tuberculosis complex in remains of 18th-19th century slaves. Braz. Emerg. Infect. Dis.
19, 837e839.
Jiménez, A.F., Gardner, S.L., Ara ujo, A., Fugassa, M., Brooks, R.H., Racz, E., Reinhard, K.,
2012. Zoonotic and human parasites of inhabitants of Cueva de los Muertos Chiquitos,
Rio Zape Valey, Durango, Mexico. J. Parasitol. 98, 304e309.
Palaeoparasitology e Human Parasites in Ancient Material 383

Jones, A.K.J., 1982. Recent finds of intestinal parasite ova at York, England. In: Presented at
the Annual Meeting of the European Chapter of the Paleopathology Association.
Jones, A.K.G., 1985. Trichurid ova in archaeological deposits: Their value as indicators of
ancient feces. In: Fieller, N.J.R., Gilbertson, D.D., Ralph, N.G.A. (Eds.), Paleobiological
investigations: Research design, methods and data analysis, BAR International Series
266. British Archaeological Reports, Oxford, pp. 105e114.
Kay, G.L., Sergeant, V.G., Bandiera, P., Milanese, M., Bramanti, B., Bianucci, R.,
Pallen, M.J., 2014. Recovery of a medieval Brucella melitensis genome using shotgun
metagenomics. MBio 5, e01337-14. http://dx.doi.org/10.1128/mBio.01337-14.
Khairat, R., Ball, M., Chang, C.C., Bianucci, R., Nerlich, A.G., Trautman, M., Ismail, S.,
Shanab, G.M., Karim, A.M., Gad, Y.Z., Pusch, C.M., 2013. First insights into the meta-
genome of Egyptian mummies using next-generation sequencing. J. Appl. Genet. 54,
309e325.
Kim, M.J., Shin, D.H., Song, M.J., Song, H.Y., Seo, M., 2013. Paleoparasitological surveys
for detection of helminth eggs in archaeological sites of Jeolla-do and Jeju-do. Korean J.
Parasitol. 51, 489e492.
Kristjansdottir, S., Collins, C., 2011. Cases of hydatid disease in Medieval Iceland. Int. J.
Osteoarchaeol. 21, 479e486.
Kuijper, W.J., Turner, H., 1992. Diet of a Roman centurion at Alphen aan den Rijn, The
Netherlands, in the first century AD. Rev. Palaeobot. Palynol. 73, 187e204.
Kumm, K., Reinhard, K., Piombino, D., Ara ujo, A., 2010. Archaeoparasitological investiga-
tion of a mummy from Sicily (18the19th century AD). Anthropol. Int. J. Sci. Ma. 48,
177e184.
Lardeux, F., 2013. Niche invasion, competition and coexistence amongst wild and domestic
Bolivian populations of Chagas vector Triatoma infestans (Hemiptera, Reduviidae,
Triatominae). C. R. Biol. 336, 183e193.
Le Bailly, M., Leuzinger, U., Schlichtherle, H., Bouchet, F., 2005. Diphyllobothrium: neolithic
parasite? J. Parasitol. 91, 957e959.
Le Bailly, M., Gonçalves, M.L.C., Harter-Lailheugue, S., Prodéo, F., Ara ujo, A.,
Bouchet, F., 2008. New finding of Giardia intestinalis (Eukaryote, Metamonad) in Old
World archaeological site using immunofluorescence and enzyme-linked immunosor-
bent assays. Mem. Inst. Oswaldo Cruz 103, 298e300.
Le Bailly, M., Bouchet, F., 2013. Diphyllobothrium in the past: review and new records. Int. J.
Paleopathol. 3, 182e187.
Leles, D., Reinhard, K., Fugassa, M., Ferreira, L.F., I~ niguez, A.M., Ara ujo, A., 2010. A para-
sitological paradox: why is ascarid infection so rare in the prehistoric Americas?
J. Archaeol. Sci. 37, 1510e1520.
Leles, D., Gardner, S.L., Reinhard, K., I~ niguez, A.M., Ara ujo, A., 2012. Are Ascaris lumbri-
coides and Ascaris suum a single species? Parasit. Vectors 5, 42. http://dx.doi.org/10.1186/
1756-3305-5-42.
Lima, V.S., I~ niguez, A.M., Otsuki, K., Ferreira, L.F., Ara ujo, A., Vicente, A.C.P.,
Jansen, A.M., 2008. Chagas disease by Trypanosoma cruzi lineage I in a hunter-gatherer
ancient population in Brazil. Em. Inf. Dis. 14, 1001e1002.
Littlewood, T.J., Donovan, S.K., 2003. Fossil parasites: a case of identity. Geol. Today 19,
136e142.
London, D., Hruschka, D., 2014. Helminths and human ancestral immune ecology:
what is the evidence for high helminth loads among foragers? Am. J. Hum. Biol. 26,
124e129.
Martill, D.M., Davis, P.G., 1998. Did dinosaurs come up to scratch? Nature 396, 528e529.
Martinson, E., Reinhard, K.J., Buikstra, J.E., Dittmar, K., 2003. Pathoecology of Chiribaya
parasitism. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 195e205.
384 Adauto Ara
ujo et al.

Marzochi, M.C., Coutinho, S.G., De Souza, W.J., Toledo, L.M., Grimaldi, G., Momen, H.,
Pacheco, R., Sabroza, P.C., Souza, M.A., Rangel, F.B., 1985. Canine visceral leishman-
iasis in Rio de Janeiro, Brazil. Clinical, parasitological, therapeutical and epidemiological
findings (1977-1983). Mem. Inst. Oswaldo Cruz 80, 349e357.
Matsui, A., Kanchara, M., Kanehara, M., 2003. Paleoparasitology in Japan e discovery of
toilet features. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 127e136.
Mitchell, P., 2013. The origins of human parasites: exploring the evidence for endoparasitism
throughout human evolution. Int. J. Paleopathol. 3, 191e198.
Mitchell, P.D., Anastasiou, E., Syon, D., 2011. Human intestinal parasites in crusader
Acre: evidence for migration with disease in the medieval period. Int. J.Paleopathol.
1, 132e137.
Montenegro, A., Ara ujo, A., Hetherington, R., Ferreira, L.F., Weaver, A., Eby, M., 2006.
Parasites, paleoclimate and the peopling of the Americas: using the hookworm to time
the Clovis migration. Cur. Anthropol. 47, 193e198.
Morris, J.N., 1988. Excavations at Casa Bisecada (Site 5MT8829), a Pueblo I field House/
Pueblo II habitation. In: Kuckelman, K.A., Morris, J.N. (Eds.), Archaeological Investi-
gations on South Canal, vol. I. Complete Archaeological Service Associates, Cortez,
Colorado, pp. 108e151. Four Corners Archaeological Project Report, no. 11.
Nagler, C., Haug, J.T., 2015. From fossil parasitoids to vectors: insects as parasites and hosts.
Adv. Parasitol. 90, 137e200.
Nerlich, A.G., Bianucci, R., Trisciuoglio, A., Sch€ onian, G., Ball, M., Giuffra, V.,
Bachmeier, B., Pusch, C.M., Ferroglio, E., Fornaciari, G., 2012. Visceral leishmaniasis
during Italian Renaissance, 1522e1562. Emerg. Infect. Dis. 18, 184e186.
Oh, C.S., Seo, M., Lim, N.J., Lee, S.J., Lee, E.J., Lee, S.D., Shin, D.H., 2010. Paleoparasi-
tological report on Ascaris aDNA from an ancient East Asian sample. Mem. Inst. Oswaldo
Cruz 105, 225e228.
Patrucco, R., Tello, R., Bonavia, D., 1983. Parasitological studies of coprolites of pre-
Hispanic Peruvian populations. Curr. Anthropol. 24, 393e394.
Pavlovsky, E.N., 1964. Natural Nidality of Transmissible Infections. Nauka, Moscow.
Pessoa, S.B., 1951. Parasitologia Médica. Ed. Guanabara, Waissman Koogan, Rio de Janeiro,
BR, p. 11.
Piombino-Mascali, D., Zink, A.R., Reinhard, K., Lein, M., Panzer, S., Aufderheide, A.C.,
Rachid, R., De Souza, W., Ara ujo, A., Chaves, S.A.M., Leroy-Toren, S., Teixeira-
Santos, I., Dutra, J.M.F., 2013. Dietary analysis of Piraino 1, Sicily, Italy: the role of
archaeopalynology in forensic science. J. Archaeol. Sci. 40, 1935e1945.
Pinto Dias, J.C., 2013. Human Chagas disease and migration in the context of globalization:
some particular aspects. J. Trop. Med. 2013, 789758. http://dx.doi.org/10.1155/2013/
789758.
Poinar, Jr., G., Poinar, R., 2004a. Paleoleishmania proterus n. gen., n. sp., (Trypanosomatidae:
Kinetoplastida) from Cretaceous Burmese Amber. Protist 155, 305e310.
Poinar, Jr., G., Poinar, R., 2004b. Evidence of vector-borne disease of Early Cretaceous
reptiles. Vector-Borne Zoonotic Dis. 4, 281e284.
Poinar, Jr., G., 2005. Triatoma dominicana sp. n. (Hemiptera: Reduviidae: Triatominae),
and Trypanosoma antiquus sp. n. (Stercoraria: Trypanosomatidae), the first fossil evi-
dence of a triatomine-trypanosomatid vector association. Vector-Borne Zoonotic
Dis. 5, 72e81.
Poinar, G., Boucot, A.J., 2006. Evidence of intestinal parasites of dinosaurs. Parasitology 133,
245e249.
Poinar, Jr., G., 2007. Early Cretaceous trypanosomatids associated with fossil sand fly larvae in
Burmese amber. Mem. Inst. Oswaldo Cruz 102, 635e637.
Poinar, H.N., Schwarz, C., Qi, J., Shapiro, B., MacPhee, R.D.E., Buigues, B., Tikhonov, A.,
Huson, D.H., Tomsho, L.P., Auch, A., Rampp, M., Miller, W., Schuster, S.C., 2006.
Palaeoparasitology e Human Parasites in Ancient Material 385

Metagenomics to Paleogenomics: large-scale sequencing of Mammoth DNA. Science


311, 392.
Poinar, G.O., 2014. Evolutionary history of terrestrial pathogens and endoparasites as
revealed in fossils and subfossils. Adv. Biol. Article ID 181353, 29 pages http://dx.doi.
org/10.1155/2014/181353.
Poulin, R., Combes, C., 1999. The concept of virulence: interpretations and implications.
Parasitol. Today 15, 474e475.
Raoult, D., Reed, D.L., Dittmar, K., Kirchman, J.J., Rolain, J.M., Guillen, S., Light, J.E.,
2008. Molecular identification of lice from pre-Columbian mummies. J. Infect. Dis.
197, 535e543.
Reed, D.L., Smith, V.S., Hammond, S.L., Rogers, A.R., Clayton, D.H., 2004. Genetic
analysis of lice supports direct contact between modern and archaic humans. PLoS
Biol. 2, e340.
Reinhard, K., 1985. Parasitism at Antelope House: a Puebloan village in Canyon de Chelly,
Arizona. In: Merbes, C.F., Miller, R.J. (Eds.), Health and Disease in the Prehistoric
Southwest. Tempe, Arizona State University, Anthropol. Papers, 34.
Reinhard, K.J., Ambler, J.R., McGuffie, M., 1985. Diet and parasitism at Dust Devil cave.
Am. Antiq. 50, 819e824.
Reinhard, R., Confalonieri, U., Ferreira, L.F., Herrmann, B., Ara ujo, A., 1986. Recovery of
parasite remains from coprolites and latrines: aspects of paleoparasitological technique.
Homo (Stuttgart) 37, 217e239.
Reinhard, K., 1988. Cultural ecology of prehistoric parasitism on the Colorado Plateau as
evidenced by coprology. Am. J. Phys. Anthropol. 77, 355e366.
Reinhard, K., 1992a. Parasitology as an interpretative tool in archaeology. Am. Antiq. 57,
231e245.
Reinhard, K., 1992b. Patterns of diet, parasitism, and anemia in prehistoric west North
America. In: Stuart- Macadam, P., Kent, S. (Eds.), Diet, Demography, and Disease:
Changing Perspectives in Anemia. Aldine de Gruyter, New York, pp. 219e258.
Reinhard, K., Bryant Jr., V.M., 1992. Coprolites analysis: a biological perspective on archae-
ology. In: Schiffer, M.,B. (Ed.), Advances in Archaeological Method and Theory, vol. 4.
Academic Press, New York, pp. 245e288.
Reinhard, K., Ara ujo, A., Ferreira, L.F., Coimbra, C.E., 2001. American hookworm
antiquity. Med. Anthropol. 20, 96e101 discussion 101e104.
Reinhard, K., Urban, O., 2003. Diagnosing ancient diphyllobothriasis from Chinchorro
mummies. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 191e193.
Reinhard, K., Fink, T.M., Skiles, J., 2003. A case of megacolon in Rio Grande Valley: pre-
liminary data. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 165e172.
Reinhard, K., Buikstra, J., 2003. Louse infestation of the Chiribaya Culture, Southern Peru:
variation in prevalence by age and sex. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 173e179.
Reinhard, K., Damon, T., Edwards, S.K., Meier, D.K., 2006. Pollen concentration analysis of
Ancestral Pueblo dietary variation. Palaeogeogr. Palaeoclimatol. Palaeoecol. 237, 92e109.
Reinhard, K.J., Ambler, J.R., Szuter, C.R., 2007. Hunter-gatherer use of small animal food
resources: coprolite evidence. Int. J. Osteoarchaeol. 17, 416e428.
Reinhard, K.J., 2008. Pathoecology of two Ancestral Pueblo villages. In: Reitz, E.J.,
Scarry, C.M., Scudder, S.J. (Eds.), Case Studies in Environmental Archaeology, second
ed. Plenum Press, New York, pp. 191e210.
Reinhard, K.J., Bryant Jr., V., 2008. Pathoecology and the future of coprolite studies in bio-
archaeology. In: Papers in Nat. Res., pp. 2e21.
Reinhard, K., Ara ujo, A., 2012. Synthesising parasitology in archaeology with
paleopathology. In: Buikstra, J., Roberts, C. (Eds.), The Global History of
Paleopathology: Pioneers and Prospects, vol. 1. Oxford University Press, Oxford,
pp. 751e764.
386 Adauto Ara
ujo et al.

Reinhard, K., Ferreira, L.F., Bouchet, F., Sianto, L., Dutra, J., I~ niguez, A.M., Leles, D., Le
Bailly, M., Fugassa, M., Pucu, E., Ara ujo, A., 2013. Food, parasites, and epidemiological
transitions: a broad perspective. Int. J. Paleopathol. 86, 1e8.
Reinhard, K., Pucu, E., 2014. Comparative parasitological perspectives on epidemiologic
transitions: the Americas and Europe. In: Zuckerman, M.K. (Ed.), Modern Environ-
ments and Human Health. Wiley Blackwell Eds., John Wiley and Sons, Inc, Hoboken,
New Jersey, pp. 321e336.
Rep, B.H., 1963. On the polyxenia of Ancylostomidae and the validity of the characters for
their differentiation. (II). Trop. Geogr. Med. 12, 271e316.
Rezende, J., 2003. Maculo e sua variada sinonímia. Rev. Patol. Trop. 32, 131e135.
Rothhammer, F., Allison, M.J., N un~ez, L., Standen, V., Arriaza, B., 1985. Chagas’ disease in
pre-Columbian South America. Am. J. Phys. Anthropol. 68, 495e498.
Ruffer, M.A., 1910. Note on the presence of Bilharzia haematobia in Egyptian mummies of
the twentieth dynasty (1250e1000). Br. Med. J. 16. Part 1.
Salgado, J.A., 1980. O centenario de Carlos Chagas e a menina Berenice. Mem. Inst.
Oswaldo Cruz 75, 193e195.
Samuels, R., 1965. Parasitological study of long dried fecal samples. In: Osborne, D.,
Katz, B.S. (Eds.), Contributions of the Wetherill Mesa Archaeological Project. Society
for American Archaeology, Memoirs, p. 19.
Santoro, C., Vinton, S.D., Reinhard, K., 2003. Inca expansion and parasitism in the Lutta
Valley: preliminary data. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 161e163.
Schmid-Hempel, P., 2011. Evolutionary Parasitology. The Integrated Study of Infectious,
Immunology, Ecology, and Genetics. Oxford University Press, UK, p. 5.
Searcey, N., Reinhard, K., Egarter-Vigl, E., Maixner, F., Piombino-Mascali, D., Zink, A.R.,
van der Sanden, W., Gardner, S.L., Bianucci, R., 2013. Parasitism of the Zweeloo
Woman: Dicrocoeliasis evidenced in a Roman period bog mummy. Int. J. Paleopathol.
3, 224e228.
Seo, M., Shin, D.H., Guk, S.M., Oh, C.S., Lee, E.J., Shin, M.H., Kim, M.J., Lee, S.D.,
Kim, Y.S., Yi, Y.S., Spigelman, M., Chai, J.Y., 2008. Gymnophalloides seoi eggs from
the stool of a 17th-century female mummy found in Hadong, Republic of Korea. J. Par-
asitol. 94, 467e472.
Seo, M., Oh, C.S., Chai, J.Y., Jeong, M.S., Hong, S.W., Seo, Y.M., Shin, D.H., 2014a. The
changing pattern of parasitic infection among Korean populations by paleoparasitological
study of Joseon Dynasty mummies. J. Parasitol. 100, 147e150.
Seo, M., Ara ujo, A., Reinhard, K.J., Chai, J.Y., Shin, D.H., 2014b. Paleoparasitological
studies on mummies of the Joseon Dynasty, Korea. Korean J. Parasitol. 52, 235e242.
Shapiro, B., Rambaut, A., Gilbert, M.T., 2006. No proof that typhoid caused the Plague of
Athens (a reply to Papagrigorakis et al.). Int. J. Infect. Dis. 10, 334e335.
Shin, D.H., Lim, D.S., Choi, K.J., Oh, C.S., Kim, M.J., Lee, I.S., Kim, S.B., Shin, J.E.,
Bok, G.D., Chai, J.Y., Seo, M., 2009. Scanning electron microscope study of ancient
parasite eggs recovered from Korean mummies of the Joseon Dynasty. J. Parasitol. 95,
137e145.
Shin, D.H., Oh, C.S., Chai, J.Y., Lee, H.J., Seo, M., 2011. Enterobius vermicularis eggs discov-
ered in coprolites from a medieval Korean mummy. Korean J. Parasitol. 49, 323e326.
Shin, D.H., Oh, C.S., Chai, J.Y., Ji, M.J., Lee, H.J., Seo, M., 2012. Sixteenth century
Gymnophalloides seoi infection on the coast of the Korean Peninsula. J. Parasitol. 98,
1283e1286.
Sianto, L., Reinhard, K., Chame, M., Chaves, S.M., Mendonça de Souza, S.M.,
Gonçalves, M.L.C., Fernandes, A., Ferreira, L.F., Ara ujo, A., 2005. The finding of Echi-
nostoma sp. (Trematoda: Digenea) and hookworm eggs in coprolites collected from a
Brazilian mummified body dated of 600e1,200 years before present. J. Parasitol. 91,
972e975.
Palaeoparasitology e Human Parasites in Ancient Material 387

Sianto, L., Chame, M., Silva, C.S.P., Gonçalves, M.L.C., Reinhard, K., Fugassa, M.H.,
Ara ujo, A., 2009. Animal helminths in human archaeological remains: a review. Rev.
Inst. Med. Trop. S~ao Paulo 51, 119e130.
Sianto, L., Santos, I.T., Chame, M., Chaves, S.M., Souza, S.M.F.M., Ferreira, L.F.,
Reinhard, K., Ara ujo, A., 2012. Eating lizards: a millenary habit evidenced by
Paleoparasitology. BMC Res. Notes 5, 586. http://dx.doi.org/10.1186/1756-0500-5-586.
Silva, P.A., Borba, V., Dutra, J., Leles, D., Da-Rosa, A.A.S., Ferreira, L.F., Ara
ujo, A., 2014.
A new ascarid species in cynodont coprolite dated of 240 million years. Acad. Bras. Ci 86,
265e270.
Siqueira, L.M., Coelho, P.M., Oliveira, A.A., Massara, C.L., Carneiro, N.F., Lima, A.C.,
Enk, M.J., 2011. Evaluation of two coproscopic techniques for the diagnosis of schisto-
somiasis in a low-transmission area in the state of Minas Gerais, Brazil. Mem. Inst.
Oswaldo Cruz 106, 844e850.
Souffez, M., 2001. Des insects et des Homes: répresentations du poux dans la culture andine.
These. Université de Paris X Nanterre, Paris.
Sousa, G.S., n.d. Notícia do Brasil. S~ao Paulo, Livraria Martins Editora (originally published
in 1587).
Stone, A.C., Wilbur, A.K., Buikstra, J.E., Roberts, C.A., 2009. Tuberculosis and leprosy in
perspective. Am. J. Phys. Anthropol. 40 (Suppl. 49), 66e94.
Szidat, L., 1944. Uber € die Erhaltungsf€ahigkeit von Helmintheneierm in vor-und
F€uhgeschichtlichen Moorleichen. Z. Parasitenkd 13, 265e274.
Trager, W., 1988. Living Together. The Biology of Animal Parasites. Plenum Press, New
York, N.Y., 467 pp., p. 1e3.
Tuon, F.F., Neto, V.A., Amato, V.S., 2008. Leishmania: origin, evolution and future since the
Precambrian. FEMS Immunol. Med. Microbiol. 54, 158e166.
Van Cleave, H.J., Ross, J.A., 1947. A method for reclaiming dried zoological specimens. Sci-
ence 105, 318.
Vlaminck, J., Levecke, B., Vercruysse, J., Geldhof, P., 2014. Advances in the diagnosis of
Ascaris suum infections in pigs and their possible applications in humans. Parasitology
28, 1e8.
Walker, P.L., Bathurst, R.R., Richman, R., Gjerdrum, T., Andrushko, V.A., 2009. The
causes of porotic hyperostosis and cribra orbitalia: a reappraisal of the iron-deficiency-
anemia hypothesis. Am. J. Phys. Anthropol. 139, 109e125.
Warnock, P.J., Reinhard, K.J., 1992. Methods for extracting pollen and parasite eggs from
latrine soils. J. Archaeol. Sci. 19, 261e264.
Wilbur, A.K., Buikstra, J.E., 2006. Patterns of tuberculosis in the Americas: how can modern
biomedicine inform the ancient past? Mem. Inst. Oswaldo Cruz 101 (Suppl. 2), 59e66.
Wilbur, A.K., Farnbach, A.W., Knudson, K.J., Buikstra, J.E., 2008. Diet, tuberculosis, and
the paleopathological record. Curr. Anthropol. 49, 963e977.
Wilke, J.P., Hall, H.G., 1975. Analysis of Ancient Feces: A Discussion and Annotated
Bibliography. Archaeological Research Facility, Department of Anthropology, Univer-
sity California, Berkeley.
Wood, J.R., Wilmshurst, J.M., Rawlence, N.J., Bonner, K.I., Worthy, T.H.,
Kinsella, J.M., Cooper, A., 2013. A megafauna’s microfauna: gastrointestinal parasites
of New Zealand’s extinct moa (Aves: Dinornithiformes). PLoS One 8, e57315.
http://dx.doi.org/10.1371/journal.pone.0057315.
CHAPTER TEN

Human Parasites in Medieval


Europe: Lifestyle, Sanitation
and Medical Treatment
Piers D. Mitchell
Department of Archaeology and Anthropology, University of Cambridge, Cambridge, United Kingdom
E-mail: pdm39@cam.ac.uk

Contents
1. Introduction 390
1.1 Sources of evidence 391
1.2 Survival of intestinal parasite eggs over archaeological time 392
1.3 Prevalence of parasites in the past 393
2. Parasites in Europe Prior to the Medieval Period 395
3. Medieval Period 397
3.1 Sanitation 398
3.2 The fishing industry 403
3.3 Farming animals 404
3.4 Detecting migrations 406
3.5 Reduction in some zoonotic parasites 408
3.6 Ectoparasites in medieval Europe 408
3.7 Medical history of medieval parasites 410
4. Conclusion 412
References 413

Abstract
Parasites have been infecting humans throughout our evolution. However, not all
people suffered with the same species or to the same intensity throughout this
time. Our changing way of life has altered the suitability of humans to infection by
each type of parasite. This analysis focuses upon the evidence for parasites from archae-
ological excavations at medieval sites across Europe. Comparison between the patterns
of infection in the medieval period allows us to see how changes in sanitation, herding
animals, growing and fertilizing crops, the fishing industry, food preparation and migra-
tion all affected human susceptibility to different parasites. We go on to explore how
ectoparasites may have spread infectious bacterial diseases, and also consider what
medieval medical practitioners thought of parasites and how they tried to treat them.
While modern research has shown the use of a toilet decreases the risk of contracting
certain intestinal parasites, the evidence for past societies presented here suggests that
the invention of latrines had no observable beneficial effects upon intestinal health.
Advances in Parasitology, Volume 90
© 2015 Elsevier Ltd.
j
ISSN 0065-308X
http://dx.doi.org/10.1016/bs.apar.2015.05.001 All rights reserved. 389
390 Piers D. Mitchell

This may be because toilets were not sufficiently ubiquitous until the last century, or
that the use of fresh human faeces for manuring crops still ensured those parasite
species were easily able to reinfect the population.

1. INTRODUCTION
The last 10,000 years has seen dramatic changes in the way humans
lived and the kinds of parasites infecting them. As a species we are thought
to have originated about 200,000 years ago in east Africa. Early humans
started to migrate around the planet from around 60,000 BP (before present)
and they had started to settle in Europe by 30e40,000 BP (Pasternak, 2007;
Holmes, 2009; Stone and Lurquin, 2007). Humans were not the first such
species to reach Europe of course, as evidence for Homo erectus and Homo
neanderthalis have been found well before Homo sapiens sapiens arrived
(Mellars, 2004; Mosquera et al., 2013). However, no faecal samples from
hominins have so far been analysed for ancient parasites, and it is only
humans who have been studied in this way.
Until about 12,000 years ago, all humans are thought to have been
hunter-gatherers, obtaining their food by collecting it from their environ-
ment. However, around that time in the Middle East a major change
took place. Somebody decided to sow some of the seeds they had gathered
and the concept of farming was born. Roughly the same time people started
to catch and herd sheep and goats rather than just eating wild caught animals,
and this again radically changed the lives of many people (Zeder, 2011).
These ideas also spread to other populations such as Europe. More reliable
food sources seem to have allowed populations to grow so that towns and
small cities developed, with higher population densities to earlier times
(Gates, 2011). As population densities increased, people began to travel in
order to trade (Simmons and DiBenedetto, 2014). From the fourth millen-
nium BC sanitation technologies were also developed in the Middle East,
such as latrines, drains and clean piped water (Mitchell, 2015).
We can see that over this 12,000-year period there have been marked
changes in those factors that may have affected our risk of contracting and
spreading different kinds of parasites. The increasing size of towns and
numbers of burials in cemeteries over this time clearly suggests that these
changes allowed the population to increase. One possible explanation is
that the populations became larger because they were healthier in general.
However, an alternative hypothesis is that fertility increased as food was
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 391

more reliable for farmers (Lawson et al., 2012), but people were actually less
healthy since they harboured more infectious diseases and parasites. If we
were to compare the archaeological evidence in Europe for parasites in pre-
history and in the medieval period, we should be able to see the effects these
changes had upon disease in the past, and hopefully enable us to understand
which of the two hypotheses is correct.
As well as looking for changes in parasite infections over time, we can
also look for differences between geographical regions and the cultural habits
of the people who lived there (Gonçalves et al., 2003; Araujo et al., 2015a).
Comparison of parasites in different parts of medieval Europe can be made
using our knowledge of the different lifestyles of the populations in those
regions. This can help us to understand which factors were most important
in allowing such infections in the past, and so allow us to understand the
health implications of being born into different past societies. Such an
approach could assist with attempts to explain why some past empires or
civilizations were more successful than others.

1.1 Sources of evidence


The potential sources for preserved human parasites vary depending upon
the kinds of parasite we wish to study. Intestinal parasites are the best under-
stood, as the eggs of many helminths are quite robust in archaeological
contexts and can survive extremely long periods of time in favourable
environments (Poinar and Boucot, 2006; Dentzien-Dias et al., 2013; Hugot
et al., 2014). The bulk of research into ancient parasites in Europe has ana-
lysed the sediment from cesspools and latrines. These provide a mixture of
faecal material from all the people who used that cesspool, and so give a pop-
ulation level indication of the species of parasites present. Another source of
preserved faeces come from the pelvic area of burials. As the intestines and
other abdominal organs decompose after death the eggs of parasites present
there will be released and become part of the soil. This may theoretically
result in both mature and immature forms of parasites eggs being detected.
A third archaeological source of intestinal parasites are preserved pieces of
stool, coprolites. Some coprolites are simply dried faeces, while mineralized
coprolites have been preserved by the seepage of minerals into the faecal
material from a waterlogged environment (Reinhard et al., 1986). Most
coprolites in Europe are of the mineralized kind. Ectoparasites such as fleas
and lice may also be preserved in mummified bodies, clothing or in water-
logged soils (Mumcuoglu, 2008). Mummies have the potential to preserve
both desiccated single-celled parasites and large multicellular helminths.
392 Piers D. Mitchell

They may also preserve the DNA of the parasites, proteins secreted by these
parasites during life and potentially the antibodies produced by the host
against the parasite (Deelder et al., 1990; Ara ujo et al., 1998; Bianucci
et al., 2008; Dittmar, 2009; Anastasiou and Mitchell, 2013a). All these
sources of evidence may be used to identify the species of parasite present
in the past.
Historical evidence in medieval texts can enable us to understand what
people thought of parasites. Sometimes their records allow us to suggest
possible diagnoses from the descriptions of worms passed from the anus,
mouth or nose (Cox, 2002; Trompoukis et al., 2007). However, since there
are many challenges associated with diagnosing diseases in the past using
written sources, we do have to be careful in our interpretation of ancient
written texts if we are to be realistic in how much we can tell about the
past (Mitchell, 2011a). Also, since we already know which species were pre-
sent in Europe at different times based upon archaeological evidence, what is
probably of much more interest to us is where texts explain what medieval
people thought of parasites. Manuscripts of most value to us today are those
that give the accepted views of the time as to how intestinal worms were
created, what it meant when parasites were seen, if the worms needed to
be treated by doctors, and how to treat them if a decision was made to treat.
This brings study of parasites in the medieval period to a new level compared
with prehistory, where by definition there were no records to explain what
people thought of their worms.

1.2 Survival of intestinal parasite eggs over


archaeological time
The eggs of some species of parasite are much more robust than others. In
the same environmental conditions fragile eggs will tend to degrade, deform,
tear or disintegrate relatively quickly. For example, roundworm eggs (Ascaris
lumbricoides) may lose the mammillated outer coat in archaeological samples
with less than perfect conditions. Some archaeological eggs can change size
in certain environments, potentially shrinking from desiccation in dry
conditions. This can make it challenging to identify them with microscopy
after thousands of years exposed to the elements. For example, the fragile
intestinal protozoa Entamoeba histolytica and Giardia duodenalis are rarely
seen in archaeological sediments when using microscopy, but are more easily
detected using enzyme-linked immunosorbent assay (ELISA) (Gonçalves
et al., 2004; Frías et al., 2013; Le Bailly and Bouchet, 2015). The eggs of
some species of helminths with fragile walls, such as pinworm (Enteriobius
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 393

vermicularis) and hookworm (Ankylostoma duodenale/Necator americanus), have


been found in many sites in the Americas where desiccated coprolites have
survived (Ara ujo et al., 2008, 2011), but have been extremely rare in Europe
and Asia where most faecal samples come from waterlogged cesspools and
mineralized coprolites. The better preservation of pinworm and hookworm
eggs in desiccated coprolites than in waterlogged samples gives the impres-
sion that pinworm and hookworm were much more common in the Amer-
icas than in the rest of the world. However, it is quite likely that what we see
really represents different levels of resistance to decay in different environ-
ments rather than different prevalence in past populations (Reinhard and
Pucu de Ara ujo, 2014).
The environment can reduce our chance of detecting parasites that were
genuinely present during the life of an individual. Perhaps the source of sam-
ples most prone to this is that of pelvic soil, where egg counts are routinely
lower than coprolites or latrine soil. While a preserved coprolite effectively
locks in and protects the eggs of those parasites present when it was formed,
pelvic soil is an amorphous mass of sediment through which worms, insects
and rodents may burrow. As rains pass through the soil they may wash away
smaller particles such as parasite eggs, so that after centuries or millennia the
original eggs may no longer be detectable.
We must also remember that over archaeological periods of time, para-
sites themselves may have evolved so that their eggs may have changed in
shape, size or their special characteristics. Some species of parasite may also
have changed their endemic regions as climate changed or zoonotic hosts
moved on, so we may find them in places we would not expect today.

1.3 Prevalence of parasites in the past


It is possible to estimate how common parasites were in the past using egg
counts, but some techniques are more reliable than others (Jones, 1982;
Warnock and Reinhard, 1992; Fugassa et al., 2006; Ara ujo et al., 2015b).
Cesspool and latrine soil is a mixture of all the faeces of those who used
that latrine in the past, so is of only limited use in determining parasite prev-
alence. However, a low concentration of eggs of one species compared with
a high concentration of eggs of other species can suggest which species may
have been more common in that population. However, we cannot know if
every individual had a few worms or if one individual had large numbers of
worms. Egg counts only give us an indication of the number of female
worms present, as male worms that may be just as troublesome to the
host do not produce any eggs. We also need to bear in mind that female
394 Piers D. Mitchell

worms of different parasite species produce different numbers of eggs,


different individual worms of the same species produce different number
of eggs to each other, and the same worm can produce different numbers
of eggs on different days (Sinniah, 1982; Walker et al., 2009).
The archaeological nature of latrines means that we cannot interpret egg
counts in the same way as we might for fresh stool in a hospital laboratory.
Egg concentrations will rise with decomposition of the nonparasite elements
of the faeces. Hence the concentration of eggs in faeces deposited in the
warmer summer months may appear higher than in the winter when little
decomposition occurs, even though counts were identical in all original
samples. Secondly, egg concentrations will fall when extra material such as
rubbish is added to the cesspool, as this will dilute out the faecal material.
For these reasons we have to be very cautious when interpreting the con-
centrations of parasite eggs in archaeological latrine sediments. However,
for prehistoric time periods before the invention of cesspools and latrines,
we have to rely on other methods to identify helminth eggs.
Coprolites are pieces of preserved faeces that can be found in any archae-
ological time period (Bryant and Dean, 2006). Dried coprolites may form in
a cave or the intestines of a natural mummy where the environment protects
the faecal material from the rain and from the action of insects. Mineralized
coprolites are those that are preserved because minerals such as calcium
carbonate soak into the coprolite and harden it. These are sometimes found
in cesspools or rubbish tips (middens). Mineralized coprolites are much
heavier than desiccated coprolites, so we cannot directly compare number
of parasite eggs per gram of coprolite in desiccated and mineralized samples.
One recent novel application of the analysis of egg counts in coprolites came
from a sequential sampling of faecal material from the colon of a water-
logged medieval burial in Belgium (Racz et al., 2015). This showed how
egg counts for whipworm progressively increased more distally along the
intestines. If coprolites are not recovered from the pelvic area of a burial,
it is often difficult to know for sure whether they are of human origin or
from an animal. Approaches to determine the origin of coprolites include
the types of parasite egg present and the fragments of food present (Reinhard
and Bryant, 1992). However, the eggs of some helminths of animals appear
similar to those of humans, for example, roundworm and whipworm in pigs
(Ascaris suis and Trichuris suis) produce eggs microscopically indistinguishable
to roundworm and whipworm in humans (A. lumbricoides and T. trichiura).
The most reliable technique to identify the species of origin of a coprolite
is the detection of bile acids and sterols within the coprolite with mass
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 395

spectrophotometry, as different animals produce different kinds of bile acids


and sterols (Shillito et al., 2011; Baeten et al., 2012). However, this relatively
new technique is not currently being widely applied to coprolite analysis due
to its cost. Coprolites tend to preserve parasite eggs reasonably well,
although it remains unknown what proportion of the eggs originally present
survive archaeological time. However, we cannot currently tell if each
coprolite recovered was passed by a different individual, or if the same per-
son passed each coprolite. This means we cannot determine true prevalence
of helminths in people using this approach, just prevalence in their
coprolites.
A more reliable technique for determining past parasite prevalence is to
analyse a series of human burials. If each skeleton undergoes pelvic soil sam-
pling from the front of the sacrum, we can see how many of those samples
contain parasite eggs. Control samples from the head and feet are also taken
to ensure any positive results in the pelvic soil represent genuine infection
during life and not generalized contamination of the soil from later use of
the site for disposal of faeces (Reinhard et al., 1992). We should not expect
to find the eggs of intestinal parasites in the feet or head during life.
However, pelvic soil tends to give a minimum prevalence rather than true
prevalence, as rainwater passing through the soil over the centuries can
wash the parasite eggs away. This can mean that someone who was infected
with intestinal parasites during their life may have a negative analysis for
parasite eggs. For the same reasons, the concentration of parasite eggs in
pelvic soil is generally much lower than the case for coprolites or cesspool
soil. That means we cannot compare egg counts from pelvic soil reliably
between archaeological sites with different environmental conditions.
However, we can make some comment as to differences in egg counts of
each parasite species in the same individual, and in different burials from
the same site with similar environmental conditions.

2. PARASITES IN EUROPE PRIOR TO THE MEDIEVAL


PERIOD
If we are going to understand parasitism in medieval Europe, we need
to know what kinds of parasites were present in people living on the conti-
nent in earlier times. This will enable us to identify changes over time, and
make suggestions as to why those changes took place.
Very little is known about parasitism in prehistoric hunter-gatherers in
Europe during the Palaeolithic Period. At Arcy-sur-Cure in Yonne, France
396 Piers D. Mitchell

sediment dating from 30,000 to 24,000 BP was found to contain the egg of a
roundworm (Ascaris sp.) (Bouchet et al., 1996). It was thought to be of
human origin due to the archaeological context, so argued to probably be
an A. lumbricoides. No other parasite eggs were found at this site. It is difficult
to analyse faecal material for parasites at European hunter-gatherer sites as
the groups were by definition mobile. This means there were no permanent
buildings to us to excavate, no rubbish tips where faeces may have been
deposited, latrines had not yet been invented, and often no cemeteries
where the dead from a tribe may be placed close to one another for archae-
ologists to find their skeletal remains. In most European societies, these only
appeared once people decided to settle in one place, and the trigger for their
staying put seems to have been practicing agriculture.
Farming started to be practiced in Europe during the Neolithic period
(about 4000e2800 BC), when both agriculture and herding was practiced.
The exact dates vary across the region as these techniques gradually spread
over time. Across Neolithic Europe a large number of parasite species
have been identified, in what is now France, Germany, Greece, the
Netherlands and Switzerland. These include beef and/or pork tapeworm
(Taenia sp.), bile duct fluke (Opisthorchis sp.), Capillaria sp., dysentery proto-
zoa (E. histolytica), fish tapeworm (Diphyllobothrium sp.), giant kidney worm
(Dioctophymidae), hookworm (Ancylostomids), lancet liver fluke (Dicrocoe-
lium sp.), fasciola liver fluke (Fasciola hepatica), roundworm (A. lumbricoides),
whipworm (T. trichiura) (Anastasiou, 2015; Bouchet et al., 2003).
It is clear that the number of sites positive for parasites, and the range of
species present, is dramatically different in the Neolithic to the earlier Palae-
olithic period. Some of these parasites are zoonoses that people probably
contracted from eating raw or undercooked wild animals (e.g. Opisthorchis
sp., Capillaria sp., Dioctophymidae, Dicrocoelium sp., Diphyllobothrium sp.).
We might expect similar findings in earlier pure hunter-gatherer groups if
such studies had been possible. Other species of parasite are those that we
associate with sedentary lifestyle in farmers. Some are typically spread by
faecal contamination and poor sanitation (Ancylostomids, A. lumbricoides,
E. histolytica, T. trichiura), and others by the consumption of undercooked
farm animals (Taenia solium and Taenia saginata).
The Bronze Age (about 2800e500 BC) and Iron Age (about 700e100
BC) heralded a move from stone tools to those made of metal, and the overlap
in dates reflects how this was a gradual process spreading across the continent.
Parasites in humans from this time have been found in what is now Austria,
Britain, the Czech Republic, Denmark, France, Germany and Poland. The
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 397

species identified are dysentery protozoa (E. histolytica), lancet liver fluke
(Dicrocoelium sp.), roundworm (A. lumbricoides), whipworm (T. trichiura)
(Anastasiou, 2015; Bouchet et al., 2003). There does appear to be a trend to-
wards parasites spread by faecal contamination and poor sanitation, and little
evidence for those parasites spread by the eating of undercooked wild animals.
Whether this means those animals are no longer being caught and consumed,
or whether they are just being cooked better, is not clear from the parasite
evidence.
With the Roman period (about 100 BCe400 AD) we move from prehis-
tory to a time when written records complement archaeological assessment of
past societies. Parasites have been identified from excavations undertaken in
what is now Austria, Britain, France, Germany, Italy and Poland. The species
found are beef and/or pork tapeworm (Taenia sp.), dog tapeworm calcified
cysts (Echinococcus granulosus), fish tapeworm (Diphyllobothrium sp.), lancet liver
fluke (Dicrocoelium sp.), roundworm (A. lumbricoides), whipworm (T. trichiura)
(Anastasiou, 2015; Bouchet et al., 2003). Again, we see a similar pattern of
parasite species to the Bronze and Iron Ages. The dominant parasites seem
to be those spread by poor sanitation, faecal contamination and working
with dogs. It is interesting that in all time periods lancet liver fluke is found.
Since this is a parasite spread by the consumption of ants, this would indicate
either that people liked eating ants in the past, or that they were inadvertently
consumed along with foods. With the fall of the Roman Empire we move to
the early medieval period.

3. MEDIEVAL PERIOD
The medieval period in Europe lasts from about 500 to 1500 AD,
varying in different parts of the continent. There have been a large number
of medieval European archaeological sites where parasites have been studied
(for example, Boersma and Jansen, 1975; Jones, 1982; Herrmann, 1988;
Anastasiou and Mitchell, 2013b; Yeh et al., 2014). However, this does
not mean that the time period is a boring one to research because there is
nothing new to discover. It is actually the very body of evidence and the
contrasting findings in different contexts that means we can look in a
more nuanced way than is the case for continents or time periods for which
there are few published studies. These excavations have provided many
distinct pieces of fascinating evidence, and we are reaching a stage where
bigger questions can be examined using this data. Here we will interpret
398 Piers D. Mitchell

the differences in parasite species and prevalence across medieval Europe by


considering variations in sanitation, population density, diet and cookery,
farming practices and the effect of migration.
There are many perceptions about parasites in medieval Europe that are
often found in the literature, such as that everyone had roundworm and
whipworm, and everyone had fleas and lice. It is also often thought that
the farming of animals led to those parasites spread in their meat to be wide-
spread. However, detailed analysis of the evidence shows that the picture is
not quite that simple.
In the medieval period, parasites have been recovered from Austria,
Belgium, Britain, Denmark, France, Germany, Greenland, Italy, Latvia,
the Netherlands, Norway and Switzerland. The species identified include
beef and/or pork tapeworm (Taenia sp.), bilharzia (Schistosoma haematobium
and Schistosoma mansoni), dog tapeworm calcified cysts (E. granulosus), dysen-
tery protozoa (E. histolytica and G. duodenalis), fasciola liver fluke (F.
hepatica), fish tapeworm (Diphyllobothrium latum), lancet liver fluke (Dicrocoe-
lium dendriticum), pinworm (E. vermicularis), roundworm (A. lumbricoides),
trichinosis (Trichinella spiralis), whipworm (T. trichiura). A summary of key
examples has been given in Table 1. This does not detail every publication
about of every parasite, but does include an example from each country
where parasites have been recovered from medieval sites. This gives a feel
for where in Europe certain species were found, and how widespread partic-
ular species were. It does need to be borne in mind that there has been very
little ancient parasite research in Portugal, Spain, Malta, Greece and parts of
Eastern Europe which means that the absence of data for these countries
does not mean that the populations living in those countries in the past
were free of parasites.

3.1 Sanitation
One of the key observations from this table is that roundworm and whip-
worm were found in more countries than any other parasites. They are
both faecal-oral parasites, spread by poor sanitation and faecal contamination
of food. This finding is well known, and has been commented upon many
times before (Herrmann, 1985; Reinhard et al., 2013; Reinhard and Pucu
de Ara ujo, 2014). Roundworm and whipworm (Figure 1) were not
restricted to the poor who may not have had clean water, clean clothes or
latrines in their houses. Even kings such as Richard III of England (ruled
1483e1485) were infected with roundworm (Figure 2) (Mitchell et al.,
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 399

Table 1 Endoparasite species found at archaeological sites in Medieval Europe


Species Date (century) Country References

Beef/pork tapeworm 8th Denmark Nansen and Jorgensen


(Taenia sp.) (1977)
11the13th Belgium Da Rocha et al. (2006)
11the15th Britain De Rouffignac (1985)
11the16th Germany Herrmann (1988)
Bilharzia (schistosomiasis)
• Schistosoma 15the16th France Bouchet and Paicheler
haematobium (1995)
• Schistosoma mansoni 15the16th France Bouchet et al. (2002)
Dog tapeworm calcified 11the12th Britain Price (1975)
cysts (Echinococcus 13the15th Iceland Kristjansdottir and
granulosus) Collins (2011)
15th Denmark Weiss and Møller-
Cristensen (1971)
Dysentery protozoa
• Entamoeba histolytica 7the9th Switzerland Le Bailly and Bouchet
(2006)
11the13th France Le Bailly and Bouchet
(2006)
14th Belgium Goncalves et al. (2004)
• Giardia duodenalis 7the9th Switzerland Le Bailly and Bouchet
(2006)
11the13th France Le Bailly et al. (2008)
Fasciola liver fluke 8th Denmark Nansen and Jorgensen
(Fasciola hepatica) (1977)
11the12th Belgium Da Rocha et al. (2006)
11the16th Germany Herrmann (1988)
Fish tapeworm 10the11th Italy Florenzano et al. (2012)
(Diphyllobothrium 11the15th Belgium Da Rocha et al. (2006)
latum) 11the15th Britain De Rouffignac (1985)
11the15th Norway Jones (1988)
11the16th Germany Herrmann (1988)
14th Latvia Yeh et al. (2014)
15the16th France Bouchet (1995)
Lancet liver fluke 11the15th Britain Pike (1967)
(Dicrocoelium 11the16th France Le Bailly and Bouchet
dendriticum) (2010)
Pinworm (Enterobius 11the16th Germany Herrmann (1988)
vermicularis) 15th Greenland Hansen (1986)
(Continued)
400 Piers D. Mitchell

Table 1 Endoparasite species found at archaeological sites in Medieval Europedcont'd


Species Date (century) Country References
Roundworm 8th Denmark Nansen and Jorgensen
(Ascaris lumbricoides) (1977)
9the5th Britain Jones (1983)
10the11th Italy Florenzano et al. (2012)
11the15th Norway Jones (1988)
11the16th Germany Herrmann (1988)
11the16th France Bouchet (1995)
14th Latvia Yeh et al. (2014)
14the15th Netherlands Boersema and Jansen
(1975)
Trichinosis 16th Greenland Lynnerup (2007)
(Trichinella spiralis)
Whipworm 8th Denmark Nansen and Jorgensen
(Trichuris trichiura) (1977)
9th Belgium Da Rocha et al. (2006)
9the15th Britain Jones (1983)
10the11th Italy Florenzano et al. (2012)
11the15th Norway Jones (1988)
11the16th Germany Herrmann (1988)
11the16th France Bouchet (1995)
14th Latvia Yeh et al. (2014)
14the15th Netherlands Boersema and Jansen
(1975)

Figure 1 Whipworm egg from Viking period latrine at York, UK. Dimensions
50  23 mm. Black bar indicates 20 mm.
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 401

Figure 2 Decorticated roundworm egg from the pelvic soil of King Richard III of
England (died 1485 AD). Dimensions 61  44 mm. Black bar indicates 20 mm.

2013). So we must ask ourselves why there seems to have been so much
roundworm and whipworm around in medieval Europe.
Modern studies have shown that when nomadic hunter-gatherers
around the world settle and change their lifestyle, the prevalence of infection
of roundworm, whipworm, hookworm and pathogenic amoebae increase
dramatically (Dounias and Froment, 2006). While these studies are generally
on forest-based groups living in the tropics, this is quite a significant finding
and starts to help us interpret why sedentary medieval people experienced so
much roundworm and whipworm too. Clinical research has also shown that
use of latrines, clean drinking water and washing hands with soap all reduce
the likelihood of contracting soil-transmitted helminths (geohelminths) such
as roundworm and whipworm (Zeigelbauer et al., 2012). Applying human
faeces to agricultural fields is known to increase crop yields (Heinonen-
Tanski and van Wijk-Sebesma, 2005). However, it has also been found
that either avoiding using human faeces as fertilizer, or at least composting
it for a minimum of 6 months prior to applying it to fields, greatly reduces
the risk of contracting these geohelminths from the food grown there (Uga
et al., 2009; Phuc et al., 2006; Jensen et al., 2008).
Archaeological evidence shows that the latrine was in use in people’s
houses in the Middle East by the late 4th millennium BC (McMahon,
2015), and in Greece by the 2nd millennium BC (Antoniou and Angelakis,
2015). The Romans adopted the sanitation technology of the Greeks,
402 Piers D. Mitchell

improved it with large multiuser latrines, water irrigation networks, bath-


houses and sewers (Antoniou and Angelakis, 2015; Taylor, 2015). With
the expansion of the Roman Empire across Europe such ideas spread
throughout much of the continent. By the medieval period we know that
there were many latrines and cesspools, as archaeologists have been exca-
vating them and their contents have been analysed for parasites and discussed
here in this article. Textual evidence for latrines in medieval cities such as
London has shown the presence of many public and private latrines (Taylor,
2015). Despite these latrines, roundworm and whipworm seems to have
been common in medieval Europe. This would suggest that the invention
of the latrine does not seem to have had any beneficial impact upon the
prevalence of those parasites spread by faecal contamination of food. This
is quite thought provoking idea.
If despite 4000 years of latrines the faecal contamination of food became
progressively more common, not less, we must ask ourselves two questions.
Firstly, why were latrines invented in the first place, and secondly why were
they so ineffective at preventing the faecal contamination of food? The first
question will remain debatable since the latrines found in Mesopotamia were
made before the development of writing, so we can never know the
reasoning behind the first toilet designer’s ideas. While it is possible they
had noticed a health risk from being near decomposing human faeces, it
may well be the first latrines were actually built to reduce smells in the
home, cut down on annoying flies or to avoid the need to walk to the
town rubbish tip with a pot of excrement every time someone in the family
opened their bowels. The second question is that of why the presence of
both private and public latrines were so ineffective in preventing the spread
of geohelminths in medieval societies. Modern research suggests latrines
should be effective for this task (Zeigelbauer et al., 2012). One possibility
is that there were just not enough toilets in existence in the medieval period
so that the prevalence of roundworm and whipworm was maintained by
those who could not access them. However, another key point is that in
the past it was common for people to fertilize their crops with human faeces
in order to increase crop yields (Jones, 2012; Magnusson, 2013). Medieval
written records show how human waste was collected from towns and taken
out to the fields. Cesspits and sewers were intermittently emptied and the
contents sold on as fertilizer (Sabine, 1934, 1937; Scobie, 1986; Taylor,
2015). There is no record of a concept of composting the faeces for a certain
period of time, so it would most likely have been applied to the fields fresh.
I suspect it was this use of faeces as crop fertilizer that was a major factor
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 403

explaining why roundworm and whipworm were so common in medieval


Europe, and why latrines failed to prevent their spread.

3.2 The fishing industry


The next most widely found parasite in medieval Europe was that of fish
tapeworm (D. latum). This has been identified from excavations in seven
countries so far, in Norway, Latvia, Belgium, Germany, France, Britain
and northern Italy (Figure 3 and Table 1). We saw earlier that fish tapeworm
has been found in Europe in the Neolithic and Roman periods, so it is quite
likely that D. latum was a zoonotic parasite originating in the wild animals of
northern Europe (Le Bailly and Bouchet, 2013). However, it does seem to
have been more widespread in its distribution during the medieval period,
especially in northern Europe. This parasite evidence for higher fish con-
sumption is supported by isotope research in human skeletal remains. For
example, people living in coastal Norway in the Viking Age (800e1050
AD) showed much higher isotopic evidence for fish consumption during
their lives that did those who lived in the same region during the earlier
Merovingian Age (550e800 AD) (Naumann et al., 2014). It seems that dur-
ing the medieval period, people in northern Europe were eating more fish
than had been the case in earlier time periods.

Figure 3 Fish Tapeworm egg from a fourteenth century AD latrine at Riga, Latvia.
Width 45 mm. Black bar indicates 20 mm.
404 Piers D. Mitchell

Fish tapeworm is contracted by humans who consume uncooked fish, as


cooking kills off the intermediate form of the parasite (Garcia, 2009, p. 354;
Gunn and Pitt, 2012, p. 104). Only fish that spend at least part of their life
cycle in freshwater can become infected with D. latum. In northern Europe
it was common to eat raw, smoked, pickled or salted fish in the medieval
period (Buxton, 1998; Hoffman, 2003; Hagen, 2006). Fishing became an
important industry both in rivers and also in the North Sea and Baltic,
with huge amounts of fish caught and preserved each year (Barrett et al.,
2004, 2011). Fish was also an important part of the diet in Christian coun-
tries at that time, even in people living well away from the sea. This is
because the church prohibited the consumption of meat on certain days
of the week, on a range of religious days and throughout lent, but fish
was allowed then (Buxton, 1998; Van Dam, 2009). It seems likely that
fish tapeworm infection became more common in the medieval period as
a consequence of the culinary practices widespread in northern Europe
coupled with the dietary regulations of the church.

3.3 Farming animals


It has been argued that herding animals in Europe also led to an increase in
impaired health as they spread animal diseases (Sianto et al., 2009). This
appears quite logical, as some diseases can infect both humans and animals,
and other diseases of humans require time in an animal to complete their life
cycle. However, modern clinical research shows that this idea is not quite as
simple as it may first appear. For example, study of children living in families
that keep livestock in Vietnam found exactly the same incidence of diar-
rhoea than was the case for those families who did not work with animals
(Thiem et al., 2012). This suggests that despite all the animal dung, living
with farm animals does not automatically lead to spread of gastrointestinal
disease and poorer health. It is helpful to look at those specific examples
where there does seem to be a link between parasitic diseases and farm
animals in medieval times, in order to see which diseases may have been
more common as a result of herding large animals.
Medieval people used domesticated dogs to guard herd animals, and
sometimes to move them between different pastures (Kelly, 1997; Lucas,
1989; O’Connor, 2008). Dogs may harbour the dog tapeworm, E. granulosus
in their intestines. Normally the eggs are passed in dog faeces and inadver-
tently consumed by herbivores such as sheep as they graze. The intermediate
form of a hydatid cyst then develops in their organs ready to be eaten by
another dog that preys upon that sheep. However, if humans consume foods
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 405

contaminated by dog faeces then these hydatid cysts develop in the chest,
abdomen or brain (Garcia, 2009, p. 362; Gunn and Pitt, 2012, p. 108).
The walls of these cysts often calcify and are preserved after death, to be
recovered at archaeological excavation of burials. Examples of these calcified
hydatid cysts have been recovered from medieval contexts in Britain, Iceland
and Denmark (Table 1). It has been proposed that hydatid disease became
endemic in Iceland following the shipment of dogs from Germany around
1200 AD (Kristjansd ottir and Collins, 2011). This shows these individuals
must have been exposed to dog faeces on their food. Those medieval people
using dogs to herd sheep or hunt wild animals would clearly have been in this
high risk group.
Some parasites specific to humans spend part of their life cycle in an an-
imal before they are spread to people who eat it. If those animals are species
we chose to farm, then it is logical to assume that those parasites will become
more common as a result. Good examples of this scenario are the beef and
pork tapeworms (T. saginata and T. solium). The eggs of these parasites
appear identical under the microscope, so are not easy to differentiate in
archaeological contexts. Genetic studies of the Taenia genus suggest beef
tapeworm evolved as a specialist parasite of humans in Africa at least 1 million
years ago, while the origins of pork tapeworm is less clear and has been
linked with both lions and bears (Hoberg et al., 2001; Mitchell, 2013; Terefe
et al., 2014). It is likely that the two species survived in wild cattle and wild
pigs that were hunted or scavenged by human hunter-gatherers. However,
when people started to farm large animals, they found that cattle and pigs
were very suitable to domestication. Taenia tapeworm eggs have been found
in medieval Denmark, Belgium, Britain and Germany (Table 1), although it
has been found in other European countries in earlier time periods (Sianto
et al., 2009). Taenia sp. seems to have been less common in medieval times
than was fish tapeworm. This may reflect the different cooking methods
used for these meats, as in Europe pork and beef were typically cooked
and not generally eaten raw, dried, smoked or pickled.
Fasciola liver fluke (F. hepatica) has been found in Belgium, Denmark and
Germany during the medieval period (Table 1), but again in other countries
in earlier time periods (Sianto et al., 2009). It was clearly not as widespread as
roundworm, whipworm or fish tapeworm in medieval times. It is a zoonotic
parasite that typically infects sheep, cattle and other ruminants that may eat
semiaquatic plants contaminated by the metacercariae. If humans eat those
plants and do not cook them first, they can become infected by the fluke
(Garcia, 2009, p. 368; Gunn and Pitt, 2012, p. 91). One might expect
406 Piers D. Mitchell

that the farming of such animals would therefore increase the likelihood of
contracting the disease. Farming sheep and cattle was a major activity in
northern Europe in the medieval period, and the numbers of animals
were significantly higher than in earlier time periods (Astill and Langdon,
1997; Hoskins, 1955; Faith, 2010; Lucas, 1989). For example, there are
thought to have been over 10 million sheep in England by the thirteenth
century (Hurst, 2005). For these reasons, we would expect the expanding
farming sector to have increasingly predisposed those living in farming areas
to contracting fasciola liver flukes, so long as they continued to eat fresh salad
leaves.

3.4 Detecting migrations


While some parasites are fairly ubiquitous in their distribution around the
world, many species are only endemic in particular areas due to the require-
ments of their life cycle. The need for a particular species of intermediate
host, or particular temperatures for a period of maturation means that
many parasites can only be contracted in particular geographic regions.
When we find the eggs of that parasite in a region where the parasite cannot
complete its life cycle, it can be used as an indicator of human migration.
Bilharzia provides a good medieval example of this concept. The eggs of
both S. haematobium and S. mansoni have been recovered from a latrine at
Montbéliard in France, in use from 1450 to 1550 AD (Table 1). Schistoso-
miasis could not have been endemic in medieval Europe, as the winters were
too cold for survival of the freshwater snails that are required as intermediate
host in the parasite’s life cycle (Garcia, 2009, p. 372; Gunn and Pitt, 2012,
p. 101). This means at least one person who used this latrine must have trav-
elled to an area endemic for both these parasites, such as Africa or the Middle
East. Potential explanations include a merchant or explorer from France
who travelled to these regions and contracted bilharzia before returning
home, or alternatively a person native to Africa or the Middle East who trav-
elled to France.
Another good example of parasites indicating medieval migration is that of
fish tapeworm and the crusades. We have discussed earlier how D. latum was
common in northern Europe during the medieval period. Diphyllobothrium sp.
eggs have been found in two thirteenth-century latrines in the coastal city of
Acre. One was a large latrine block of the Order of St. John with 35 seats, and
the other a cesspool in a private house in the residential quarter of the city
(Mitchell and Stern, 2001; Mitchell et al., 2008a, 2011). In the twelfth and
thirteenth centuries, Acre was under the control of the crusaders from
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 407

Europe, and lay within the Frankish Kingdom of Jerusalem. The Order of St.
John was a military and hospitaller order, founded in 1112 AD to care for sick
pilgrims in the Holy Land (Nicholson, 2007; Riley-Smith, 2012). Not all
crusader latrines have been positive for fish tapeworm eggs, suggesting that
infection with Diphyllobothrium was by no means ubiquitous among crusaders
(Mitchell and Tepper, 2007; Anastasiou and Mitchell, 2013b). In the medie-
val period fish was not eaten raw, pickled or smoked in the Middle East as the
hot temperatures meant it decomposed more quickly than in northern
Europe. Medieval textual sources from the Middle East show that it was al-
ways eaten cooked (Lewicke, 2011). Cooking kills the intermediate forms of
the parasite and so prevents its transmission. Finding the eggs of fish tapeworm
in two separate medieval latrines used by crusaders from Europe suggests they
probably contracted the parasite in northern Europe where it was common,
and travelled to the east on crusade taking the tapeworms with them in their
intestines. The parasite does not seem to have ever become endemic in the
freshwater fish of the region (such as at Lake Tiberias), suggesting the local
environment probably did not meet all the requirements for its life cycle.
Such parasitological evidence complements written descriptions in textual
sources of the period, recording how epidemics and other infectious diseases
seem to have been spread by crusaders and pilgrims as they travelled (Mitchell,
2011b).
Potential evidence for long distance travel can be found following the
analysis of a Mamluk period cesspool from the Christian Quarter of Jerusalem
(Yeh et al., 2015). Radiocarbon dating and pottery fragments show the
cesspool was in use during the late 1400s and early 1500s. The pottery in
the latrine came from both the Jerusalem region and also from northern Italy.
Six species of parasites were found in coprolites and cesspool sediment, four
helminths (roundworm, whipworm, Taenia tapeworm and fish tapeworm)
and two protozoa (E. histolytica and G. duodenalis). Entamoeba histolytica seems
to have originated in Europe as all the early examples come from there
(Le Bailly and Bouchet, 2015). We have already seen how fish tapeworm
was common in northern Europe during the medieval period, and has
recently been identified in northern Italy (Florenzano et al., 2012). There
are a number of potential explanations as to how this picture may have
come about. One possibility is that merchants from Jerusalem travelled to
northern Italy, bought pottery and contracted parasites before returning
home. Another option to consider is that E. histolytica may have become
endemic in the region after its spread there by crusaders in the thirteenth cen-
tury, but the same would be unlikely for fish tapeworm as fish was typically
408 Piers D. Mitchell

eaten cooked in the Middle East and this would prevent parasite transmission.
A further possibility is that this was a hostel providing accommodation for
European traders or pilgrims travelling to the Christian quarter of the holy
city (Yeh et al., 2015).

3.5 Reduction in some zoonotic parasites


It can be seen that some of the parasites found in humans during the
Neolithic period (4000e2800 BC) are not found in the medieval period.
These all seem to be zoonoses, parasites that are primarily found in animals
but can be spread to humans in certain circumstances (Sianto et al., 2009).
Examples found in the Neolithic but not in medieval times include the
bile duct fluke (Opisthorchis sp.), Capillaria sp. and the giant kidney worm
(Dioctophymidae) (Roever-Bonnet et al., 1979; Bouchet, 1997; Le Bailly
et al., 2003). Unlike the zoonoses of large herbivores discussed above (Taenia
sp., F. hepatica), these are found in animals that were not farmed in medieval
times. For example, the giant kidney worm is contacted by eating freshwater
frogs or fish, capillariasis by eating rodents or other mammals, and opis-
thorchis bile duct fluke by eating freshwater fish. It seems likely that these
were more common in the Neolithic as people still supplemented the
farmed components of their diet with hunting and gathering from the
wild, coupled with eating some of this raw or undercooked. The fact that
the zoonotic parasites of animals that were not farmed seem to have been
much less common in the medieval period may indicate that gathering
wild foods was less common in medieval Europe, or that it was cooked thor-
oughly before being eaten.

3.6 Ectoparasites in medieval Europe


A number of ectoparasites of humans have been found in archaeological
deposits in Europe. These have been recovered from soil sediment, textiles,
mummies and combs. Head lice (Pediculus humanus capitis), body lice
(Pediculus humanus corporis) and crab/pubic lice (Phthirus pubis) have been
found in countries such as Britain, Iceland, Greenland, the Netherlands
and Italy (Kenward, 2001; Mumcuoglu, 2008; Fornaciari et al., 2009; Forbes
et al., 2013). Fleas (Pulex irritans) were also widespread and have been recov-
ered from excavations in Britain, Iceland, Ireland, France, Greenland and the
Netherlands (Buckland et al., 1998; Kenward and Hall, 1995; Sadler, 1990;
Schelvis, 1994; Yvinec et al., 2000). Bed bugs (Cimex lectularius) have been
identified in medieval Britain as well as other time periods (Kenward and
Allison, 1994; Bain, 2004).
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 409

Some studies have assessed the number of ectoparasites in sediments in


the same town over the centuries in order to see if they were becoming
more or less prevalent over time. In York (Britain), the concentration of
ectoparasites in Roman, Viking and medieval layers did not seem to show
any significant changes. For example, human lice (Pediculus humanus) were
found in over 50 archaeological layers and fleas (P. irritans) were also consis-
tently found (Kenward and Hall, 1995, pp. 488e91 and 700e3; Hall and
Kenward, 2015, p. 114). While it is hard to know for sure how representa-
tive these findings may have been for other regions of medieval Europe, the
similar attitudes to personal hygiene at that time means there may well have
been similar problems with infection by human ectoparasites across northern
Europe in the past.
While skin itching will often be annoying to those infected with ectopar-
asites, it is thought that head lice, public lice and the European bed bug do not
spread infectious diseases. However the body louse can spread epidemic
typhus (Rickettsia prowazekii), trench fever (Bartonella quintana), louse-borne
relapsing fever (Borrelia recurrentis) and potentially bubonic plague (Gunn
and Pitt, 2012, p. 157; Houhamdi et al., 2006). The ancient DNA of
B. quintana, the agent of trench fever, has been identified from dental pulp
of the teeth of medieval skeletons from both eleventh- to fifteenth-century
Bondy in France and fifteenth- to sixteenth-century Venice in Italy (Tran
et al., 2011a,b). Similarly, ancient DNA analysis of French soldiers and their
lice in a Napoleonic War mass grave in Vilnius (Lithuania) dating from 1812
has confirmed the presence of both epidemic typhus and trench fever in the
army (Raoult et al., 2006). Three out of five body lice analysed were positive
for B. quintana (the agent of trench fever). Dental pulp from the teeth of 35
soldiers was also tested. Of these, seven were positive for B. quintana and a
further three soldiers were positive for R. prowazekii (the agent of epidemic
typhus). It seems highly probably that the same infections were spread by
body lice in medieval Europe too.
Bites from rodent fleas such as Xenopsylla cheopis have been implicated in
the spread of bubonic plague (Yersinia pestis) to humans in recent outbreaks.
It is thought that when the rats die from plague infection their hungry fleas
have little choice but to feed from humans, and this spreads the disease
(Schotthoefer et al., 2011; Chouikha and Hinnebusch, 2012; Bobrov
et al., 2014). The Black Death was the second pandemic of bubonic plague,
and swept across medieval Europe from 1348 to 1352. It is thought that 30e
60% of the populations died as a result (Green, 2014; Carmichael, 2014).
Ancient DNA analysis of skeletons in Black Death mass graves has confirmed
410 Piers D. Mitchell

that the pandemic was caused by Y. pestis (Bos et al., 2011; Cui et al., 2013).
However, it is not entirely clear if rat fleas were also responsible for the
spread of the Black Death between humans. Medieval written records do
not describe dead rats in the streets as generally happens in recent epidemics
(Slack, 1989; Cohn, 2008). Bubonic plague swept across fourteenth-century
Scandinavia just as it did elsewhere on the continent, but archaeological
studies have shown the black rat (Rattus rattus) to have been rare or
completely absent in Nordic countries at that time (Hufthammer and
Walløe, 2013). If there were few or no rats in the region, this would suggest
it could not have been the bite of the rat flea that was spreading the
infection. It has been argued that in the Black Death pandemic the Yersinia
bacteria may have been spread instead by fleas normally found on other
mammals including humans (e.g. P. irritans), the body louse (Pediculus
humanus humanus), or as pneumonic plague by coughing (Houhamdi
et al., 2006; Hufthammer and Walløe, 2013; Carmichael, 2014).

3.7 Medical history of medieval parasites


Hippocrates was an ancient Greek medical practitioner who lived on the
island of Kos in the Aegean Sea during the fifth and fourth centuries
BC. He and his students wrote a number of medical texts that attempted
to give a framework to their understanding of disease at that early time
(Mann, 2012; Totelin, 2009). This is relevant to the study of the medieval
period as the Hippocratic Corpus of texts was still regarded as the funda-
mental basis for the understanding of disease and its treatment right
through the medieval period. Several of these texts in the Hippocratic
Corpus describe what we now know as parasites, although they did not
realize this at the time. Three worms were described, which were referred
to in ancient Greek as ‘helmins strongyle’, helmins plateia’ and ‘ascaris’.
Modern assessment of their descriptions has suggested these may be equiv-
alent to roundworm (A. lumbricoides), tapeworm (Taenia sp.) and pinworm
(E. vermicularis) (Trompoukis et al., 2007).
Maurus of Salerno was head master of the famous medieval centre of
medical learning at Salerno in Italy from the 1170s to the 1190s (Moulinier
et al., 2007; Nutton, 1971). He wrote his Commentary on the Prognostics of
Hippocrates, to give his opinions on some of the key ideas Hippocrates and
his students had written one and a half thousand years previously (Saffron,
1972). In this text he described intestinal worms as being round or flat,
and as being long or short. He did not regard these worms as being a disease
or parasite, but rather a sign that there was an imbalance of the four humours
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 411

in the body that had been defined by Hippocrates (blood, phlegm, black bile
and yellow bile). Maurus believed that when there was an excess of phlegm
because the person was not eating the right balance of foods or was getting
elderly, the body would try and get rid of the phlegm by producing worms
and expelling them. Hence worms were not regarded as a disease in their
own right, but rather as a sign of humoural imbalance due to excess phlegm
(Saffron, 1972). Medieval medical practitioners would try to treat this imbal-
ance by modifying the diet, by bloodletting and with medicines that were
thought to reduce the levels of phlegm in the body and so return the
humoural balance to normal (Harington, 1953; York, 2012; Mitchell,
2014).
Medieval texts sometimes describe outbreaks of dysentery in armies on
campaign. The 7th Crusade took place from 1248 to 1252, and armies
from across Europe led by King Louis IX of France invaded Egypt. After
some months of campaigning the army was surrounded by Egyptian troops,
trapped between canals clogged with corpses (Bartlett, 2007; Jackson, 2007).
Personal hygiene, latrines and clean drinking water were not available, and
widespread dysentery took hold. We hear how King Louis had diarrhoea so
frequently that he had the bottom of his breeches cut off, so his liquid stool
could run down his legs without having to remove his clothes (Hague, 1955,
p. 24). The fact that Louis was reduced to this low level of personal hygiene,
compared with that normally expected for a medieval French king, high-
lights the challenging position of the whole army. It shows how accepted
attitudes to smells, sights and sounds can be lost in desperate times. The
army later surrendered to the Egyptians, and after ransoming themselves
they travelled to the city of Acre, a crusader city on the Mediterranean coast
of the Frankish Kingdom of Jerusalem. ELISA analysis of thirteenth-century
sediments from the latrines of the crusading Order of St. John at Acre has
found evidence for two parasites that can cause dysentery, E. histolytica
and G. duodenalis (Mitchell et al., 2008b). It is possible that either or both
of these species may have been responsible for the outbreak of dysentery
on the 7th Crusade.
There is some evidence for the use of topical medications containing
mercury to treat ectoparasites in 1400s Europe. Ferdinand II of Aragon
was the King of Naples in Italy, and he lived from 1467 to 1496. Analysis
of his naturally mummified body shows that he was infected with both
head lice (P. humanus capitis) and pubic lice (P. pubis). Tests of the hair
from his head showed very high levels of mercury (827 parts per million),
suggestive of the topical application of a mercury-containing medicine to
412 Piers D. Mitchell

try and kill off the head lice. However, his pubic hair did not show these
very high levels (10 parts per million), implying that he did not undergo
treatment for his pubic lice (Fornaciari et al., 2011).

4. CONCLUSION
In the medieval period, Europe was a diverse continent in terms of its
geography, climate, population density and lifestyles. This resulted in quite
different patterns of parasitism in different regions. Past work has tended to
amalgamate the evidence for the continent in order to gain an overview, and
to provide a data source with which to compare with parasitism in other
continents. Here we do the opposite, by exploring the differences within
the continent and attempting to understand why these differences existed.
This approach has shown that roundworm and whipworm were found
right across the continent, with little evidence for significant variability
between regions. Despite the modern evidence that latrine use decreases the
prevalence of these faecal-oral parasites, there does not seem to be any archae-
ological evidence for a decrease in these parasites with increasing numbers of
latrines. It is quite possible that the widespread use of human faeces as a crop
fertilizer in medieval Europe meant that food plants became contaminated
with parasite eggs and this negated any health benefits from latrine use.
However, a number of parasite species demonstrated marked variability
in prevalence across Europe. In some cases it was due to local culinary prac-
tices, such as the eating of raw, smoked or pickled fish predisposing the pop-
ulation to widespread fish tapeworm infection in many parts of northern
Europe. Farming herbivores such as pigs, cattle and sheep seems to have
facilitated the spread of parasites such as pork tapeworm, beef tapeworm
and liver fluke to medieval people. In regions where herding sheep was
common, and dogs were widespread, we see evidence for hydatid disease
in humans too. The medieval archaeological evidence for hydatid cysts
seems to be concentrated in northern Europe.
Migrations can be detected when parasites are identified in locations
incompatible with their life cycle, or where they had not previously been
identified. Examples of such parasites include two species of schistosomiasis
in France, and both fish tapeworm and entamoeba dysentery in crusaders
and Mamluk period pilgrims.
A range of ectoparasites have been recovered from medieval Europe.
Some would have just led to itchy skin and annoyance, but others have
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 413

been implicated in the spread of a range of bacterial infectious diseases such


as epidemic typhus, trench fever, louse-borne relapsing fever and bubonic
plague. In these circumstances, human ectoparasites clearly played a major
role in some of the most serious disease outbreaks of the Middle Ages.
In contrast with today, medieval medical practitioners viewed both
helminths and ectoparasites in the context of humoural theory. Parasites
were regarded as material created as the body tried to remove excess phlegm,
and not as an infective organism that could be spread from one host to
another. There is evidence for medical treatments including dietary modifi-
cation, bloodletting, oral medications and also topical medications contain-
ing mercury for ectoparasites.

REFERENCES
Anastasiou, E., 2015. Parasites in European populations from prehistory to the industrial
revolution. In: Mitchell, P.D. (Ed.), Sanitation, Latrines and Intestinal Parasites in Past
Populations. Ashgate, Farnham, pp. 203e217.
Anastasiou, E., Mitchell, P.D., 2013a. Evolutionary anthropology and genes: investigating
the genetics of human evolution from excavated skeletal remains. Gene 528 (1), 27e32.
Anastasiou, E., Mitchell, P.D., 2013b. Human intestinal parasites from a latrine in the 12th
century Frankish castle of Saranda Kolones in Cyprus. Int. J. Paleopathol. 3, 218e223.
Antoniou, G.P., Angelakis, A.N., 2015. Latrines and wastewater sanitation technologies in
ancient Greece. In: Mitchell, P.D. (Ed.), Sanitation, Latrines and Intestinal Parasites in
Past Populations. Ashgate, Farnham, pp. 41e68.
Araujo, A., Reinhard, K., Bastos, O.M., Costa, L.C., Pirmez, C., Iniguez, A.M.,
Vicente, A.C., Morel, C.M., Ferreira, L.F., 1998. Paleoparasitology: perspectives with
new techniques. Rev. Inst. Med. Trop. S~ao Paulo 40, 371e376.
Araujo, A., Ferreira, L.F., Fugassa, M., Leles, D., Sianto, L., de Souza, S.M.M., Dutra, J.,
I~
niguez, A., Reinhard, K., 2015a. New World paleoparasitology. In: Mitchell, P.D.
(Ed.), Sanitation, Latrines and Intestinal Parasites in Past Populations. Ashgate, Farnham,
pp. 165e202.
Araujo, A., Reinhard, K., Ferreira, L.F., 2015b. Paleoparasitology e human parasites in
ancient material. Adv. Parasitol. 90, 349e387.
Araujo, A., Reinhard, K., Ferreira, L.F., Gardner, S., 2008. Parasites as probes for prehistoric
migrations? Trends Parasitol. 24, 112e115.
Araujo, A., Reinhard, K., Leles, D., Sianto, L., Iniguez, A., Fugassa, M., Arriaza, B.,
Orellana, N., Ferreira, L.F., 2011. Paleoepidemiology of intestinal parasites and lice in
pre-Columbian South America. Chungara Rev. Antropol. Chil. 42 (2), 303e313.
Astill, G., Langdon, J. (Eds.), 1997. Medieval Farming and Technology: The Impact of
Agricultural Change in Northwest Europe. Brill, Leiden.
Baeten, J., Marinova, E., De Laet, V., Degryse, P., De Vos, D., Waelkens, M., 2012. Faecal
biomarker and archaeobotanical analyses of sediments from a public latrine shed new
light on ruralisation in Sagalassos, Turkey. J. Archaeol. Sci. 39 (4), 1143e1159.
Bain, A., 2004. Irritating intimates: the archaeoentomology of lice, fleas and bed bugs.
Northeast Hist. Archaeol. 33 (1), 81e90.
Barrett, J.H., Locker, A.M., Roberts, C.M., 2004. The origins of intensive marine fishing in
medieval Europe: the English evidence. Proc. R. Soc. B: Biol. Sci. 271, 2417e2421.
Barrett, J.H., Orton, D., Johnstone, C., Harland, J., van Neer, W., Ervynck, A., Roberts, C.,
Locker, A., Amundsen, C., Enghoff, I.B., Hamilton-Dyer, S., Heinrich, D.,
414 Piers D. Mitchell

Hufthammer, A.K., Jones, A.K.G., Jonsson, L., Makowiecki, D., Pope, P.,
O’Connell, T.C., de Roo, T., Richards, M., 2011. Interpreting the expansion of sea
fishing in medieval Europe using stable isotope analysis of archaeological cod bones.
J. Archaeol. Sci. 38 (7), 1516e1524.
Bartlett, W.B., 2007. The Last Crusade: The Seventh Crusade and the Final Battle for the
Holy Land. Tempus Books, Stroud.
Bianucci, R., Mattutino, G., Lallo, R., Charlier, P., Jouin-Spriet, H., Peluso, A., Higham, T.,
Torre, C., Massa, A.R., 2008. Immunological evidence of Plasmodium falciparum infec-
tion in an Egyptian child mummy from the Early Dynastic period. J. Archaeol. Sci.
35, 1880e1885.
Bobrov, A.G., Kirillina, O., Vadyvaloo, V., Koestler, B.J., Hinz, A.K., Mack, D.,
Waters, C.M., Perry, R.D., 2014. The Yersinia pestis HmsCDE regulatory system is
essential for blockage of the oriental rat flea (Xenopsylla cheopis), a classic plague vector.
Environ. Microbiol. 17, 947e959. http://dx.doi.org/10.1111/1462-2920.12419.
Boersma, J.H., Jansen, J., 1975. Helminth infection in medieval Utrecht. Trop. Geogr. Med.
27, 441.
Bos, K.I., Schuenemann, V.J., Golding, G.B., Burbano, H.A., Waglechner, N.,
Coombes, B.K., McPhee, J.B., DeWitte, S.N., Meyer, M., Schemes, S., Wood, J.,
Earn, D.J.D., Herring, D.A., Bauer, P., Poinar, H.N., Krause, J., 2011. A draft genome
of Yersinia pestis from victims of the black death. Nature 478 (7370), 506e510.
Bouchet, F., 1995. Recovery of helminth eggs from archaeological excavation of the Grand
Louvre (Paris, France). J. Parasitol. 81, 785e787.
Bouchet, F., 1997. Intestinal capillariasis in neolithic inhabitants of Chalain (Jura, France).
Lancet 349, 256.
Bouchet, F., Harter, S., Paicheler, J.C., Ara ujo, A., Ferreira, L.F., 2002. First recovery of
Schistosoma mansoni eggs from a latrine in Europe (15the16th centuries). J. Parasitol.
88 (2), 404e405.
Bouchet, F., Baffier, D., Girard, M., Morel, P., Paicheler, J.C., David, F., 1996. Paléopara-
sitologie en contexte Pleistocene: premiere observations a la Grande Grotte d’Arcy-sur-
Cure (Yonne), France. C. R. Acad. Sci. 319, 147e151.
Bouchet, F., Harter, S., Le Bailly, M., 2003. The state of the art of palaeoparasitology research
in the Old World. Mem. Inst. Oswaldo Cruz 98 (Suppl. 1), 92e102.
Bouchet, F., Paicheler, J.C., 1995. Paleoparasitologie: présumption d’un cas de bilharziose au
XVe siecle a Montbéliard (Doubs, France). C. R. Acad. Sci. 318, 811e814.
Bryant, V.M., Dean, G.W., 2006. Archaeological coprolite science: the legacy of Eric O.
Callen (1912-1970). Palaeogeogr. Palaeoclimatol. Palaeoecol. 237, 51e66.
Buckland, P.C., Buckland, P.I., Skidmore, P., 1998. Insect remains from GUS: and interim
report. In: Arneborg, J., Gulløv, H.C. (Eds.), Man, Culture and Environment in Ancient
Greenland. Dansk Polar Centre & Danish National Museum, Copenhagen, pp. 74e79.
Buxton, M., 1998. Fish-eating in medieval England. In: Walker, H. (Ed.), Fish: Food from
the Waters. Prospect Books, Totnes, pp. 51e59.
Carmichael, A.G., 2014. Plague persistence in Western Europe: a hypothesis. Medieval
Globe 1, 157e191.
Chouikha, I., Hinnebusch, B.J., 2012. Yersinia-flea interactions and the evolution of the
arthropod-borne transmission route of plague. Curr. Opin. Microbiol. 15 (3), 239e246.
Cohn, S.K., 2008. Epidemiology of the Black Death and successive waves of plague. Med.
Hist. 52 (Suppl. 27), 74e100.
Cox, F.E.G., 2002. History of human parasitology. Clin. Microbiol. Rev. 15, 595e612.
Cui, Y., Yu, C., Yan, Y., Li, D., Li, Y., Jombart, T., Weinert, L.A., Wang, Z., Guo, Z.,
Xu, L., Zhang, Y., Zheng, H., Qin, N., Xiao, X., Wu, M., Wang, X., Zhou, D.,
Qi, Z., Du, Z., Wu, H., Yang, X., Cao, H., Wang, H., Wang, J., Yao, S., Rakin, A.,
Li, Y., Falush, D., Balloux, F., Achtman, M., Song, Y., Wang, J., Yang, R., 2013.
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 415

Historical variations in mutation rate in an epidemic pathogen, Yersinia pestis. Proc. Natl.
Acad. Sci. U.S.A. 110 (2), 577e582.
Da Rocha, G.C., Harter, S., Le Bailly, M., Ara ujo, A., Ferreira, L.F., Serra-Freire, M.,
Bouchet, F., 2006. Paleoparasitological remains revealed by seven historic contexts from
“Place d’Armes”, Namur, Belgium. Mem. Inst. Oswaldo Cruz 101 (Suppl. 2), 43e52.
Deelder, A.M., Miller, R.L., de Jonge, N., Krijger, F.W., 1990. Detection of schistosome
antigen in mummies. Lancet 335, 724e725.
Dentzien-Dias, P.C., Poinar Jr., G., De Figueiredo, A.E.Q., Pacheco, A.C.L., Horn, B.L.D.,
Schultz, C.L., 2013. Tapeworm eggs in a 270 million-year-old shark coprolite. PLoS
One 8, e55007.
De Rouffignac, C., 1985. Parasite egg survival and identification from Hibernia Warf,
Southwark. Lond. Archaeol. 5, 103e105.
Dittmar, K., 2009. Old parasites for a new world: the future of paleoparasitological research: a
review. J. Parasitol. 95, 365e371.
Dounias, E., Froment, A., 2006. When forest-based hunter-gatherers become sedentary:
consequences for diet and health. Unasylva 57, 26e33.
Faith, R., 2010. Farms and families in ninth-century Provence. Early Medieval Eur. 18 (2),
175e201.
Florenzano, A., Mercuri, A.M., Pederzoli, A., Torri, P., Bosi, G., Olmi, L., Rinaldi, R.,
Mazzanti, M.B., 2012. The significance of intestinal parasite remains in pollen samples
from medieval pits in the Piazza Garibaldi of Parma, Emilia Romagna, Northern Italy.
Geoarchaeology 27 (1), 34e47.
Forbes, V., Dussault, F., Bain, A., 2013. Contributions of ectoparasite studies in archaeology
with two examples from the North Atlantic region. Int. J. Paleopathol. 3 (3), 158e164.
Fornaciari, G., Giuffra, V., Marinozzi, S., Picchi, M.S., Masetti, M., 2009. “Royal” pedicu-
losis in Renaissance Italy: lice in the mummy of the King of Naples Ferdinand II of
Aragon (1467-1496). Mem. Inst. Oswaldo Cruz 104, 671e672.
Fornaciari, G., Marinozzi, S., Gazzaniga, V., Giuffra, V., Picchi, M.S., Giusiani, M.,
Masetti, M., 2011. The use of mercury against pediculosis in the Renaissance: the case
of Ferdinand II of Aragon, King of Naples, 1467-91. Med. Hist. 55, 109e115.
Frías, L., Leles, D., Ara ujo, A., 2013. Studies on protozoa in ancient remains e a review.
Mem. Inst. Oswaldo Cruz 108, 1e12.
Fugassa, M.H., Ara ujo, A., Guichon, R.A., 2006. Quantitative paleoparasitology applied to
archaeological sediments. Mem. Inst. Oswaldo Cruz 101 (Suppl. 2), 29e33.
Garcia, L.S., 2009. Practical Guide to Diagnostic Parasitology, second ed. ASM Press,
Washington.
Gates, C., 2011. Ancient Cities: The Archaeology of Urban Life in the Ancient Near East and
Egypt, Greece and Rome, second ed. Routledge, Abingdon.
Gonçalves, M.L.C., Ara ujo, A., Ferreira, L.F., 2003. Human intestinal parasites in the past:
new findings and a review. Mem. Inst. Oswaldo Cruz 98, 103e118.
Gonçalves, C.L.M., da Silva, V.L., de Andrade, C.M., Reinhard, K., de Rocha, G.C., Le
Bailly, M., Bouchet, F., Ferreira, L.F., Ara ujo, A., 2004. Amoebiasis distribution in
the past: first steps using an immunoassay technique. Trans. R. Soc. Trop. Med. Hyg.
98, 88e91.
Green, M.H., 2014. Taking “pandemic” seriously: making the black death global. Medieval
Globe 1, 27e61.
Gunn, A., Pitt, S.J., 2012. Parasitology: An Integrated Approach. Wiley-Blackwell,
Chichester.
Hagen, A., 2006. Anglo-Saxon Food and Drink: Production, Processing, Distribution and
Consumption. Anglo-Saxon Books, Hockwold Cum Wilton, pp. 275e276.
Hague, R. (Ed.), 1955. The Life of Saint Louis, by John of Joinville. Sheed and Ward,
London.
416 Piers D. Mitchell

Hall, A.R., Kenward, H.K., 2015. Sewers, cesspits and middens: a survey of the evidence for
2000 years of waste disposal in York, UK. In: Mitchell, P.D. (Ed.), Sanitation, Latrines
and Intestinal Parasites in Past Populations. Ashgate, Farnham, pp. 99e119.
Hansen, J., 1986. Les momies du Groenland. La Rech. 183, 1490e1498.
Harington, J. (Ed.), 1953. Regimen Sanitatis Salerni: The School of Salernum. Ente Provin-
ciale Per Il Tourismo, Salerno.
Heinonen-Tanski, H., van Wijk-Sebesma, C., 2005. Human excreta for plant production.
Bioresour. Technol. 96, 403e411.
Herrmann, B., 1985. Parasitologisch-epidemiologische auswertungen mittelalterlicher
kloaken. Z. Arch€aologie des Mittelaltes 13, 131e161.
Herrmann, B., 1988. Parasite remains from medieval latrine deposits: an epidemiological
and ecologic approach. Actes des Troisiemes J. Anthropol. Notes Monogr. Tech. 24,
135e142.
Hoberg, E.P., Alkire, N.L., Queiroz, A., Jones, A., 2001. Out of Africa: origins of the Taenia
tapeworms in humans. Proc. R. Soc. B Biol. Sci. 268, 781e787.
Hoffman, R.C., 2003. A brief history of aquatic resource use in medieval Europe. Helgol.
Mar. Res. 59 (1), 22e30.
Holmes, T., 2009. Early Humans: The Pleistocene and Holocene Epochs. Chelsea House
Publishers, New York.
Hoskins, W.G., 1955. Sheep Farming in Saxon and Medieval England. Department of
Education of the International Wool Secretariat, London.
Houhamdi, L., Lepidi, H., Drancourt, M., Raoult, D., 2006. Experimental model to evaluate
the human body louse as a vector of plague. J. Infect. Dis. 194 (11), 1589e1596.
Hufthammer, A.K., Walløe, L., 2013. Rats cannot have been intermediate hosts for Yersinia
pestis during medieval plague epidemics in northern Europe. J. Archaeol. Sci. 40 (4),
1752e1759.
Hugot, J.-P., Gardner, S., Borba, V., Ara ujo, P., Leles, D., Stock Da-Rosa, A., Dutra, J.,
Ferreira, L., Ara ujo, A., 2014. Discovery of a 240 million year old nematode parasite
egg in a cynodont coprolite sheds light on the early origin of nematode parasites in
vertebrates. Parasites & Vectors 7, 486.
Hurst, D., 2005. Sheep in the Cotswolds: The Medieval Wool Trade. Tempus Books,
Stroud, p. 57.
Jackson, P., 2007. The Seventh Crusade, 1244e1254: Sources and Documents. Ashgate,
Aldershot.
Jensen, P.K., Phuc, P.D., Knudsen, L.G., Dalsgaard, A., Konradsen, F., 2008. Hygiene versus
fertiliser: the use of human excreta in agriculture e a Vietnamese example. Int. J. Hyg.
Environ. Health 211, 432e439.
Jones, A.K.G., 1982. Human parasite remains: prospects for a quantitative approach. In:
Hall, A.R., Kenward, K.H. (Eds.), Environmental Archaeology in the Urban
Context. The Council for British Archaeology, London, pp. 66e70.
Jones, A.K.G., 1983. A coprolite from 6-8 Pavement. In: Hall, A.R., Kenward, H.K.,
Williams, D., Grieg, J.R.A. (Eds.), Environment and Living Conditions at Two
Anglo-Scandinavian Sites, The Archaeology of York series, vol. 14(4). Council for
British Archaeology, London, pp. 225e229.
Jones, A.K.G., 1988. Parasitological investigations. In: Schia, E., Griffin, K. (Eds.), De
Arkeologiske Utgravninger I Gamlegyen, Oslo, Series no. 5. Alheim and Eide,
Oslo, pp. 134e137.
Jones, R. (Ed.), 2012. Manure Matters: Historical, Archaeological and Ethnographic
Perspectives. Ashgate, Farnham.
Kelly, F., 1997. Early Irish Farming: A Study Based Mainly on the Law Texts of the 7th and
8th Century. Dublin Institute for Advanced Studies, Dublin.
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 417

Kenward, H., 2001. Pubic lice in Roman and medieval Britain. Trends Parasitol. 17 (4),
167e168.
Kenward, H., Allison, E., 1994. Rural origins of the urban insect fauna. In: Hall, A.R.,
Kenward, H.K. (Eds.), Urban-Rural Connexions: Perspectives from Environmental
Archaeology, Oxbow monograph, vol. 47. Oxbow, Oxford, pp. 55e77.
Kenward, H.K., Hall, A.R., 1995. Biological Evidence from 16e22 Coppergate. Archae-
ology of York series no.14/7. Council for British Archaeology, York.
Kristjansd
ottir, S., Collins, C., 2011. Cases of hydatid disease in Medieval Iceland. Int. J.
Osteoarchaeol. 21, 479e486.
Lawson, D.W., Alvergne, A., Gibson, M.A., 2012. The life-history trade off between fertility
and child survival. Proc. R. Soc. Part B Biol. Sci. 279, 4755e4764.
Le Bailly, M., Bouchet, F., 2006. Paléoparasitologie et immunologie: l’exemple d’Entamoeba
histolytica. Archéosci. Rev. Archaeom. 30, 129e135.
Le Bailly, M., Bouchet, F., 2010. Ancient dicrocoeliasis: occurrence, distribution and
migration. Acta Trop. 115, 175e180.
Le Bailly, M., Bouchet, F., 2013. Diphyllobothrium in the past: review and new records. Int. J.
Paleopathol. 3, 182e187.
Le Bailly, M., Bouchet, F., 2015. A first attempt to retrace the history of dysentery caused by
Entamoeba histolytica. In: Mitchell, P.D. (Ed.), Sanitation, Latrines and Intestinal Parasites
in Past Populations. Ashgate, Farnham, pp. 219e228.
Le Bailly, M., Gonçalves, C.L.M., Harter, S., Prodeo, F., Ara ujo, A., Bouchet, F., 2008. New
finding of Giardia intestinalis (Eukaryote, Metamonad) in Old World archaeological site
using immunofluorescence and enzyme-linked immunosorbent assays. Mem. Inst.
Oswaldo Cruz 103, 298e300.
Le Bailly, M., Leuzinger, U., Bouchet, F., 2003. Dioctophymidae eggs in coprolites from
Neolithic site of Arbon-Bleiche 3 (Switzerland). J. Parasitol. 89 (5), 1073e1076.
Lewicke, P.B., 2011. Food and Foodways of Medieval Cairenes. Brill, Ledien, pp. 209e225.
Lucas, A.T., 1989. Cattle in Ancient Ireland. Boethius Press, Kilkenny.
Lynnerup, N., 2007. Mummies. Yearb. Phys. Anthropol. 50, 162e190.
Magnusson, R.J., 2013. Medieval urban environmental history. Hist. Compass 11 (3),
189e200.
Mann, J.E., 2012. Hippocrates, on the Art of Medicine. Brill, Leiden.
McMahon, A., 2015. Waste management in early urban southern Mesopotamia. In:
Mitchell, P.D. (Ed.), Sanitation, Latrines and Intestinal Parasites in Past Populations. Ash-
gate, Farnham, pp. 19e39.
Mellars, P., 2004. Neanderthals and the modern human colonization of Europe. Nature 432,
461e465.
Mitchell, P.D., 2011a. Retrospective diagnosis, and the use of historical texts for investigating
disease in the past. Int. J. Paleopathol. 1, 81e88.
Mitchell, P.D., 2011b. The spread of disease with the crusades. In: Nance, B., Glaze, E.F.
(Eds.), Between Text and Patient: The Medical Enterprise in Medieval and Early Mod-
ern Europe. Sismel, Florence, pp. 309e330.
Mitchell, P.D., 2013. The origins of human parasites: exploring the evidence for endopara-
sitism throughout human evolution. Int. J. Paleopathol. 3, 191e198.
Mitchell, P.D., 2014. Bloodletting: when, where and how. In: Dourish, E. (Ed.), Early
Printed Treasures from Cambridge University Library. Cambridge University Press,
Cambridge, p. 142.
Mitchell, P.D. (Ed.), 2015. Sanitation, Latrines and Intestinal Parasites in Past Populations.
Ashgate, Farnham.
Mitchell, P.D., Anastasiou, E., Syon, D., 2011. Human intestinal parasites in crusader Acre: evi-
dence for migration with disease in the Medieval Period. Int. J. Paleopathol. 1, 132e137.
418 Piers D. Mitchell

Mitchell, P.D., Huntley, J., Sterns, E., 2008a. Bioarchaeological analysis of the 13th century
latrines of the crusader hospital of St. John at Acre, Israel. In: Mallia-Milanes, V. (Ed.),
The Military Orders: Volume 3. Their History and Heritage. Ashgate, Aldershot,
pp. 213e223.
Mitchell, P.D., Stern, E., Tepper, Y., 2008b. Dysentery in the crusader kingdom of Jerusa-
lem: an ELISA analysis of two medieval latrines in the city of Acre (Israel). J. Archaeol.
Sci. 35 (7), 1849e1853.
Mitchell, P.D., Stern, E., 2001. Parasitic intestinal helminth ova from the latrines of the 13th
century crusader hospital of St. John in Acre, Israel. In: La Verghetta, M., Capasso, L.
(Eds.), Proceedings of the XIIIth European Meeting of the Paleopathology Association,
Chieti Italy. Edigrafital S.p.A., Teramo, pp. 207e213.
Mitchell, P.D., Tepper, Y., 2007. Intestinal parasitic worm eggs from a crusader period cess-
pool in the city of Acre (Israel). Levant 39, 91e95.
Mitchell, P.D., Yeh, H.-Y., Appleby, J., Buckley, R., 2013. The intestinal parasites of King
Richard III. Lancet 382, 888.
Mosquera, M., Ollé, A., Rodríguez, X.P., 2013. From Atapuerca to Europe: tracing the
earliest peopling of Europe. Quat. Int. 295, 130e137.
Moulinier, L., Jacquart, D., Paravicini Bagliani, A., 2007. Le science des urines de Maurus de
Salerne et les Sinthomata Magistri Mauri inedits. In: Jacquart, D., Paravicini Bagliani, A.
(Eds.), La Scuola Medica Salernitana. Sismel, Florence, pp. 261e282.
Mumcuoglu, K.Y., 2008. Human lice: pediculus and pthirus. In: Raoult, D.,
Drancourt, M. (Eds.), Paleomicrobiology: Past Human Infections. Springer-Verlag,
Berlin, pp. 215e222.
Nansen, P., Jorgensen, R.J., 1977. Fund af parasitæg i arkæologisk material fra det vikinge-
tidige Ribe. Nord. Veterinaer Med. 29, 263e266.
Naumann, E., Price, T.D., Richards, M.P., 2014. Changes in dietary practices and social
organization during the pivotal late iron age period in Norway (AD 550-1030): isotope
analyses of merovingian and viking age human remains. Am. J. Phys. Anthropol. 155,
322e331.
Nicholson, H., 2007. The Knights Hospitaller. Boydell Press, Woodbridge.
Nutton, V., 1971. Velia and the school of Salerno. Med. Hist. 15 (1), 1e11.
O’Connor, T.P., 2008. Pets and pests in Roman and medieval Britain. Mammal Rev. 22 (2),
107e113.
Pasternak, C. (Ed.), 2007. What Makes Us Human. Oneworld, Oxford.
Phuc, P.D., Konradsen, F., Phuong, P.T., Cam, P.D., Dalsgaard, A., 2006. Practice of using
human excreta as fertilizer and implications for health in Nghean Province, Vietnam.
Southeast Asian J. Trop. Med. Public Health 37, 222e229.
Pike, A.W., 1967. The recovery of parasite eggs from ancient cesspit and latrine deposits: an
approach to the study of early parasite infections. In: Brothwell, D., Sandison, A.T.
(Eds.), Diseases in Antiquity. C.C. Thomas, Springfield, pp. 184e188.
Poinar, G., Boucot, A.J., 2006. Evidence of intestinal parasites of dinosaurs. Parasitology 133,
245e249.
Price, J.L., 1975. The radiology of excavated saxon and medieval human remains from
Winchester. Clin. Radiol. 26, 363e370.
Raoult, D., Dutour, O., Houhamdi, L., Jankauskas, R., Fournier, P.-E., Ardagna, Y.,
Drancourt, M., Signoli, M., Dang La, V., Macia, Y., Aboudharam, G., 2006. Evidence
for louse-transmitted diseases in soldiers of Napoleon’s Grand Army in Vilnius. J. Infect.
Dis. 193, 112e120.
Racz, E.E., Pucu de Araujo, E., Jensen, E., Mostek, C., Morrow, J.J., Van Hove, M.L.,
Bianucci, R., Willems, D., Heller, F., Araujo, A., Reinhard, K.J., 2015. Parasitology
in an archaeological context: analysis of medieval burials in Nivelles, Belgium.
J. Archaeol. Sci. 53, 304e315.
Human Parasites in Medieval Europe: Lifestyle, Sanitation and Medical Treatment 419

Reinhard, K., Bryant, V.M., 1992. Coprolite analysis: a biological perspective on


archaeology. In: Schiffer, M.B. (Ed.), Advances in Archaeological Method and
Theory. University of Arizona, Tucson, pp. 245e288.
Reinhard, K., Confalonieri, U., Ferreira, L.F., Herrmann, B., Ara ujo, A., 1986. Recovery of
parasite remains from coprolites and latrines: aspects of paleoparasitological technique.
Homo 37, 217e239.
Reinhard, K., Ferreira, L.F., Bouchet, F., Sianto, L., Dutra, J., I~niguez, A.M., Leles, D., Le
Bailly, M., Fugassa, M., Pucu, E., Ara ujo, A., 2013. Food, parasites, and epidemiological
transitions: a broad perspective. Int. J. Paleopathol. 86, 1e8.
Reinhard, K., Geib, P.R., Callahan, M.M., Hevley, R.H., 1992. Discovery of colon
contents in a skeletonised burial: soil sampling for dietary remains. J. Archaeol. Sci.
19, 697e705.
Reinhard, K., Pucu de Ara ujo, E., 2014. Comparative parasitological perspectives on epide-
miological transitions: the Americas and Europe. In: Zuckerman, M.K. (Ed.), Modern
Environments and Human Health: Revisiting the Second Epidemiological Transition.
Wiley Blackwell, Chichester, pp. 311e326.
Riley-Smith, J., 2012. The Knights Hospitaller in the Levant, c.1070e1309. Palgrave,
Basingstoke.
Roever-Bonnet, H., Rijpstra, C., Van Renesse, M.A., Peen, C.H., 1979. Helminth eggs and
gregarines from coprolites from the excavations at Swifterbant. Helinium 19, 7e12.
Sabine, E., 1934. Latrines and cesspools of mediaeval London. Speculum 9, 303e321.
Sabine, E., 1937. City cleaning in mediaeval London. Speculum 12, 19e43.
Sadler, J.P., 1990. Records of ectoparasites on humans and sheep from viking-age deposits in
the former western settlement of Greenland. J. Med. Entomol. 27, 628e631.
Saffron, M.H., 1972. Maurus of Salerno: a twelfth century “optimus physicus” with his
commentary on the prognostics of Hippocrates. Trans. Am. Philos. Soc. 62, 5e104.
Schelvis, J., 1994. Caught between the teeth: a review of Dutch finds of archaeological
remains of ectoparasites in combs. In: Proceedings of the Section Experimental and
Applied Entomology: 5th Meeting of Experimental and Applied Entomologists in the
Netherlands. Nederlandse Entomologische Vereniging, Amsterdam, pp. 131e132.
Schotthoefer, A.M., Bearden, S.W., Holmes, J.L., Vetter, S.M., Montenieri, J.A.,
Williams, S.K., Graham, C.B., Woods, M.E., Eisen, R.J., Gage, K.L., 2011. Effects of
temperature on the transmission of Yersinia pestis by the flea, Xenopsylla cheopis, in the
late phase period. Parasites Vectors 4, 191.
Scobie, A., 1986. Slums, sanitation, and mortality. Klio 68, 399e433.
Shillito, L.-M., Bull, I.D., Matthew, W., Almond, M.J., Williams, J.M., Evershed, R.P.,
2011. Biomolecular and micromorphological analysis of suspected faecal deposits at
Neolithic Çatalh€oy€uk, Turkey. J. Archaeol. Sci. 38, 1869e1877.
Sianto, L., Chame, M., Silva, C.S.P., Gonçalves, M.L.C., Reinhard, K., Fugassa, M.,
Araujo, A., 2009. Animal helminths in human archaeological remains: a review of
zoonoses in the past. Rev. Inst. Med. Trop. S~ao Paulo 51 (3), 119e130.
Simmons, A.H., DiBenedetto, K., 2014. Stone Age Sailors: Paleolithic Seafaring in the
Mediterranean. Left Coast Press, Walnut Creek.
Sinniah, B., 1982. Daily egg production of Ascaris lumbricoides: the distribution of eggs in the
faeces and the variability of egg counts. Parasitology 84 (1), 167e175.
Slack, P., 1989. The black death past and present. 2. Some historical problems. Trans. R. Soc.
Trop. Med. Hyg. 83 (4), 461e463.
Stone, L., Lurquin, P.F. (Eds.), 2007. Genes, Culture and Human Evolution: A Synthesis.
Blackwell, Oxford.
Taylor, C., 2015. A tale of two cities: the efficacy of ancient and medieval sanitation methods.
In: Mitchell, P.D. (Ed.), Sanitation, Latrines and Intestinal Parasites in Past Populations.
Ashgate, Farnham, pp. 69e97.
420 Piers D. Mitchell

Terefe, Y., Hailemariam, Z., Menkir, S., Nakao, M., Lavikainen, A., Haukisalmi, V.,
Iwaki, T., Okamoto, M., Ito, A., 2014. Phylogenetic characterisation of Taenia tape-
worms in spotted hyenas and reconsideration of the “Out of Africa” hypothesis of Taenia
in humans. Int. J. Parasitol. 44, 533e541.
Thiem, V.D., Schmidt, W.P., Suziki, M., Tho, L.H., Yanai, H., Ariyoshi, K., Anh, D.D.,
Yoshida, L.M., 2012. Animal livestock and the risk of hospitalised diarrhoea in children
under 5 years in Vietnam. Trop. Med. Int. Health 17, 613e621.
Totelin, L.M.V., 2009. Hippocratic Recipes: Oral and Written Transmission of Pharmaco-
logical Knowledge in Fifth- and Fourth-Century Greece. Brill, Leiden.
Tran, T.-N.-N., Forestier, C.L., Drancourt, D., Raoult, D., Aboudharam, G., 2011a.
Co-detection of Bartonella quintana and Yersinia pestis in an 11th-15th century burial
site in Bondy, France. Am. J. Phys. Anthropol. 145 (3), 489e494.
Tran, T.-N.-N., Signoli, L., Fozzati, G., Aboudharam, G., Raoult, D., Drancourt, M.,
2011b. High throughput, multiplexed pathogen detection authenticates plague waves
in medieval Venice, Italy. PLoS One 6 (3), e16735.
Trompoukis, C., German, C., Falagas, M.E., 2007. From the roots of parasitology: Hippo-
crates’ first scientific observations in helminthology. J. Parasitol. 93 (4), 970e972.
Uga, S., Hoa, N.T., Noda, S., Moji, K., Cong, L., Aoki, Y., Rai, S.K., Fujimaki, Y., 2009.
Parasite egg contamination of vegetables from a suburban market in Hanoi, Vietnam.
Nepal Med. Coll. J. 11, 75e78.
Van Dam, P.J.E.M., 2009. Fish for feast and fast: fish consumption in the Netherlands in the
late Middle Ages. In: Sicking, L., Abreu-Ferreira, D. (Eds.), Beyond the Catch: Fish-
eries of the North Atlantic, the North Sea and the Baltic, 900e1850. Brill, Leiden,
pp. 309e336.
Walker, M., Hall, A., Anderson, R.M., Basanez, M.G., 2009. Density-dependent effects on
the weight of female Ascaris lumbricoides infections of humans and its impact on patterns of
egg production. Parasites & Vectors 2, 11.
Warnock, P.J., Reinhard, K.J., 1992. Methods for extracting pollen and parasite eggs from
latrine soils. J. Archaeol. Sci. 19, 261e264.
Weiss, D.L., Møller-Christensen, V., 1971. Leprosy, echinococcosis and amulets: a study of a
medieval Danish inhumation. Med. Hist. 15, 260e267.
Yeh, H.-Y., Prag, K., Clamer, C., Humbert, J.B., Mitchell, P.D., 2015. Human intestinal
parasites from a Mamluk Period cesspool in the Christian Quarter of Jerusalem: potential
indicators for long distance travel in the 15th century AD. Int. J. Paleopathol. 9, 69e75.
Yeh, H.-Y., Pluskowski, A., Kalejs, U., Mitchell, P.D., 2014. Intestinal parasites in a mid-
14th century latrine from Riga, Latvia: fish tapeworm and the consumption of uncooked
fish in the medieval eastern Baltic region. J. Archaeol. Sci. 49, 83e89.
York, W.H., 2012. Health and Wellness in Antiquity through the Middle Ages. Greenwood
Publishing, Santa Barbara.
Yvinec, J.H., Ponel, P., Beaucournu, J.-C., 2000. Premier apports archéoentomologiques de
l’étude des puces aspects historiques et anthropologiques (Siphonaptera). Bull. Soc. Ento-
mol. Fr. 105 (4), 419e425.
Zeder, M.A., 2011. The origins of agriculture in the Near East. Curr. Anthropol. 52 (S4),
S221eS235.
Zeigelbauer, K., Speich, B., M€ausezahl, D., Bos, R., Keiser, J., Utzinger, J., 2012. Effect of
sanitation on soil-transmitting helminth infection: systematic review and meta-analysis.
PLoS Med. 9, e1001162.
INDEX
‘Note: Page numbers followed by “f” indicate figures and “t” indicate tables.’

A B
Acanthocephala, 109–110, 246–247 Barnacles, 249–252, 250f
Acrothoracica, 265–266 Benthic colonies, 331
Aculeata, 165 Bioarchaeology, 358
aDNA. See Ancient DNA (aDNA) Bioclaustrations, 242, 334–338
Amber Bivalve skeletons, 218–219
Baltic amber, 25, 60–65, 145, 170, 173 Blister beetles, 161–162
Dominican amber, 16t, 65–75, 109–110, Body fossils, 14–17, 55–60, 96–97, 140,
157, 160, 179, 375 203, 237–239, 249, 339
Mexican amber, 75–77, 160 Bopyroids, 237–239
Amblycera, 143 Borings, 292–293, 340–341
Amblyceropsis indica, 145 Brachylophosaurus canadensis, 7
Amphoracrinus gilbertsoni, 300f, 301 Branchiura, 270–271
Ancient biomolecules
aDNA, 6–7 C
palaeoproteomics, 7 Cambrian, 5–6, 79–80, 108–112, 140,
stabilized haem, 6 211–216, 261–262, 340–341
Ancient DNA (aDNA), 4–7, 369 Cambrian–Ordovician species, 29–30,
Ancient parasite discovery 262–264
ancient biomolecules Capillaria sp., 408
aDNA. See Ancient DNA (aDNA) Carboniferous, 82, 105, 113, 140, 161, 173
palaeoproteomics, 7 Carboniferous coprolite. See Coprolites
stabilized haem, 6 Cascofilaria baltica, 64
computed tomography, 5–6 Castexia douvillei, 252–253
See also Computed tomography (CT) Catastrophic climate events, 359–360
destructive/nondestructive methods, Catellocaula vallata, 337
4–5 Caunopores, 335, 335f
thin sections, 5–6 Celleporaria palmata, 338
Anodonta piscinalis, 222–223 Cephalotes serratus, 67
Anoigmaichnus odinsholmensis, 337 Ceratoconcha, 266–267
Anoplura, 142–143 Ceratopogonidae, 152, 157
Antedon bifida, 297 Cercomer theory, 94–95
Antliophora Cestoda, 9t–13t, 94–95, 97–100, 104f
Diptera, 152–159 Chaeotosalpinx tapanilai, 337
Mecopteroidea, 147 Chironomidae, 152
Siphonaptera, 139, 147–152 Chrysidoidea, 165, 168–169
Aphelenchoides marinus, 83 Ciliates, 246
Archicnephia ornithoraptor, 157 Cladocyclus gardneri, 253–258
Ascaris lumbricoides, 362–363, 369–371 Clonorchis sinensis, 361–362
Ascarites priscus, 58, 59f Cobelodus aculeatus, 105
Asterolepis ornata, 97–99 Coleoptera
Astya crassus, 249 Adiphlebia lacoana, 161

421 j
422 Index

Coleoptera (Continued ) Cymothooidea, 241, 267–269


fossil representatives, 162 Cyrenodonax formosana, 210
general aspects, 161–162 Cystomyzon dimerum, 252–253
phylogenetic inference, 162
Computed tomography (CT), 5–6, 25–26, D
32, 35 Dakoticancer overanus, 244–246
Copepods, 252–259 DCA. See Detrended correspondence
in echinoderms, 252–253, 254t–257t analysis (DCA)
in fish, 253–259 Dendrogaster, 249
Coprolites, 22–26, 55, 103–105, 293, Detrended correspondence analysis
367–368, 391–392, 394–395 (DCA), 220
Coral–crinoid association Devonian, 97–100, 140, 176, 209–210,
description, 299 241, 258–259, 265–266, 336–338
material/locality and horizon, 299 Devonian fossil hook circlets, 98f–99f
reliability, 301 Acanthocephala, 99–100
Corethrellidae, 152 in A. ornata, 97–99
Corkcrew shaped symbionts, 334–335 Cestoda, 99–100
Cosmocercoides dukae, 57 in L. gaujicus, 97–99
Cosmocercoids, 81 Monogenea, 99–100
Cretaceous, 9t–13t, 54, 80–81, 105–106, Diabolepis, 116–117
146, 151, 162–163, 165, 167–168, Dickinsonia, 29–30, 108–109
171–173, 178–179, 204, 237–239, Diphyllobothrium sp., 373–374
238f, 241–242, 251t, 313–319, D. pacificum, 108, 359
331, 340, 375, 377 Diptera
Cretachododes burmitis, 175 fossil representatives, 154–159, 176, 182
Cretaciaphelenchoides burmensis, 57 general aspects, 152–153
Cretacimermis protus, 58, 59f phylogenetic inference, 153–154
Cretoscelis burmitica, 143, 144f Dominican amber nematodes, 16t, 65–75,
Crinoid columnals, 296 67f–75f, 109–110, 157, 160, 179,
Crustaceans 375
cysts/swellings, 234–235 Bursaphelenchus similus, 70
detrimental effects, 234–235 Cascofilaria dominicana, 73
intermediate host, 235 Cephalotes, 67
lineages, 235 Chalcidonema paradoxa, 74
parasites Heydenius
equivocal fossil evidence, 246 H. dipterophilus, 69
fossil evidence, 236–246 H. lamprophilus, 73
modern evidence, 246–248 H. myrmecophilia, 65
noncrustacean. See Noncrustacean H. neotropicus, 72–73
hosts H. saprophilus, 73
stratigraphic ranges, 273, 275f H. scatophilus, 68–69
surface scars, 234–235 Myrmeconema
whale barnacles, 234–235 M. antiqua, 65–67
Cryptochirus coralliodytes, 259–260 M. neotropicum, 65–67
Culicidae, 152, 156 Oligaphelenchoides maximus, 70–71
Curvolithus, 108–109 Palaeoanguina dominicana, 71
Cylindrical shaped symbionts, 334–335 Palaeoparasitylenchus dominicana, 67–68
Index 423

Paleoiotonchium dominicanum, 69–70 E. histolytica, 357, 392–393, 407–408


Setonema protera, 71–72 E. invadens, 351–352
Enterobius vermicularis, 352–353
E Enzyme-linked immunosorbent assay
Echinococcus granulosus, 362, 397–398, (ELISA), 392–393
399t–400t, 404–405 Eocene, 9t–13t, 60, 101–103, 102f,
Echinoderms, 252–253, 254t–257t 106–107, 156, 160–161, 167, 170,
borings, 292–293 205, 207f, 213t–215t, 260–261,
confidence limits, 294–298 269, 340
coral–crinoid association, 299–301, 300f Epizoobionts infesting
description, 299 description, 305
material/locality and horizon, 299 material/locality and horizon, 304, 304f
reliability, 301 reliability, 305–306
epizoobionts infesting
description, 305 F
material/locality and horizon, 304, Facetotecta, 271
304f Flatworms. See Parasitic flatworms, free-
reliability, 305–306 living flatworms; Platyhelminthes
growth deformity Fossil colonial animals
description, 302, 302f colonial animals, 331–333
material/locality and horizon, 302 parasites, 341–342
reliability, 303 putative parasites
Synbathocrinus, 303 borings, 340–341
interpretation, 294–298, 295f galls, 340
Linnean classification, 293 graptolites, supposed parasites of, 341
palaeoparasitology, 294 recognition, 333–334
pits site selectivity symbiotic intergrowths and
description, 314–316, 315f bioclaustrations. See Symbiotic
material/locality and horizon, intergrowths/bioclaustrations
313–314, 314f Free-living flatworms, 108–110
reliability, 319
platyceratid gastropods infesting, G
306–313, 307f–308f Gaimardia trapesina, 206–207
coprophagic/parasitic relationship, Gall crabs, 259–261
309–310 Galls, 340
description, 306–309 Genomic fossils, 3–4
Lacrimichnus, 311 Giardia
material/locality and horizon, 306 G. duodenalis, 392–393
Neoplatycrinus, 310, 313 G. intestinalis, 357, 363
reliability, 313 Graptolites, supposed parasites of, 341
putative parasitic infestations, 298 Growth deformity, 292–294
trace fossils, 293 description, 302, 302f
Ectoparasites, 139, 142, 145, material/locality and horizon, 302
147–148, 158–159, 181, reliability, 303
391–392 Synbathocrinus, 303
Endosacculus moltkiae, 249 Gymnophallidae, 101–103, 206–207, 208t
Entamoeba Gymnophalloides seoi, 361–362
424 Index

H Insect parasitism s.str.


Haemolymph, 236–237 Antliophora
Halloween pumpkin-mask, 252–253, 253f Diptera, 152–159
Hamarilipora minima, 342 Mecopteroidea, 147
Hapalocarcinus marsupialis, 259–260 Siphonaptera, 148–152
Helicosalpinx, 337 Neuropteroida, 159
Hemiptera Coleopterida, 161–162
fossil representatives, 146–147 Neuroptera, 159–161
general aspects, 146 Paraneoptera
phylogenetic inference, 146 Hemiptera, 146–147
Heydenius Phthiraptera, 142–146
H. araneus, 65 Insects
H. brownii, 61–62 in fossil record, 140–141
H. cecidomyae, 62–63 as hosts
H. matutinus, 60, 61f mites, 175–176
H. sciarophilus, 62 Nematoida, 174–175
H. tabanae, 77 pseudoscorpions, 177–179
Hoekia monticulariae, 266–267 parasites and hosts, 139–140
Holocene, 9t–13t, 77–79, 106–108, 207f, as vectors
210, 212f fossil representatives, 182
Horizontal gene transfer (HGT), 34–35 general aspects, 179–181
Howardula helenoschini, 63–64 phylogenetic inference, 181
Human parasites Interpolating/extrapolating extant
archaeological time, intestinal parasite parasite
eggs survival over, 392–393 conodonts, 110–112
medieval period, 397–412 extant and fossil host ranges, 110–112,
detecting migrations, 406–408 111f
Europe, ectoparasites in, 408–410 pentastomids, 110–112
Europe prior to, 395–397 Irregular calcareous deposits (ICD), 206,
farming animals, 404–406 210
fishing industry, 403–404, 403f Ischnocera, 143
parasites, medical history of, 410–412 Isopod parasites
sanitation, 398–403, 400f–401f in decapod crustaceans
zoonotic parasites, reduction in, 408 age, 241
past parasites prevalence, 393–395 biogeography, 241–242
sources of evidence, 391–392 fossil record, 237–239
Hydnophora, 266–267 ichnotaxon, erecting, 242–244
Hymenoptera infestation patterns, 239–240, 240f
fossil representatives, 167–169 life cycle and parasitism, 236–237
general aspects, 163–166 modern evidence, 236
phylogenetic inference, 166–167 quantitative data, 239
Hyperparasitoidism, 139–140 in fishes and squids, 267–269

I J
Ichneumonoidea, 165 Jurassic, 9t–13t, 80, 148, 151–152, 156,
Ichnology, 293 158, 160–161, 173, 212f, 237–241,
Insect-parasitic tylenchs, 81 238f, 253f, 254t–257t, 313
Index 425

K Montemagrechirus, 260–261
Kabatarina, 253–258, 258f Myxozoa, 27
K. patersoni, 9t–13t, 253–258
Kanthyloma crusta, 237–239, 238f N
Kato-Katz technique, 363–364 Nematode parasites
Kerodon rupestris, 77 from cenozoic
baltic amber, 60–65, 61f–66f
L fossil nematodes
Latrines, 394 amber, 54
Light acetolysis, 365–366 coprolites, 55
Liposcelidae, 143 rock fossils, 54–55
Lodeacanthus gaujicus, 97–99, 246 invertebrates, 79–81
Lutz spontaneous sedimentation method, mesozoic, body fossils from, 55–60,
366–367 56f–58f
Oligocene–Miocene
M Dominican amber nematodes, 65–75,
Maculo, 371 67f–75f
Malacostraca, 272 Mexican amber nematodes, 75–77, 76f
Mallophaga, 142 Palaeozoic parasitic nematodes, 55
Marine crustaceans, 9t–13t, 235, 248 plants, 83
Mass spectrometry, 3 Pleistocene and Holocene, 77–79
Megalosaurus dunkeri, 105–106 humans, 78–79
Megalyroidea, 167–168 from Pliocene, 77
Megamenopon rasnitsyni, 145–146 vertebrates, 81–82
Metagonimus yokogawai, 361–362 Nematoida
Microendoliths, 341 fossil representatives, 175
Microniscus larva, 236–237 general aspects, 174
MicroRNAs, 94–95 phylogenetic inference, 174
Mineralized coprolites, 394–395 Neolithic Revolution, 360–361
Miocene, 9t–13t, 108–109, 115, 149, 158, Neuroptera, 159
162, 205, 237–239, 339, 375 fossil representatives, 160–161
Mites general aspects, 159–160
fossil representatives, 176 phylogenetic inference, 160
general aspects, 175 Neuropteroida, 159
phylogenetic inference, 176 Coleopterida, 161–162
Molecular clock methods Neuroptera, 159–161
biogeographic calibrations, 116 Nilonema, 32–34
cophylogenetic approach, 113 Nothrotheriops shastensis, 78
fossil calibrations, criteria, 113–114 Nucleotide sequencing, 369
Polystomatidae, 116–117 Nuculodonta gotlandica, 211–216
pseudo-congruence, 116
Schistosoma, 115 O
taxon sampling, 117–119 Oligaphelenchoides atrebora, 75–76, 76f
Westollrhynchus, 116–117 Oligocene, 156, 160–161, 266–267,
Moltkia minuta, 249 339–340
Monogenea, 94–95, 119–120 Ordovician, 29–30, 81–82, 110–112, 114,
Monopisthocotylea, 94–95 262–264, 296, 333–334, 338
426 Index

Ordovician crinoids, 296 Paleogene, 101–103, 160, 168, 212f


Orsten preservation, 14–15 Paleothelastoma tipulae, 57, 57f
Orussidae, 163 Paraneoptera
Ostracods, 246 Hemiptera
fossil representatives, 146–147
P general aspects, 146
Palaega phylogenetic inference, 146
P. lamnae, 267–268 Phthiraptera
P. nusplingensis, 267–268, 268f fossil representatives, 145–146
Palaeoallantonema baltica, 64 general aspects, 142–143
Palaeoaphelenchoides balticus, 60 phylogenetic inference, 143–145,
Palaeocosmocerca burmanicum, 57 144f
Palaeoparasitology Psocodea/Psocoptera, 142
Ascaris lumbricoides, 354 Parasite fossil record
Clonorchis sinensis, 361–362 body fossils, 16t
coprolites, 354 Eocene mosquito, 15
defined, 353–363, 354f–355f galls, 17
ectoparasites, 376–377 Orsten preservation, 14–15
humans, 350 Pleistocene, 14
light microscopy techniques, 363–365, Rhynie Chert preservation, 14–15
364f coprolites, 22–26, 23t–24t
microscope, counting remains under, helminths, 8
365–368 host–parasite associations, 8–14, 9t–13t
analysis of sediments, 365–367 trace fossils and pathologies
mummies, coprolites in, 367–368 gastropods, 19
molecular techniques gymnophallid-induced pits, 18–19
molecular diagnosis, 368–369 igloo-shaped concretions, 18–19
Oswaldo Cruz Foundation, 354–356 pearls, 101–103, 208t, 209–210
parasite finds, 369–374 unicellular eukaryotes, 20–21
Ascaris lumbricoides, 369–371 Parasite phylogeny/evolution
Diphyllobothrium sp., 373–374 acanthocephalans, 28
Enterobius vermicularis, 372–373 aDNA techniques, 29
hookworms, 371–372 Chromera velia, 27
parasitism, 350 contamination, 27
prehistoric parasitic infections and human HGT and parasitic DNA, 34–35
coprolites, animals in, 374 molecular clocks
Salmonella enterica, 357 Cambrian–Ordovician, 29–30
Schistosoma haematobium, 353 computed tomography, 32
Trichuris trichiura, 354 Dickinsonia, 29–30
trypanosomatids and paradigm shift, Nilonema, 32–34
375–376 Spinther, 29–30
Palaeoproteomics, 7 Myxozoa, 27
Palaeotetradonema Parasitic flatworms, free-living flatworms,
P. phlebotomae, 76–77 94
P. sciarae, 62 Cretaceous egg, 105–106
Palaeozoic crinoids, 297, 306–313, Devonian fossil hook circlets, 98f–99f
321 Acanthocephala, 99–100
Index 427

A. ornata, 97–99 Pionodesmotes domhainfharraigeanus,


Cestoda, 99–100 252–253
L. gaujicus, 97–99 Plantago, 4
Monogenea, 99–100 Plant parasitism
Eocene shell pits, 106–107 nematode parasites, 83
extant and fossil host ranges, 110–112, vs. phytophagy
111f fossil representatives, 173–174
free-living flatworms, 108–110 general aspects, 172
Holocene evidence, 107–108 phylogenetic inference, 173
Permo-Carboniferous egg, 103–105, Platyceratid gastropods infesting, 306–313,
104f 307f–308f
Pleistocene mammal coprolite, 107 coprophagic/parasitic relationship,
Silurian blister pearls and calcareous 309–310
concretions, 101–103, 102f description, 306–309
Parasitoids Lacrimichnus, 311
Hymenoptera, 163–169 material/locality and horizon, 306
Strepsiptera, 169–172 Neoplatycrinus, 310, 313
Partial mortality, 332 reliability, 313
Pathologies Platyhelminthes, 95–96, 246
blister pearls, 101–103, 207–210 Platyzoa, 95–96, 119–120
galls Pleistocene, 9t–13t, 77–79, 107, 216,
in colonial organisms, 340 260–261, 370
in echinoderms, 319 Pleistocene mammal coprolite, 107
growth deformities, 292–293, 302–303 Pliocene, 77, 204, 212f, 267–268, 335f,
Halloween pumpkin-mask, 252–253, 339–340
254t–257t Polyopisthocotylea, 94–95
igloo-shaped concretions, 18–19, Porotic hyperostosis, 358
101–103 Potamocorbula amurensis, 221–222
irregular calcareous deposits (ICD), 206, Precambrian, 29–30, 108–109, 174,
208t 219–220
pits Proctotrupomorpha, 164–165, 168
in bivalves, 106–107, 211–222 Proheterorhabditis burmanicus, 58, 58f
in echinoderms, 293 Pseudopulex jurassicus, 151
Pelagic colonies, 331–332 Pseudoscorpions
Pentastomida, 5–6, 9t–13t, 27, 110–112, fossil representatives, 179
111f, 116–117, 140, 235, 261–265, general aspects, 177–178
263t phylogenetic inference, 178
Permian, 105, 108–109, 114–115, 147, Psychodidae, 153
153–154, 241, 303, 306, 307f, 310 Putative parasites
Phase contrast tomography, 5–6 borings, 340–341
Phthipodochiton thraivensis, 295–296, 295f galls, 340
Phthiraptera graptolites, supposed parasites of, 341
fossil representatives, 145–146 recognition, 333–334
general aspects, 142–143 symbiotic intergrowths and
phylogenetic inference, 143–145, 144f bioclaustrations. See Symbiotic
Psocodea/Psocoptera, 142 intergrowths/bioclaustrations
Pinworm infection, 359 Pyrgomatidae, 266–267, 266f
428 Index

R T
Rhizocephalan barnacles, 244–246 Tantulocarida, 248, 270
Rhynchophthirina, 143 Tarbellastrea reussiana, 266–267
Rhynie Chert, 14–15, 55, 56f Tarwinia australis, 149–150
Torquaysalpinx, 336
S Trace fossils/pathologies, 18–22, 108–109,
Salmonella enterica, 357 242, 293, 311, 337
Saurodectes vrsanskyi, 145 blister pearls, 209–210
Saurophthirus borings
S. exquisitus, 151 in colonial organisms, 340–341
S. longipes, 149 in crustaceans, 340–341
Schistosoma, 115 in echinoderms, 292–293
S. haematobium, 108 gymnophallid-induced pits, 18–19
Schistosomiasis mansoni, 363–364 igloo-shaped concretions, 18–19
Sedilichnus unicellular eukaryotes, 20–21
S. excavatus, 316–317 ‘zombie’ ants, 20–21
S. paraboloides, 302, 305 Trapeziidae, 269
Shergoldana australiensis, 175 Trematodes
Silurian, 5–6, 57, 101–103, 140, 209–216, Abra tenuis, 209–210
241, 334–335, 335f bivalve molluscs, 203, 206–210
Siphonaptera, 9t–13t, 15–17, 377, blister pearls, 209–210
391–392, 398 detrimental effects of, 222–225
fossil representatives, 149–152 Anodonta piscinalis, 222–223
general aspects, 148 Chamalea gallina, 223–225, 225f
phylogenetic inference, 148–149 Transenella, 223–225
Smilium zancleanum, 248 fossil and subfossil bivalves, trematode-
Smithsoninema inaequale, 60 induced pits occurrences, 211–222
Spinther, 29–30 paleoenvironmental indicators,
Strashila incredibilis, 149 220–222, 221f
Strepsiptera taphonomy and origin, 217–220
fossil representatives, 171–172 taxonomic/temporal and ecological,
general aspects, 169–170 211–216, 212f, 213t–215t, 217f
phylogenetic inference, 170 Gymnophallus rebecqui, 209–210
Strongylus edentatus, 77 parasite–host interaction, 204–205
Symbiodinium, 342 Triaenodes balticus, 64–65
Symbiotic intergrowths/bioclaustrations, Triassic, 9t–13t, 101–103, 140, 145,
334–340, 334f 152–154, 156, 167, 209–210, 313
caunopores, 335, 335f Trichuris trichiura, 354, 369–371
Celleporaria, 338 Trigonalyoidea, 168
Chaetosalpinx and bioclaustrations, Trisodium phosphate aqueous solution,
336–338 363
cornulitids and stromatoporoids, 338 Troglocarcinus corallicola, 259–260, 259f
Culicia, 338 Trypanosoma
pyrgomatid barnacles, 339–340 T. antiquae, 375
rugose corals and stromatoporoids, 336 T. cruzi, 352, 375
Synbathocrinus conicus, 302, 302f Tumidocarcinus giganteus, 244–246
Syringonomous typicus, 60 Tyrannosaurus rex, 7, 296
Index 429

U X
Ulophysema, 249 Xenopsylla cheopis, 409–410

V Z
Venericor clarendonensis, 102f, 106, Zooids, 330
213t–215t, 216, 219–220
Vespoidea, 165–166
CONTENTS OF VOLUMES IN THIS SERIES

Volume 41 Volume 43
Drug Resistance in Malaria Parasites of Genetic Exchange in the Trypanosomatidae
Animals and Man W. Gibson and J. Stevens
W. Peters
The Host-Parasite Relationship in Neosporosis
Molecular Pathobiology and Antigenic A. Hemphill
Variation of Pneumocystis carinii
Y. Nakamura and M. Wada Proteases of Protozoan Parasites
P.J. Rosenthal
Ascariasis in China
P. Weidono, Z. Xianmin and Proteinases and Associated Genes of Parasitic
D.W.T. Crompton Helminths
J. Tort, P.J. Brindley, D. Knox,
The Generation and Expression of Immunity K.H. Wolfe, and J.P. Dalton
to Trichinella spiralis in Laboratory Rodents
R.G. Bell Parasitic Fungi and their Interaction with the
Insect Immune System
Population Biology of Parasitic Nematodes: A. Vilcinskas and P. G€otz
Application of Genetic Markers
T.J.C. Anderson, M.S. Blouin Volume 44
and R.M. Brech
Cell Biology of Leishmania
Schistosomiasis in Cattle B. Handman
J. De Bont and J. Vercruysse
Immunity and Vaccine Development in the
Volume 42 Bovine Theilerioses
N. Boulter and R. Hall
The Southern Cone Initiative Against Chagas
The Distribution of Schistosoma bovis Sonaino,
Disease
1876 in Relation to Intermediate Host
C. J. Schofield and J.C.P. Dias
Mollusc-Parasite Relationships
Phytomonas and Other Trypanosomatid H. Moné, G. Mouahid, and S. Morand
Parasites of Plants and Fruit
The Larvae of Monogenea (Platyhelminthes)
E.P. Camargo
H.D. Whittington, L.A. Chisholm, and
Paragonimiasis and the Genus Paragonimus K. Rohde
D. Blair, Z.-B. Xu, and T. Agatsuma
Sealice on Salmonids: Their Biology and
Immunology and Biochemistry of Hymenolepis Control
diminuta A.W. Pike and S.L. Wadsworth
J. Anreassen, E.M. Bennet-Jenkins, and
C. Bryant Volume 45
Control Strategies for Human Intestinal The Biology of some Intraerythrocytic
Nematode Infections Parasites of Fishes, Amphibia and Reptiles
M. Albonico, D.W.T. Cromption, and A.J. Davies and M.R.L. Johnston
L. Savioli
The Range and Biological Activity of FMR
DNA Vaocines: Technology and Applications Famide-related Peptides and Classical
as Anti-parasite and Anti-microbial Agents Neurotransmitters in Nematodes
J.B. Alarcon, G.W. Wainem and D. Brownlee, L. Holden-Dye,
D.P. McManus and R. Walker

431 j
432 Contents of Volumes in This Series

The Immunobiology of Gastrointestinal Forecasting Diseases Risk for Increased


Nematode Infections in Ruminants Epidemic Preparedness in Public Health
A. Balic, V.M. Bowles, and E.N.T. Meeusen M.F. Myers, D.J. Rogers, J. Cox, A. Flauhalt,
and S.I. Hay

Volume 46 Education, Outreach and the Future of


Remote Sensing in Human Health
Host-Parasite Interactions in Acanthocephala: B.L. Woods, L.R. Beck, B.M. Lobitz, and
A Morphological Approach M.R. Bobo
H. Taraschewski
Eicosanoids in Parasites and Parasitic Infections Volume 48
A. Daugschies and A. Joachim The Molecular Evolution of
Trypanosomatidae
Volume 47 J.R. Stevens, H.A. Noyes, C.J. Schofield, and
W. Gibson
An Overview of Remote Sensing and Geodesy
for Epidemiology and Public Health Transovarial Transmission in the Microsporidia
Application A.M. Dunn, R.S. Terry, and J.E. Smith
S.I. Hay Adhesive Secretions in the Platyhelminthes
Linking Remote Sensing, Land Cover and I.D. Whittington and B.W. Cribb
Disease The Use of Ultrasound in Schistosomiasis
P.J. Curran, P.M. Atkinson, G.M. Foody, and C.F.R. Hatz
E.J. Milton
Ascaris and Ascariasis
Spatial Statistics and Geographic Information D.W.T. Crompton
Systems in Epidemiology and Public
Health Volume 49
T.P. Robinson
Antigenic Variation in Trypanosomes:
Satellites, Space, Time and the African Enhanced Phenotypic Variation in a
Trypanosomiases Eukaryotic Parasite
D.J. Rogers H.D. Barry and R. McCulloch
Earth Observation, Geographic Information The Epidemiology and Control of Human
Systems and Plasmodium falciparum Malaria African Trypanosomiasis
in Sub-Saharan Africa J. Pépin and H.A. Méda
S.I. Hay, J. Omumbo, M. Craig, and
R.W. Snow Apoptosis and Parasitism: from the Parasite to
the Host Immune Response
Ticks and Tick-borne Disease Systems in Space G.A. DosReis and M.A. Barcinski
and from Space
S.E. Randolph Biology of Echinostomes Except Echinostoma
B. Fried
The Potential of Geographical Information
Systems (GIS) and Remote Sensing in the
Epidemiology and Control of Human
Volume 50
Helminth Infections The Malaria-Infected Red Blood Cell:
S. Brooker and E. Michael Structural and Functional Changes
B.M. Cooke, N. Mohandas, and R.L. Coppel
Advances in Satellite Remote Sensing
of Environmental Variables for Schistosomiasis in the Mekong Region:
Epidemiological Applications Epidemiology and Phytogeography
S.J. Goetz, S.D. Prince, and J. Small S.W. Attwood
Contents of Volumes in This Series 433

Molecular Aspects of Sexual Development Enzymes Involved in the Biogenesis of the


and Reproduction in Nematodes and Nematode Cuticle
Schistosomes A.P. Page and A.D. Winter
P.R. Boag, S.E. Newton, and R.B. Gasser
Diagnosis of Human Filariases (Except
Antiparasitic Properties of Medicinal Plants and Onchocerciasis)
Other Naturally Occurring Products M. Walther and R. Muller
S. Tagboto and S. Townson
Volume 54
Volume 51 Introduction d Phylogenies, Phylogenetics,
Aspects of Human Parasites in which Surgical Parasites and the Evolution of Parasitism
Intervention May Be Important D.T.J. Littlewood
D.A. Meyer and B. Fried
Cryptic Organelles in Parasitic Protists and Fungi
Electron-transfer Complexes in Ascaris B.A.P. Williams and P.J. Keeling
Mitochondria
K. Kita and S. Takamiya Phylogenetic Insights into the Evolution
of Parasitism in Hymenoptera
Cestode Parasites: Application of In Vivo and J.B. Whitfield
In Vitro Models for Studies of the
Host-Parasite Relationship Nematoda: Genes, Genomes and the
M. Siles-Lucas and A. Hemphill Evolution of Parasitism
M.L. Blaxter

Volume 52 Life Cycle Evolution in the Digenea: A New


Perspective from Phylogeny
The Ecology of Fish Parasites with Particular T.H. Cribb, R.A. Bray, P.D. Olson, and
Reference to Helminth Parasites and their D.T. J. Littlewood
Salmonid Fish Hosts in Welsh Rivers:
Progress in Malaria Research: The Case for
A Review of Some of the Central
Phylogenetics
Questions
S.M. Rich and F.J. Ayala
J.D. Thomas
Phylogenies, the Comparative Method and
Biology of the Schistosome Genus
Parasite Evolutionary Ecology
Trichobilharzia
S. Morand and R. Poulin
P. Horak, L. Kolarova, and C.M. Adema
Recent Results in Cophylogeny Mapping
The Consequences of Reducing Transmission
M.A. Charleston
of Plasmodium falciparum in Africa
R.W. Snow and K. Marsh Inference of Viral Evolutionary Rates from
Molecular Sequences
Cytokine-Mediated Host Responses during
A. Drummond, O.G. Pybus, and A. Rambaut
Schistosome Infections: Walking the Fine
Line Between Immunological Control Detecting Adaptive Molecular Evolution:
and Immunopathology Additional Tools for the Parasitologist
K.F. Hoffmann, T.A. Wynn, and J.O. Mclnerney, D.T.J. Littlewood, and
D.W. Dunne C.J. Creevey

Volume 53 Volume 55
Interactions between Tsetse and Contents of Volumes 28–52
Trypanosomes with Implications for the Cumulative Subject Indexes for Volumes
Control of Trypanosomiasis 28–52
S. Aksoy, W.C. Gibson, and M.J. Lehane Contributors to Volumes 28–52
434 Contents of Volumes in This Series

Volume 56 Variation in Giardia: Implications for


Taxonomy and Epidemiology
Glycoinositolphospholipid from Trypanosoma R.C.A. Thompson and P.T. Monis
cruzi: Structure, Biosynthesis and
Immunobiology Recent Advances in the Biology of Echinostoma
J.O. Previato, R. Wait, C. Jones, species in the “revolutum” Group
G.A. DosReis, A.R. Todeschini, N. Heise B. Fried and T.K. Graczyk
and L.M. Previata Human Hookworm Infection in the
Biodiversity and Evolution of the Myxozoa 21st Century
E.U. Canning and B. Okamura S. Brooker, J. Bethony, and P.J. Hotez

The Mitochondrial Genomics of Parasitic The Curious Life-Style of the Parasitic Stages
Nematodes of Socio-Economic of Gnathiid Isopods
Importance: Recent Progress, and N.J. Smit and A.J. Davies
Implications for Population Genetics
and Systematics
M. Hu, N.B. Chilton, and R.B. Gasser
Volume 59
Genes and Susceptibility to Leishmaniasis
The Cytoskeleton and Motility in
Emanuela Handman, Colleen Elso, and Simon
Apicomplexan Invasion
Foote
R.E. Fowler, G. Margos, and G.H. Mitchell
Cryptosporidium and Cryptosporidiosis
Volume 57 R.C.A. Thompson, M.E. Olson, G. Zhu,
S. Enomoto, Mitchell S. Abrahamsen and
Canine Leishmaniasis N.S. Hijjawi
J. Alvar, C. Ca~navate, R. Molina, J. Moreno,
and J. Nieto Ichthyophthirius multifiliis Fouquet and
Ichthyophthiriosis in Freshwater Teleosts
Sexual Biology of Schistosomes R.A. Matthews
H. Moné and J. Boissier
Biology of the Phylum Nematomorpha
Review of the Trematode Genus Ribeiroia B. Hanelt, F. Thomas, and A. Schmidt-Rhaesa
(Psilostomidae): Ecology, Life History,
and Pathogenesis with Special Emphasis
on the Amphibian Malformation Problem Volume 60
P.T.J. Johnson, D.R. Sutherland, J.M. Kinsella Sulfur-Containing Amino Acid Metabolism
and K.B. Lunde in Parasitic Protozoa
The Trichuris muris System: A Paradigm of Tomoyoshi Nozaki, Vahab Ali, and Masaharu
Resistance and Susceptibility to Intestinal Tokoro
Nematode Infection The Use and Implications of Ribosomal DNA
L.J. Cliffe and R.K. Grencis Sequencing for the Discrimination of
Scabies: New Future for a Neglected Disease Digenean Species
S.F. Walton, D.C. Holt, B.J. Currie, Matthew J. Nolan and Thomas H. Cribb
and D.J. Kemp Advances and Trends in the Molecular
Systematics of the Parasitic
Volume 58 Platyhelminthes
Peter D. Olson and Vasyl V. Tkach
Leishmania spp.: On the Interactions they
Establish with Antigen-Presenting Cells Wolbachia Bacterial Endosymbionts of Filarial
of their Mammalian Hosts Nematodes
J.-C. Antoine, E. Prina, N. Courret, and Mark J. Taylor, Claudio Bandi, and
T. Lang Achim Hoerauf
Contents of Volumes in This Series 435

The Biology of Avian Eimeria with an Emphasis Implementation of Human Schistosomiasis


on their Control by Vaccination Control: Challenges and Prospects
Martin W. Shirley, Adrian L. Smith, and Alan Fenwick, David Rollinson, and
Fiona M. Tomley Vaughan Southgate

Volume 62
Volume 61
Models for Vectors and Vector-Borne Diseases
Control of Human Parasitic Diseases: Context D.J. Rogers
and Overview
David H. Molyneux Global Environmental Data for Mapping
Infectious Disease Distribution
Malaria Chemotherapy S.I. Hay, A.J. Tatem, A.J. Graham,
Peter Winstanley and Stephen Ward S.J. Goetz, and D.J. Rogers
Insecticide-Treated Nets Issues of Scale and Uncertainty in the Global
Jenny Hill, Jo Lines, and Mark Rowland Remote Sensing of Disease
Control of Chagas Disease P.M. Atkinson and A.J. Graham
Yoichi Yamagata and Jun Nakagawa Determining Global Population Distribution:
Human African Trypanosomiasis: Methods, Applications and Data
Epidemiology and Control D.L. Balk, U. Deichmann, G. Yetman,
E.M. Févre, K. Picozzi, J. Jannin, F. Pozzi, S.I. Hay, and A. Nelson
S.C. Welburn and I. Maudlin Defining the Global Spatial Limits of Malaria
Chemotherapy in the Treatment and Control Transmission in 2005
of Leishmaniasis C.A. Guerra, R.W. Snow and
Jorge Alvar, Simon Croft, and Piero Olliaro S.I. Hay

Dracunculiasis (Guinea Worm Disease) The Global Distribution of Yellow Fever and
Eradication Dengue
Ernesto Ruiz-Tiben and Donald R. Hopkins D.J. Rogers, A.J. Wilson, S.I. Hay, and
A.J. Graham
Intervention for the Control of
Soil-Transmitted Helminthiasis in Global Epidemiology, Ecology and Control
the Community of Soil-Transmitted Helminth Infections
Marco Albonico, Antonio Montresor, S. Brooker, A.C.A. Clements and
D.W.T. Crompton, and Lorenzo D.A.P. Bundy
Savioli Tick-borne Disease Systems: Mapping
Control of Onchocerciasis Geographic and Phylogenetic Space
Boakye A. Boatin and S.E. Randolph and D.J. Rogers
Frank O. Richards, Jr. Global Transport Networks and Infectious
Lymphatic Filariasis: Treatment, Control and Disease Spread
Elimination A.J. Tatem, D.J. Rogers and S.I. Hay
Eric A. Ottesen Climate Change and Vector-Borne Diseases
Control of Cystic Echinococcosis/Hydatidosis: D.J. Rogers and S.E. Randolph
1863-2002
P.S. Craig and E. Larrieu Volume 63
Control of Taenia solium Cysticercosis/ Phylogenetic Analyses of Parasites in the New
Taeniosis Millennium
Arve Lee Willingham III and Dirk Engels David A. Morrison
436 Contents of Volumes in This Series

Targeting of Toxic Compounds to the T. Sicheritz-Ponten, P. G. Foster,


Trypanosome’s Interior J. Samuelson, C.J. Noël, R.P. Hirt,
Michael P. Barrett and Ian H. Gilbert T.M. Embley, C. A. Gilchrist,
B.J. Mann, U. Singh, J.P. Ackers,
Making Sense of the Schistosome Surface S. Bhattacharya, A. Bhattacharya,
Patrick J. Skelly and R. Alan Wilson A. Lohia, N. Guillén, M. Duchene,
Immunology and Pathology of Intestinal T. Nozaki, and N. Hall
Trematodes in Their Definitive Hosts Epidemiological Modelling for Monitoring
Rafael Toledo, José-Guillermo Esteban, and and Evaluation of Lymphatic Filariasis
Bernard Fried Control
Systematics and Epidemiology of Trichinella Edwin Michael, Mwele N. Malecela-Lazaro,
Edoardo Pozio and K. Darwin Murrell and James W. Kazura
The Role of Helminth Infections in
Volume 64 Carcinogenesis
David A. Mayer and Bernard Fried
Leishmania and the Leishmaniases: A Parasite
A Review of the Biology of the
Genetic Update and Advances in
Parasitic Copepod Lernaeocera
Taxonomy, Epidemiology and
branchialis (L., 1767)(Copepoda:
Pathogenicity in Humans
Pennellidae
Anne-Laure Ba~nuls, Mallorie Hide and
Adam J. Brooker, Andrew P. Shinn, and
Franck Prugnolle
James E. Bron
Human Waterborne Trematode and
Protozoan Infections Volume 66
Thaddeus K. Graczyk and Bernard Fried
Strain Theory of Malaria: The First 50 Years
The Biology of Gyrodctylid Monogeneans: F. Ellis McKenzie,* David L. Smith,
The “Russian-Doll Killers” Wendy P. O’Meara, and
T.A. Bakke, J. Cable, and P.D. Harris Eleanor M. Riley
Human Genetic Diversity and the Advances and Trends in the Molecular
Epidemiology of Parasitic and Other Systematics of Anisakid Nematodes, with
Transmissible Diseases Implications for their Evolutionary
Michel Tibayrenc Ecology and HostParasite
Co-evolutionary Processes
Simonetta Mattiucci and Giuseppe Nascetti
Volume 65
Atopic Disorders and Parasitic Infections
ABO Blood Group Phenotypes and Aditya Reddy and Bernard Fried
Plasmodium falciparum Malaria: Unlocking
a Pivotal Mechanism Heartworm Disease in Animals and Humans
María-Paz Loscertales, Stephen Owens, John W. McCall, Claudio Genchi, Laura
James O’Donnell, James Bunn, H. Kramer, Jorge Guerrero, and
Xavier Bosch-Capblanch, and Luigi Venco
Bernard J. Brabin
Structure and Content of the Entamoeba Volume 67
histolytica Genome Introduction
C.G. Clark, U.C.M. Alsmark, Irwin W. Sherman
M. Tazreiter, Y. Saito-Nakano, V. Ali,
S. Marion, C. Weber, C. Mukherjee, An Introduction to Malaria Parasites
I. Bruchhaus, E. Tannich, M. Leippe, Irwin. W. Sherman
Contents of Volumes in This Series 437

The Early Years Polyamines


Irwin W. Sherman Irwin W. Sherman
Show Me the Money New Permeability Pathways and Transport
Irwin W. Sherman Irwin W. Sherman
In Vivo and In Vitro Models Hemoglobinases
Irwin W. Sherman Irwin W. Sherman
Malaria Pigment Erythrocyte Surface Membrane Proteins
Irwin W. Sherman Irwin W. Sherman
Chloroquine and Hemozoin Trafficking
Irwin W. Sherman Irwin W. Sherman
Invasion of Erythrocytes Erythrocyte Membrane Lipids
Irwin W. Sherman Irwin W. Sherman
Vitamins and Anti-Oxidant Defenses
Irwin W. Sherman
Volume 68
HLA-Mediated Control of HIV and HIV
Shocks and Clocks
Adaptation to HLA
Irwin W. Sherman
Rebecca P. Payne, Philippa C. Matthews,
Transcriptomes, Proteomes and Data Julia G. Prado, and Philip J.R. Goulder
Mining
An Evolutionary Perspective on Parasitism as a
Irwin W. Sherman
Cause of Cancer
Mosquito Interactions Paul W. Ewald
Irwin W. Sherman
Invasion of the Body Snatchers: The Diversity
Isoenzymes and Evolution of Manipulative Strategies
Irwin W. Sherman in HostParasite Interactions
Thierry Lefévre, Shelley A. Adamo, David G.
The Road to the Plasmodium falciparum Biron, Dorothée Missé , David Hughes, and
Genome Frédéric Thomas
Irwin W. Sherman
Evolutionary Drivers of Parasite-Induced
Carbohydrate Metabolism Changes in Insect Life-History Traits:
Irwin W. Sherman From Theory to Underlying Mechanisms
Pyrimidines and the Mitochondrion Hilary Hurd
Irwin W. Sherman Ecological Immunology of a Tapeworms’
The Road to Atovaquone Interaction with its Two Consecutive
Irwin W. Sherman Hosts
Katrin Hammerschmidt and
The Ring Road to the Apicoplast Joachim Kurtz
Irwin W. Sherman
Tracking Transmission of the Zoonosis
Ribosomes and Ribosomal Ribonucleic Acid Toxoplasma gondii
Synthesis Judith E. Smith
Irwin W. Sherman
Parasites and Biological Invasions
De Novo Synthesis of Pyrimidines and Folates Alison M. Dunn
Irwin W. Sherman
Zoonoses in Wildlife: Integrating Ecology into
Salvage of Purines Management
Irwin W. Sherman Fiona Mathews
438 Contents of Volumes in This Series

Understanding the Interaction Between an Volume 70


Obligate Hyperparasitic Bacterium,
Pasteuria penetrans and its Obligate Ecology and Life History Evolution of
Plant-Parasitic Nematode Host, Frugivorous Drosophila Parasitoids
Meloidogyne spp. Frédéric Fleury, Patricia Gibert, Nicolas Ris, and
Keith G. Davies Roland Allemand

HostParasite Relations and Implications for Decision-Making Dynamics in Parasitoids of


Control Drosophil
Alan Fenwick Andra Thiel and Thomas S. Hoffmeister

OnchocercaSimulium Interactions and the Dynamic Use of Fruit Odours to Locate Host
Population and Evolutionary Biology of Larvae: Individual Learning, Physiological
Onchocerca volvulus State and Genetic Variability as Adaptive
María-Gloria Basa~nez, Thomas S. Churcher, Mechanisms
and María-Eugenia Grillet Laure Kaiser, Aude Couty, and
Raquel Perez-Maluf
Microsporidians as Evolution-Proof Agents of
Malaria Control? The Role of Melanization and Cytotoxic
Jacob C. Koella, Lena Lorenz, and By-Products in the Cellular Immune
Irka Bargielowski Responses of Drosophila Against Parasitic
Wasps
A. Nappi, M. Poirié, and Y. Carton
Virulence Factors and Strategies of Leptopilina
Volume 69 spp.: Selective Responses in Drosophila
Hosts
The Biology of the Caecal Trematode
Mark J. Lee, Marta E. Kalamarz,
Zygocotyle lunata
Indira Paddibhatla, Chiyedza Small,
Bernard Fried, Jane E. Huffman, Shamus Keeler,
Roma Rajwani, and Shubha Govind
and Robert C. Peoples
Variation of Leptopilina boulardi Success in
Fasciola, Lymnaeids and Human Fascioliasis,
Drosophila Hosts: What is Inside the Black
with a Global Overview on Disease
Box?
Transmission, Epidemiology,
A. Dubuffet, D. Colinet, C. Anselme,
Evolutionary Genetics, Molecular
S. Dupas, Y. Carton, and M. Poirié
Epidemiology and Control
Santiago Mas-Coma, María Adela Valero, and Immune Resistance of Drosophila Hosts Against
María Dolores Bargues Asobara Parasitoids: Cellular Aspects
Patrice Eslin, Genevieve Prévost, Sebastien
Recent Advances in the Biology of
Havard, and Géraldine Doury
Echinostomes
Rafael Toledo, José-Guillermo Esteban, and Components of Asobara Venoms and their
Bernard Fried Effects on Hosts
Sébastien J.M. Moreau, Sophie Vinchon, Anas
Peptidases of Trematodes
Cherqui, and Genevieve Prévost
Martin Kasný, Libor Mikes, Vladimír Hampl,
Jan Dvorak, Conor R. Caffrey, John P. Strategies of Avoidance of Host Immune
Dalton, and Petr Horak Defenses in Asobara Species
Geneviéve Prevost, Géraldine Doury, Alix D.N.
Potential Contribution of
Mabiala-Moundoungou, Anas Cherqui, and
Sero-Epidemiological Analysis for
Patrice Eslin
Monitoring Malaria Control and
Elimination: Historical and Current Evolution of Host Resistance and Parasitoid
Perspectives Counter-Resistance
Chris Drakeley and Jackie Cook Alex R. Kraaijeveld and H. Charles J. Godfray
Contents of Volumes in This Series 439

Local, Geographic and Phylogenetic Scales of Coordinating Research on Neglected Parasitic


Coevolution in DrosophilaParasitoid Diseases in Southeast Asia Through
Interactions Networking
S. Dupas, A. Dubuffet, Y. Carton, and Remi Olveda, Lydia Leonardo, Feng Zheng,
M. Poirié Banchob Sripa, Robert Bergquist, and
Xiao-Nong Zhou
DrosophilaParasitoid Communities as
Model Systems for HostWolbachia Neglected Diseases and Ethnic Minorities in
Interactions the Western Pacific Region: Exploring
Fabrice Vavre, Laurence Mouton, and Bart the Links
A. Pannebakker Alexander Schratz, Martha
Fernanda Pineda, Liberty G. Reforma,
A Virus-Shaping Reproductive Strategy in a Nicole M. Fox, Tuan Le Anh,
Drosophila Parasitoid L. Tommaso Cavalli-Sforza,
Julien Varaldi, Sabine Patot, Mackenzie K. Henderson,
Maxime Nardin, and Raymond Mendoza, J€urg Utzinger,
Sylvain Gandon John P. Ehrenberg, and
Ah Sian Tee
Volume 71 Controlling Schistosomiasis in Southeast Asia:
A Tale of Two Countries
Cryptosporidiosis in Southeast Robert Bergquist and Marcel Tanner
Asia: What’s out There?
Yvonne A.L. Lim, Aaron R. Jex, Schistosomiasis Japonica: Control and
Huw V. Smith, and Robin B. Gasser Research Needs
Xiao-Nong Zhou, Robert Bergquist,
Human Schistosomiasis in the Economic Lydia Leonardo, Guo-Jing Yang,
Community of West African States: Kun Yang, M. Sudomo, and
Epidemiology and Control Remigio Olveda
Hélené Moné, Moudachirou Ibikounlé,
Achille Massougbodji, Schistosoma mekongi in Cambodia and Lao
and Gabriel Mouahid People’s Democratic Republic
Sinuon Muth, Somphou Sayasone,
The Rise and Fall of Human Sophie Odermatt-Biays,
Oesophagostomiasis Samlane Phompida, Socheat Duong, and
A.M. Polderman, M. Eberhard, S. Baeta, Peter Odermatt
Robin B. Gasser, L. van Lieshout,
P. Magnussen, A. Olsen, N. Spannbrucker, Elimination of Lymphatic Filariasis in
J. Ziem, and J. Horton Southeast Asia
Mohammad Sudomo, Sombat
Chayabejara, Duong Socheat,
Volume 72 Leda Hernandez, Wei-Ping Wu, and
Robert Bergquist
Important Helminth Infections in Southeast
Asia: Diversity, Potential for Control and Combating Taenia solium Cysticercosis in
Prospects for Elimination Southeast Asia: An Opportunity for
J€urg Utzinger, Robert Bergquist, Remigio Improving Human Health and Livestock
Olveda, and Xiao-Nong Zhou Production Links
A. Lee Willingham III, Hai-Wei Wu,
Escalating the Global Fight Against James Conlan, and Fadjar Satrija
Neglected Tropical Diseases Through
Interventions in the Asia Pacific Echinococcosis with Particular Reference to
Region Southeast Asia
Peter J. Hotez and John P. Ehrenberg Donald P. McManus
440 Contents of Volumes in This Series

Food-Borne Trematodiases in Southeast Asia: Towards Improved Diagnosis of


Epidemiology, Pathology, Clinical Zoonotic Trematode Infections in
Manifestation and Control Southeast Asia
Banchob Sripa, Sasithorn Kaewkes, Pewpan M. Maria Vang Johansen, Paiboon
Intapan, Wanchai Maleewong, and Sithithaworn, Robert Bergquist, and
Paul J. Brindley J€urg Utzinger
Helminth Infections of the Central Nervous The Drugs We Have and the Drugs
System Occurring in Southeast Asia and We Need Against Major Helminth
the Far East Infections
Shan Lv, Yi Zhang, Peter Steinmann, Jennifer Keiser and J€urg Utzinger
Xiao-Nong Zhou, and J€urg Utzinger
Research and Development of
Less Common Parasitic Infections in Southeast Antischistosomal Drugs in the People’s
Asia that can Produce Outbreaks Republic of China: A 60-Year Review
Peter Odermatt, Shan Lv, and Somphou Sayasone Shu-Hua Xiao, Jennifer Keiser,
Ming-Gang Chen, Marcel Tanner,
and J€urg Utzinger
Volume 73
Control of Important Helminthic Infections:
Concepts in Research Capabilities Vaccine Development as Part of the
Strengthening: Positive Experiences of Solution
Network Approaches by TDR in the Robert Bergquist and Sara Lustigman
People’s Republic of China and
Eastern Asia Our Wormy World: Genomics, Proteomics
Xiao-Nong Zhou, Steven Wayling, and and Transcriptomics in East and Southeast
Robert Bergquist Asia
Jun Chuan, Zheng Feng,
Multiparasitism: A Neglected Reality on Paul J. Brindley, Donald P. McManus,
Global, Regional and Local Scale Zeguang Han, Peng Jianxin, and Wei Hu
Peter Steinmann, J€urg Utzinger,
Zun-Wei Du, and Xiao-Nong Zhou Advances in Metabolic Profiling of
Experimental Nematode and Trematode
Health Metrics for Helminthic Infections Infections
Charles H. King Yulan Wang, Jia V. Li, Jasmina Saric, Jennifer
Implementing a Geospatial Health Data Keiser, Junfang Wu, J€urg Utzinger, and
Infrastructure for Control of Asian Elaine Holmes
Schistosomiasis in the People’s Republic Studies on the Parasitology, Phylogeography
of China and the Philippines and the Evolution of HostParasite
John B. Malone, Guo-Jing Yang, Lydia Interactions for the Snail Intermediate
Leonardo, and Xiao-Nong Zhou Hosts of Medically Important Trematode
The Regional Network for Asian Genera in Southeast Asia
Schistosomiasis and Other Helminth Stephen W. Attwood
Zoonoses (RNAS+): Target Diseases
in Face of Climate Change
Guo-Jing Yang, J€urg Utzinger, Shan Lv,
Volume 74
Ying-Jun Qian, Shi-Zhu Li, Qiang Wang, The Many Roads to Parasitism: A Tale of
Robert Bergquist, Penelope Vounatsou, Convergence
Wei Li, Kun Yang, and Xiao-Nong Zhou Robert Poulin
Social Science Implications for Control of Malaria Distribution, Prevalence, Drug
Helminth Infections in Southeast Asia Resistance and Control in Indonesia
Lisa M. Vandemark, Tie-Wu Jia, and Iqbal R.F. Elyazar, Simon I. Hay, and
Xiao-Nong Zhou J. Kevin Baird
Contents of Volumes in This Series 441

Cytogenetics and Chromosomes of Advances in Imaging of Animal Models of


Tapeworms (Platyhelminthes, Cestoda) Chagas Disease
Marta Spakulova, Martina Orosova, and Linda A. Jelicks and Herbert B. Tanowitz
John S. Mackiewicz
The Genome and Its Implications
Soil-Transmitted Helminths of Humans in Santuza M. Teixeira, Najib M. El-Sayed,
Southeast AsiadTowards Integrated and Patrícia R. Araujo
Control
Aaron R. Jex, Yvonne A.L. Lim, Jeffrey Genetic Techniques in Trypanosoma cruzi
Bethony, Peter J. Hotez, Martin C. Taylor, Huan Huang, and
Neil D. Young, and Robin B. Gasser John M. Kelly

The Applications of Model-Based Geostatistics Nuclear Structure of Trypanosoma cruzi


in Helminth Epidemiology and Control Sergio Schenkman, Bruno dos Santos
Ricardo J. Soares Magalh~aes, Archie C.A. Pascoalino, and Sheila C. Nardelli
Clements, Anand P. Patil, Aspects of Trypanosoma cruzi Stage
Peter W. Gething, and Simon Brooker Differentiation
Samuel Goldenberg and Andrea
Volume 75 
Rodrigues Avila

Epidemiology of American Trypanosomiasis The Role of Acidocalcisomes in the Stress


(Chagas Disease) Response of Trypanosoma cruzi
Louis V. Kirchhoff Roberto Docampo, Veronica Jimenez,
Sharon King-Keller, Zhu-hong Li, and
Acute and Congenital Chagas Disease Silvia N.J. Moreno
Caryn Bern, Diana L. Martin, and
Robert H. Gilman Signal Transduction in Trypanosoma cruzi
Huan Huang
Cell-Based Therapy in Chagas Disease
Antonio C. Campos de Carvalho,
Adriana B. Carvalho, and Regina Volume 76
C.S. Goldenberg Bioactive Lipids in Trypanosoma cruzi Infection
Targeting Trypanosoma cruzi Sterol Fabiana S. Machado, Shankar Mukherjee,
14a-Demethylase (CYP51) Louis M. Weiss, Herbert B. Tanowitz, and
Galina I. Lepesheva, Fernando Villalta, Anthony W. Ashton
and Michael R. Waterman Mechanisms of Host Cell Invasion by
Experimental Chemotherapy and Approaches Trypanosoma cruzi
to Drug Discovery for Trypanosoma cruzi Kacey L. Caradonna and Barbara
Infection A. Burleigh
Frederick S. Buckner Gap Junctions and Chagas Disease
Vaccine Development Against Trypanosoma Daniel Adesse, Regina Coeli Goldenberg, Fabio
cruzi and Chagas Disease S. Fortes, Jasmin, Dumitru A. Iacobas, Sanda
Juan C. Vazquez-Chagoyan, Iacobas, Antonio Carlos Campos de
Shivali Gupta, and Nisha Jain Garg Carvalho, Maria de Narareth
Meirelles, Huan Huang, Milena B. Soares,
Genetic Epidemiology of Chagas Disease Herbert B. Tanowitz, Luciana Ribeiro
Sarah Williams-Blangero, John L. VandeBerg, Garzoni, and David C. Spray
John Blangero, and Rodrigo Corrêa-Oliveira
The Vasculature in Chagas Disease
Kissing Bugs. The Vectors of Chagas Cibele M. Prado, Linda A. Jelicks,
Lori Stevens, Patricia L. Dorn, Justin O. Schmidt, Louis M. Weiss, Stephen M. Factor,
John H. Klotz, David Lucero, and Herbert B. Tanowitz, and
Stephen A. Klotz Marcos A. Rossi
442 Contents of Volumes in This Series

Infection-Associated Vasculopathy in Assessment and Monitoring of Onchocerciasis


Experimental Chagas Disease: in Latin America
Pathogenic Roles of Endothelin and Mario A. Rodríguez-Pérez,
Kinin Pathways Thomas R. Unnasch, and
Julio Scharfstein and Daniele Andrade Olga Real-Najarro
Autoimmunity
Edecio Cunha-Neto, Priscila Camillo Volume 78
Teixeira, Luciana Gabriel Nogueira, Gene Silencing in Parasites: Current Status and
and Jorge Kalil Future Prospects
ROS Signalling of Inflammatory Cytokines Raul Manzano-Roman, Ana Oleaga,
During Trypanosoma cruzi Infection Ricardo Pérez-Sanchez, and
Shivali Gupta, Monisha Dhiman, Jian-jun Mar Siles-Lucas
Wen, and Nisha Jain Garg GiardiadFrom Genome to Proteome
Inflammation and Chagas Disease: Some R.C. Andrew Thompson and Paul Monis
Mechanisms and Relevance Malaria Ecotypes and Stratification
André Talvani and Mauro M. Teixeira Allan Schapira and Konstantina Boutsika
Neurodegeneration and Neuroregeneration in The Changing Limits and Incidence of Malaria
Chagas Disease in Africa: 19392009
Marina V. Chuenkova and Mercio Robert W. Snow, Punam Amratia,
PereiraPerrin Caroline W. Kabaria, Abdisalan M. Noor,
Adipose Tissue, Diabetes and Chagas Disease and Kevin Marsh
Herbert B. Tanowitz, Linda A. Jelicks,
Fabiana S. Machado, Lisia Esper, Volume 79
Xiaohua Qi, Mahalia S. Desruisseaux,
Streamson C. Chua, Philipp E. Scherer, Northern Host  Parasite Assemblages:
and Fnu Nagajyothi History and Biogeography on the
Borderlands of Episodic Climate and
Environmental Transition
Volume 77 Eric P. Hoberg, Kurt E. Galbreath,
Joseph A. Cook, Susan J. Kutz, and
Coinfection of Schistosoma (Trematoda) with Lydden Polley
Bacteria, Protozoa and Helminths
Amy Abruzzi and Bernard Fried Parasites in Ungulates of Arctic North America
and Greenland: A View of Contemporary
Trichomonas vaginalis Pathobiology: New Diversity, Ecology and Impact in a World
Insights from the Genome Sequence Under Change
Robert P. Hirt, Natalia de Miguel, Susan J. Kutz, Julie Ducrocq, Guilherme
Sirintra Nakjang, Daniele Dessi, G. Verocai, Bryanne M. Hoar, Doug D.
Yuk-Chien Liu, Nicia Diaz, Colwell, Kimberlee B. Beckmen,
Paola Rappelli, Alvaro Acosta-Serrano, Lydden Polley, Brett T. Elkin, and
Pier-Luigi Fiori, and Jeremy C. Mottram Eric P. Hoberg
Cryptic Parasite Revealed: Improved Prospects Neorickettsial Endosymbionts of the Digenea:
for Treatment and Control of Human Diversity, Transmission and Distribution
Cryptosporidiosis Through Advanced Jefferson A. Vaughan, Vasyl V. Tkach, and
Technologies Stephen E. Greiman
Aaron R. Jex, Huw V. Smith, Matthew
J. Nolan, Bronwyn E. Campbell, Priorities for the Elimination of Sleeping
Neil D. Young, Cinzia Cantacessi, and Sickness
Robin B. Gasser Susan C. Welburn and Ian Maudlin
Contents of Volumes in This Series 443

Scabies: Important Clinical Consequences Natural Acquisition of Immunity to


Explained by New Molecular Studies Plasmodium vivax: Epidemiological
Katja Fischer, Deborah Holt, Bart Currie, and Observations and Potential Targets
David Kemp Ivo Mueller, Mary R. Galinski, Takafumi
Tsuboi, Myriam Arevalo-Herrera,
Review: Surveillance of Chagas Disease William E. Collins, and
Ken Hashimoto and Kota Yoshioka Christopher L. King
Volume 80 G6PD Deficiency: Global Distribution,
Genetic Variants and Primaquine Therapy
The Global Public Health Significance of Rosalind E. Howes, Katherine E. Battle,
Plasmodium vivax Ari W. Satyagraha, J. Kevin Baird, and
Katherine E. Battle, Peter W. Gething, Simon I. Hay
Iqbal R.F. Elyazar,
Catherine L. Moyes, Maríanne E. Sinka, Genomics, Population Genetics and
Rosalind E. Howes, Carlos A. Guerra, Evolutionary History of Plasmodium vivax
Ric N. Price, J. Kevin Baird, and Jane M. Carlton, Aparup Das, and
Simon I. Hay Ananias A. Escalante
Relapse Malariotherapy  Insanity at the Service of
Nicholas J. White and Mallika Imwong Malariology
Georges Snounou and Jean-Louis Pérignon
Plasmodium vivax: Clinical Spectrum, Risk
Factors and Pathogenesis
Nicholas M. Anstey, Nicholas M. Douglas, Volume 82
Jeanne R. Poespoprodjo, and
Ric N. Price Recent Developments in Blastocystis Research
C. Graham Clark, Mark van der Giezen,
Diagnosis and Treatment of Plasmodium vivax Mohammed A. Alfellani, and
Malaria C. Rune Stensvold
J. Kevin Baird, Jason D. Maguire, and
Ric N. Price Tradition and Transition: Parasitic Zoonoses of
People and Animals in Alaska, Northern
Chemotherapeutic Strategies for Canada, and Greenland
Reducing Transmission of Plasmodium Emily J. Jenkins, Louisa J. Castrodale,
vivax Malaria Simone J.C. de Rosemond, Brent R. Dixon,
Nicholas M. Douglas, George K. John, Stacey A. Elmore, Karen M. Gesy,
Lorenz von Seidlein, Nicholas M. Anstey, Eric P. Hoberg, Lydden Polley,
and Ric N. Price Janna M. Schurer, Manon Simard, and
R.C. Andrew Thompson
Control and Elimination of Plasmodium vivax
G. Dennis Shanks The Malaria Transition on the Arabian
Peninsula: Progress toward a Malaria-Free
Volume 81 Region between 1960-2010
Robert W. Snow, Punam Amratia, Ghasem
Plasmodium vivax: Modern Strategies Zamani, Clara W. Mundia, Abdisalan M.
to Study a Persistent Parasite’s Noor, Ziad A. Memish, Mohammad H. Al
Life Cycle Zahrani, Adel AlJasari, Mahmoud Fikri, and
Mary R. Galinski, Esmeralda V.S. Meyer, and Hoda Atta
John W. Barnwell
Microsporidia and ‘The Art of Living
Red Blood Cell Polymorphism and Together’
Susceptibility to Plasmodium vivax Jir í Vavra and Julius Lukes
Peter A. Zimmerman, Marcelo U. Ferreira,
Rosalind E. Howes, and Odile Patterns and Processes in Parasite Co-Infection
Mercereau-Puijalon Mark E. Viney and Andrea L. Graham
444 Contents of Volumes in This Series

Volume 83 Volume 85
IronSulphur Clusters, Their Biosynthesis, Diversity and Ancestry of Flatworms Infecting
and Biological Functions in Protozoan Blood of Nontetrapod Craniates “Fishes”
Parasites Raphael Orélis-Ribeiro, Cova R. Arias,
Vahab Ali and Tomoyoshi Nozaki Kenneth M. Halanych,Thomas H. Cribb,
and Stephen A. Bullard
A Selective Review of Advances in Coccidiosis
Research Techniques for the Diagnosis of Fasciola
H. David Chapman, John R. Barta, Infections in Animals: Room for
Damer Blake, Arthur Gruber, Mark Jenkins, Improvement
Nicholas C. Smith, Xun Suo, and Cristian A. Alvarez Rojas, Aaron R. Jex,
Fiona M. Tomley Robin B. Gasser, and
Jean-Pierre Y. Scheerlinck
The Distribution and Bionomics of
Anopheles Malaria Vector Mosquitoes in Reevaluating the Evidence for Toxoplasma
Indonesia gondii-Induced Behavioural Changes in
Iqbal R.F. Elyazar, Marianne E. Sinka, Rodents
Peter W. Gething, Siti N. Tarmidzi, Amanda R. Worth, R.C. Andrew Thompson,
Asik Surya, Rita Kusriastuti, Winarno, and Alan J. Lymbery
J. Kevin Baird, Simon I. Hay, and
Michael J. Bangs Volume 86
Next-Generation Molecular-Diagnostic Tools Historical Patterns of Malaria Transmission in
for Gastrointestinal Nematodes of China
Livestock, with an Emphasis on Small Jian-Hai Yin, Shui-Sen Zhou, Zhi-Gui Xia,
Ruminants: A Turning Point? Ru-Bo Wang, Ying-Jun Qian, Wei-Zhong
Florian Roeber, Aaron R. Jex, and Yang, and Xiao-Nong Zhou
Robin B. Gasser
Feasibility and Roadmap Analysis for Malaria
Elimination in China
Volume 84 Xiao-Nong Zhou, Zhi-Gui Xia, Ru-Bo Wang,
Joint Infectious Causation of Human Ying-Jun Qian, Shui-Sen Zhou, J€urg
Cancers Utzinger, Marcel Tanner, Randall Kramer,
Paul W. Ewald and Holly A. Swain Ewald and Wei-Zhong Yang

Neurological and Ocular Fascioliasis in Lessons from Malaria Control to Elimination:


Humans Case Study in Hainan and Yunnan
Santiago Mas-Coma, Veronica H. Agramunt, Provinces
and María Adela Valero Zhi-Gui Xia, Li Zhang, Jun Feng, Mei Li,
Xin-Yu Feng, Lin-Hua Tang, Shan-Qing
Measuring Changes in Plasmodium falciparum Wang, Heng-Lin Yang, Qi Gao, Randall
Transmission: Precision, Accuracy and Kramer, Tambo Ernest, Peiling Yap, and
Costs of Metrics Xiao-Nong Zhou
Lucy S. Tusting, Teun Bousema, David
L. Smith, and Chris Drakeley Surveillance and Response to Drive the
National Malaria Elimination Programme
A Review of Molecular Approaches for Xin-Yu Feng, Zhi-Gui Xia, Sirenda Vong,
Investigating Patterns of Coevolution in Wei-Zhong Yang, and Shui-Sen Zhou
Marine HostParasite Relationships
G€otz Froeschke and Sophie von der Heyden Operational Research Needs Toward Malaria
Elimination in China
New Insights into Clonality and Panmixia in Shen-Bo Chen, Chuan Ju, Jun-Hu Chen,
Plasmodium and Toxoplasma Bin Zheng, Fang Huang, Ning Xiao, Xia
Michel Tibayrenc and Francisco J. Ayala Zhou, Tambo Ernest, and Xiao-Nong Zhou
Contents of Volumes in This Series 445

Approaches to the Evaluation of Malaria Andrea Bosman, Robert David Newman,


Elimination at County Level: Case Study Tambo Ernest, Michael O’leary, and
in the Yangtze River Delta Region Ning Xiao
Min Zhu, Wei Ruan, Sheng-Jun Fei,
Jian-Qiang Song, Yu Zhang, Xiao-Gang
Mou, Qi-Chao Pan, Ling-Ling Zhang, Volume 87
Xiao-Qin Guo, Jun-Hua Xu, Tian-Ming The Allee Effect and Elimination of Neglected
Chen, Bin Zhou, Peiling Yap, Li-Nong Tropical Diseases: A Mathematical
Yao, and Li Cai Modelling Study
Surveillance and Response Strategy in the Manoj Gambhir, Brajendra K. Singh, and
Malaria Post-elimination Stage: Case Edwin Michael
Study of Fujian Province Mathematical Modelling of Leprosy and Its
Fa-Zhu Yang, Peiling Yap, Shan-Ying Zhang, Control
Han-Guo Xie, Rong Ouyang, Yao-Ying David J. Blok, Sake J. de Vlas,
Lin, and Zhu-Yun Chen Egil A.J. Fischer, and Jan Hendrik Richardus
Preparation of Malaria Resurgence in China: Mathematical Models of Human African
Case Study of Vivax Malaria Trypanosomiasis Epidemiology
Re-emergence and Outbreak in Kat S. Rock, Chris M. Stone,
Huang-Huai Plain in 2006 Ian M. Hastings, Matt J. Keeling,
Hong-Wei Zhang, Ying Liu, Shao-Sen Zhang, Steve J. Torr, and Nakul Chitnis
Bian-Li Xu, Wei-Dong Li, Ji-Hai Tang,
Shui-Sen Zhou, and Fang Huang Ecology, Evolution and Control of Chagas
Disease: A Century of Neglected
Preparedness for Malaria Resurgence in Modelling and a Promising Future
China: Case Study on Imported Cases in Pierre Nouvellet, Zulma M. Cucunuba,
2000–2012 and Sébastien Gourbiere
Jun Feng, Zhi-Gui Xia, Sirenda Vong,
Wei-Zhong Yang, Shui-Sen Zhou, and Mathematical Inference on Helminth Egg
Ning Xiao Counts in Stool and Its Applications in
Mass Drug Administration Programmes
Preparation for Malaria Resurgence in China: to Control Soil-Transmitted
Approach in Risk Assessment and Rapid Helminthiasis in Public Health
Response Bruno Levecke, Roy M. Anderson,
Ying-Jun Qian, Li Zhang, Zhi-Gui Xia, Dirk Berkvens, Johannes Charlier,
Sirenda Vong, Wei-Zhong Yang, Brecht Devleesschauwer, Niko Speybroeck,
Duo-Quan Wang, and Ning Xiao Jozef Vercruysse, and Stefan Van Aelst
Transition from Control to Elimination: Modelling Lymphatic Filariasis Transmission
Impact of the 10-Year Global Fund and Control: Modelling Frameworks,
Project on Malaria Control and Lessons Learned and Future Directions
Elimination in China Wilma A. Stolk, Chris Stone, and Sake J. de Vlas
Ru-Bo Wang, Qing-Feng Zhang, Bin Zheng,
Zhi-Gui Xia, Shui-Sen Zhou, Lin-Hua Modelling the Effects of Mass Drug
Tang, Qi Gao, Li-Ying Wang, and Administration on the Molecular
Rong-Rong Wang Epidemiology of Schistosomes
Poppy H.L. Lamberton, Thomas Crellen,
China–Africa Cooperation Initiatives in James A. Cotton, and Joanne P. Webster
Malaria Control and Elimination
Zhi-Gui Xia, Ru-Bo Wang, Duo-Quan Economic and Financial Evaluation of
Wang, Jun Feng, Qi Zheng, Chang-Sheng Neglected Tropical Diseases
Deng, Salim Abdulla, Ya-Yi Guan, Wei Bruce Y. Lee, Sarah M. Bartsch, and
Ding, Jia-Wen Yao, Ying-Jun Qian, Katrin M. Gorham
446 Contents of Volumes in This Series

Volume 88 A Perspective on Cryptosporidium and Giardia,


with an Emphasis on Bovines and Recent
Recent Developments in Malaria Vaccinology Epidemiological Findings
Benedict R. Halbroth and Simon J. Draper Harshanie Abeywardena, Aaron R. Jex, and
Robin B. Gasser
PfEMP1 – A Parasite Protein Family of Key
Importance in Plasmodium falciparum
Malaria Immunity and Pathogenesis Volume 89
Lars Hviid and Anja TR. Jensen
Ecology of Free-Living Metacercariae
Prospects for Vector-Based Gene Silencing to (Trematoda)
Explore Immunobiological Features of Neil J. Morley
Schistosoma mansoni
Cross-Border Malaria: A Major Obstacle for
Jana Hagen, Jean-Pierre Y. Scheerlinck,
Malaria Elimination
Neil D. Young, Robin B. Gasser, and
Kinley Wangdi, Michelle L. Gatton, Gerard C.
Bernd H. Kalinna
Kelly, Archie CA. Clements
Chronobiology of Trematode Cercarial
Development of Malaria Transmission-
Emergence: from Data Recovery to
Blocking Vaccines: From Concept to
Epidemiological, Ecological and
Product
Evolutionary Implications
Yimin Wu, Robert E. Sinden, Thomas S.
André Théron
Churcher, Takafumi Tsuboi, Vidadi
Strongyloidiasis with Emphasis on Human Yusibov
Infections and Its Different Clinical Forms
Rafael Toledo, Carla Mu~noz-Antoli, and
José-Guillermo Esteban

You might also like