You are on page 1of 74

Chemical and Mechanical

Methods for Pipeline


Integrity

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 1 03/11/17 6:50 PM


BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 2 03/11/17 6:50 PM
Chemical and Mechanical
Methods for Pipeline
Integrity

Wayne W. Frenier, FNACE


Frenier Chemistry Consultants

Written in Collaboration with T. D. Williamson,


Incorporated

Foreword by Richard B. Williamson


Chairman (Ret.), T. D. Williamson,
Incorporated

Society of Petroleum Engineers

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 3 03/11/17 6:50 PM


© Copyright 2018 Society of Petroleum Engineers

All rights reserved. No portion of this book may be reproduced in any form or by any means,
including electronic storage and retrieval systems, except by explicit, prior written permission of
the publisher except for brief passages excerpted for review and critical purposes.

Printed in the United States of America.

Disclaimer

This book was prepared by members of the Society of Petroleum Engineers and their well-
qualified colleagues from material published in the recognized technical literature and from their
own individual experience and expertise. While the material presented is believed to be based on
sound technical knowledge, neither the Society of Petroleum Engineers nor any of the authors or
editors herein provide a warranty either expressed or implied in its application. Correspondingly,
the discussion of materials, methods, or techniques that may be covered by patents implies no
freedom to use such materials, methods, or techniques without permission through appropriate
licensing. Nothing described within this book should be construed to lessen the need to apply
sound engineering judgment nor to carefully apply accepted engineering practices in the design,
implementation, or application of the techniques described herein.

ISBN 978-1-61399-496-2

First Printing 2018

Society of Petroleum Engineers


222 Palisades Creek Drive
Richardson, TX 75080-2040 USA

http://www.spe.org/store
service@spe.org
1.972.952.9393

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 4 03/11/17 6:50 PM


Dedication
I dedicate this book to my family. My wife, Dolores, our children Andrew Frenier and Kathleen Turner
and our grandchildren are the inspiration for the work that I have accomplished.
– Wayne Frenier

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 5 03/11/17 6:50 PM


BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 6 03/11/17 6:50 PM
Foreword
Critical Needs
Pipelines are the safest, most reliable, and most cost-effective means of transporting hazardous
fluids. In comparison to other forms of transportation, their construction and operations have
substantially less initial and ongoing impact on the environment. With approximately 2.3 million
miles of hazardous substance pipelines operating in as many as 120 countries, these lines are some
of the most important infrastructures in the world. The number will continue to grow with popula-
tion growth, changing patterns of fuel usages for power generation, and industrial, commercial, and
residential heating and transportation.
In just the last 50 years, these pipeline systems have grown from approximately 0.5 million to
2.3 million miles of pipelines built and in operation for crude oil, refined liquids, and natural gas
and gas liquids in regions across the world. These more recently constructed pipelines reflect the
growing industrial and post-war generations’ requirements of the industrializing and industrial
economies. They also reflect the critical role that pipelines continue to play in scaling the ability
of economies to expand their energy consumption in a safe and cost-effective manner. However,
these newer pipelines—built with new and improved material and construction techniques—as well
as those constructed pre-1970 all require continuing efforts to assure that they remain safe from
external forces and that they are resistant to corrosion, for example, and to ensure that the pipelines
remain free of materials and conditions that would impede the effective transport of fluids.

A Short History of Petroleum Pipelines


The history of the pipeline industry is largely the story of industrialized societies’ growing use of
and reliance on hydrocarbon liquids and gas for energy and essential chemicals.
While it is not completely clear when the first iron-based pipelines were used, the development
of the first drilled oil/gas wells in Pennsylvania circa 1865 was a major driver. There is also a claim
that iron-based lines were used in Canada circa 1860 to transport crude oil. Gas pipelines made of
wood are claimed to have been used as early as 1830. A Canadian source also says that a 25-km iron
gas pipeline was constructed in 1856 (NRC 2016).
At this early point in time, the market for distilled crude oil (kerosene) and gas was for lighting.
However, in 1885 Robert Bunsen’s invention of what is now known as the Bunsen burner opened
vast new opportunities to use natural gas. Once effective and numerous pipelines began to be built
in the 20th century, the use of natural gas expanded to home heating and cooking, manufacturing
and processing plants, and utility plants to generate electricity. The development of the internal
combustion engine and motor vehicles in the early 20th century dramatically changed the markets
for distilled oil-based products.
An interesting note is that during the 19th century, there were significant competitions between
transportation of petroleum products in pipelines and transportation using horse-drawn wagons (oil
in whiskey barrels, for example) and then in rail cars. This competition greatly affected the price of
the delivered products. Similar types of competitions between pipelines and rail delivery of crude
oil continue today.

Technical, Regulatory, and Societal Challenges


Pipeline maintenance and integrity management will become more necessary in hostile climates and
remote locations such as the Arctic and deep water. Additional challenges will include producing
additional appropriate devices for “unpiggable” lines. Chemical issues will include formulating,

vii

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 7 03/11/17 6:50 PM


viii Foreword

sourcing, and injecting “greener” corrosion inhibitors and other treatment chemicals into longer line
sections. Waste disposal of water and treatment of chemicals, as well as ecological requirements
and problems with heavy and waxy crude oils and high-water-cut wells, will require new solutions.
We are living in an age of almost instant communication of news about transportation failures.
Societies around the world will require that hazardous liquids and gases be transported safely and
reliably. The result will be more laws and regulations and the development of best practices to
reduce these failures. The growing experience in the pipeline industry as well as the current and
future infrastructure improvements can be used to continue to reduce spills and failures.

The Importance of Training


The various parts of the pipeline systems are physically connected from the wells to the consumers.
They are also connected by a relatively small set of mechanical and chemical principles. Once these
principles are understood and taught to pipeline engineers and operators, consistent and improved
practices can be developed.
To this end, this book—inspired, authored, compiled, and edited by Wayne Frenier, one of the
true experts in our industry—is a guide to the practicing pipeline engineer and others who serve and
support the pipeline industry.
Richard B. Williamson, Chairman (Ret.), T. D. Williamson, Incorporated

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 8 03/11/17 6:50 PM


Preface
Pipelines are a major part of the circulatory and transportation systems of the oil/gas and petrochem-
ical industries. Fluids are produced from the wells and flow through gathering lines or/and flowlines
to central points where some degree of processing and separation may take place (in onshore and
offshore facilities). The liquids and gases are then placed into pipelines for conveyance to refineries,
to chemical plants, or to power producers. The refined products also may pass through pipelines to
the ultimate users. At various points wastewater streams (including vast amounts of produced and
hydraulic fracture water) are generated, and these very diverse fluids may pass through pipelines
before disposal. This frequently means reinjection of a fraction of the fluids into the earth through
the injection-well system of a field or into licensed wastewater wells.
The author and technical consultants of this book (see “Acknowledgments”) understand that some
of the transportation of well fluids as well as the finished products is accomplished using tanker
trucks, barges, trains, and ships of many designs. At the date of the compilation of this book, there
was a shortage of pipelines to transport liquids or gas. The cost of rail transport was USD 10–15
more for a barrel of crude oil than by pipeline (Batheja 2014). The book, however concentrates on
the piping-conveyed segments of the transportation system.
Although the analogy of a circulatory system breaks down to some extent because all the fluids
are not returned to the producing formations, very large volumes of liquids and gases may be rein-
jected as part of pressure maintenance, enhanced oil recovery, stimulation activities (see Frenier and
Ziauddin 2014)., and waste disposal. The “heart” of the processes includes the natural pressure of
the producing formations as well as innumerable pumps and gas compressors of all descriptions that
maintain the flowing pressure of the systems.
The pipelines are vital and integral parts of the oil/gas systems of the world. Because of recent
advances in production of gas and oil from shale formations (Boyer et al. 2011; PI 2012), the future
growth in pipeline infrastructure will be significant. INGAA (2011) has predicted the growth for
North America through 2035. The values in the following predictions are from this reference.

• Natural gas transmission infrastructure


 43 Bcf/D of new natural gas transmission capability

 400,000 miles of gathering lines (at 16,000 miles/yr)

 1,400 miles/yr of new gas transmission mainline

 600 miles/yr of new laterals to and from natural-gas-fired power plants, processing facili-

ties, and storage fields


 24 Bcf/yr of new working gas capacity in storage

 197,000 hp/yr for pipeline compression

• Natural gas liquids and oil infrastructure


 USD 0.6 billion/year or a total of USD 14.5 billion over the study period for natural gas

liquids pipeline expenditures


 USD 1.3 billion/year or USD 31.4 billion over the study period of capital expenditures for

oil pipelines

Fig. P-1 is a photograph of a pipeline under construction showing some of the welded segments.
Fig. P-1a shows sections in the construction ditch with welded and unwelded sections. Fig. P-1b
demonstrates the large equipment needed for laying pipeline segments.

ix

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 9 03/11/17 6:50 PM


x Preface

(a) Pipeline showing sections

(b) Equipment laying sections

Fig. P-1—Pipelines under construction: (a) individual sections, (b) handling equipment.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 10 03/11/17 6:50 PM


Preface xi

An important driver for current construction of additional pipelines is the new light oil as well as
the loss of associated natural gas produced in several of the shale oil fields (Bakken in North Dakota,
USA, and the Eagle Ford in Texas, USA). Because of the lack of pipelines and gas-treating facilities,
as much as 30% of the gas is flared (Fig. P-2) or is used for directly powering hydraulically operated
equipment that then vents the gas into the atmosphere (Rahim 2013). A report by Aleklett (2013)
includes satellite photos claimed to have been taken by the National Aeronautics and Space Admin-
istration of multiple oilfield gas flares in the Bakken and Eagle Ford plays. The illumination from
the flares seems to compare in intensity with the illumination of major cities in the regions near the
flares. The state of North Dakota as well as a major producer (described in Rahim 2013) are claimed
to be committed to dramatically reducing the waste of these hydrocarbon streams by constructing
adequate transportation and other ways to use the products in the near future.
This new oil and gas also affects the location of any new pipelines. Speakers (Solomon et al. 2015;
Banerjee 2015) at the 2015 Pipeline + Energy Conference and Exposition in Tulsa, Oklahoma, USA,
noted that oil quality (light vs. heavy) also plays a role in the geographic direction of new lines as
well as the mode of transportation. They noted that the refineries best suited for heavy oil are on the
US coast on the Gulf of Mexico and that those designed for light oil are mostly on the US East Coast.
Thus, in some cases, rail transportation is currently needed, even if it is costlier than in-place pipelines.
In addition, methane is a potent “greenhouse gas.” USEPA (2013a) noted that the lifetime of
methane (CH4) in the atmosphere is much shorter than that of carbon dioxide (CO2), but CH4 is more
efficient at trapping radiation than is CO2. The report claims that pound for pound, the comparative
impact of CH4 on climate change is over 20 times greater than that of CO2 over a 100-year period.
Johnson (2014) has published a short review of the role that oil/gas production plays in contributing
to the release of methane to the atmosphere and the actual role it has in global climate change. He
cites information that up to 29% of the annual methane loss to the atmosphere comes from oil/gas

Fig. P-2—Orvis State natural gas flare, Arnegard North Dakota (Wikimedia 2013).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 11 03/11/17 6:50 PM


xii Preface

production and transportation. He claims that because of the trapping-of-heat factor of methane vs.
CO2, reducing methane emissions in the short term (i.e., in the next 10 years) should be a global
priority.
The needed construction of new gas-handling facilities is in turn driven by the economic value of
these resources as well as US Environmental Protection Agency (EPA) regulations (EPA 2012), 40
CFR 98 Subpart W (USEPA 2012), and 40 CFR 60 Subpart OOOO (USEPA 2013b). Details regard-
ing the difficulties of maintaining these current types of oil/gas facilities and new lines are described
in Sections 2.2.3 and 2.3.1.
The overall result for some sectors is “Midstream Mania.” Walton (2013) claims that companies
that construct, own, or maintain pipelines, tankage, and midstream processing plants are working
overtime to meet demand. One section of the article is titled “The Golden Age of Pipelines.” How-
ever, the author and consultants for this book note that significant problems have already arisen that
must be addressed and thus are the major subject of this book. Note that economic conditions as well
as market forces also will affect demand for new pipelines.

Goal and Organization of This Book


This current publication provides an overview of the science and technology of the use of a wide
range of pipeline chemicals and mechanical equipment for a general technical audience, with
emphasis placed on the basic chemical/physical principles by which the chemicals and devices can
enhance or maintain product delivery. Some knowledge of chemistry (equivalent to an introductory
college general chemistry course) is assumed. More-advanced concepts are introduced in Chapter 1.
The introductory chapter describes the varied pipeline environments, problems that require chem-
ical or mechanical intervention, and thus the need for the thousands of different chemicals and
devices that are in use. This chapter also reviews the important chemical/physical principles that are
common to most if not all the enhancement treatments. The applications of interventions described
are primarily in the upstream and midstream oil/gas business, but many of the methods can also be
used in refineries or product pipelines. This book is limited to internal pipeline flow assurance and
reliability issues and does not specifically address external corrosion, internal coatings, or cathodic
protection. See Goldschmidt and Streitberger (2003) for references to pipeline coatings and Lazzari
and Pedeferri (2006) for cathodic protection information, as well as NACE (2011) for a general
review of pipeline corrosion control.
The following is an outline for analysis of potential internal problems in pipelines and facilities:

1. Is there a problem that requires an intervention?


2. If there is a problem, how bad will it be?
3. Can the problem be managed through engineering and/or chemical means?
4. Evaluate the results of an intervention or control strategy.

Each major section and most subsections will include reviews of current literature as well as sum-
maries of the consensus understandings from the literature cited.
Chapter 1, “Introduction to the Technology of Flow and Integrity Management,” provides an over-
view of the reasons pipelines require intervention to enhance or maintain product delivery and intro-
duces the various types of chemical and mechanical interventions in use. This chapter also reviews
basic chemical and engineering processes that occur in pipeline operations. It also emphasizes the
commonality shared among many of the chemical and engineering processes across the pipeline
systems as well as the well production processes.
Chapter 2, “From the Well to the Consumer,” describes how the aqueous fluids, hydrocarbon
liquids, and gases change from the wellhead, through the gathering lines and surface or subsur-
face facilities, then through transmission (trunk) pipelines to a consumer. It outlines the chemical/

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 12 03/11/17 6:50 PM


Preface xiii

physical forces that ultimately affect the delivery of the products as well as the plans to anticipate
and alleviate problems.
Chapter 3, “Corrosion Processes in Pipelines and Facilities,” provides a review of internal corro-
sion and corrosion mechanisms that affect pipeline/facility operations. Here also are reviewed perti-
nent texts. In addition, the chapter provides a short introduction to integrity management processes.
Chapter 4, “Chemistry of Product Flow Impairment in Pipelines and Facilities,” describes pro-
cesses that impair the flow of oil, gas, and water in pipelines and facilities. These include inorganic
solids, organic solids, mixed deposits, and emulsions.
Chapter 5, “Mechanical Methods of Enhancement and Assessment of Pipeline Operations,”
reviews the many pigs (scrapers), moles, coiled tubing, and jetting equipment used to maintain the
lines and to place chemicals in them. The use of hydrostatic testing as well as in-line detection and
inspection devices is also reviewed.
Chapter 6, “Chemical and Mechanical Treatments To Enhance/Maintain Pipeline Operations,”
shows how chemicals and mechanical devices perform to prevent and inhibit the formation of
deposits, corrosion, and emulsions. This chapter also reviews gas dehydration methods, acid gas
removal, and flow enhancement chemicals.
Chapter 7, “Cleaning of Pipelines and Facilities,” describes the use of pigging and various chemi-
cal cleaning solvents to clear fouled pipelines and units in the facilities. Methods reviewed include
chemicals, testing, evaluation, and application of solvents.
Chapter 8, “Pipeline/Facility Maintenance Health, Safety, and Environmental Issues,” reviews
issues of health, safety, and the environment related to the maintenance of pipelines and facilities.
Chapters 1 through 8 each concludes with a “Summary and Lessons Learned” section that sum-
marizes the major findings revealed by the review of the technologies discussed and how this knowl-
edge can be applied to pipeline management projects. Chapters 3, 4, 5, 6, and 7 also have a section
titled “Best Practices and Case Studies for Chemical/Mechanical Management of Pipelines.” Here,
the science and engineering principles described in the earlier sections are illustrated through practi-
cal demonstrations of chemical/mechanical intervention and remediation.
Thus, Chapters 1 through 4 describe the root problems and Chapters 5 through 8 provide a large
range of mechanical and chemical solutions that can be accomplished in safe and environmentally
acceptable ways.

Definition of the Pipeline Treatment Environment and Summary


The scope of this book is limited primarily to chemical/mechanical intervention and enhancement
in the production (upstream) and transfer (midstream) oilfield environment. This includes flowlines
and gathering lines, associated-gas/liquid-treating facilities, and US Department of Transportation–
regulated crude-oil trunk and gas transmission pipelines. The discussion will not specifically include
problems in refineries or finished product pipelines. However, many of the needed techniques and
technologies, especially those for treating separation/gas-treating units, are similar to those described
in this book and could be applied with appropriate modifications. See Frenier (2001) for a review of
the cleaning of industrial equipment, including downstream oil/gas equipment. The scope is limited
primarily to internal flow assurance and integrity issues. Note that external corrosion problems are
usually addressed using cathodic protection and coating methodologies that are beyond the scope of
this book. However, some in-line inspections will detect and assess exterior corrosion as well as the
effectiveness of cathodic protection.
See Lazzari and Pedeferri (2006) for an explanation of cathodic protection and Munger and Vin-
cent (1999) for discussions of protective coatings. Also see the comprehensive books on pipeline
activities by Revie (2015), Peabody (2001), and Hevle (2012) and the Petrowiki (2014) web page.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 13 03/11/17 6:50 PM


BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 14 03/11/17 6:50 PM
Acknowledgments
The author acknowledges the extensive help and advice of T. D. Williamson, Incorporated. The
engineers at T. D. Williamson participated in the planning, evolution, and technical reviews of this
book. I wish especially to thank Richard B. Williamson, Chairman (Ret.) of T. D. Williamson, Incor-
porated, for his continuing support and help with the project. I also recognize the extensive advice
of Lee Shouse, Chuck Harris, Woody Smith, Eric Freeman, and Gordon Blair of T. D. Williamson.
In addition, I acknowledge the help and guidance of Rick Underwood of Enable Midstream Partners
(Enable 2013) (Ret.), Michael Volk Jr. of the University of Tulsa, Mohsen Achour of ConocoPhillips,
and Murtaza Ziauddin of Schlumberger. I also acknowledge the help and encouragement of David
Wint, Jeff Foote, and Ian Lisko as well as the members of the Tulsa Section of NACE, International.

xv

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 15 03/11/17 6:50 PM


BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 16 03/11/17 6:50 PM
Table of Contents
Foreword������������������������������������������������������������������������������������������������������������������������� vii
Preface������������������������������������������������������������������������������������������������������������������������������ ix
Acknowledgments���������������������������������������������������������������������������������������������������������� xv
1  Introduction to the Technology of Flow and Integrity Management������������������������ 1
1.1  Description of Pipelines and Operating Environments�������������������������������������������� 2
1.2 Need for Chemical and Mechanical Enhancements to Pipelines
and Facilities������������������������������������������������������������������������������������������������������������ 4
1.3  Economic and Market-Related Forces Affecting Pipeline Maintenance������������������ 5
1.3.1  General Economic Issues for Oilfield Treatments����������������������������������������� 5
1.3.2  Pipeline-Specific Economic Considerations������������������������������������������������� 6
1.4  Review of the Physics and Chemistry of Pipeline Interventions������������������������������ 7
1.4.1  Pipeline Materials for Construction and Pressure Requirements����������������� 7
1.4.2 Fluid Mechanics and the Effects of Fluids and Phases on Pipeline
Operations���������������������������������������������������������������������������������������������������� 9
1.4.3  Viscosity and Rheology of Fluids��������������������������������������������������������������� 17
1.4.4  Thermodynamics and Kinetics of Pipeline-Fouling Reactions������������������� 21
1.4.5  Surface Chemistry�������������������������������������������������������������������������������������� 24
1.4.6  Testing of Pipeline Fluids���������������������������������������������������������������������������� 28
1.5  Summary and Lessons Learned��������������������������������������������������������������������������� 29
2  From the Well to the Consumer��������������������������������������������������������������������������������� 31
2.1  Description of Well Production Fluids Entering the Pipeline System��������������������� 31
2.1.1  Aqueous Phases���������������������������������������������������������������������������������������� 31
2.1.2  Hydrocarbon Liquids���������������������������������������������������������������������������������� 33
2.1.3  Gaseous Phases���������������������������������������������������������������������������������������� 33
2.1.4 Solids��������������������������������������������������������������������������������������������������������� 34
2.1.5  Emulsions, Foams, and Solid Dispersions������������������������������������������������� 34
2.2  Effects of the Life Cycle and Reservoir Type on Pipeline Maintenance����������������� 34
2.2.1  Life Cycle of a Hydrocarbon-Producing Reservoir������������������������������������� 34
2.2.2  Conventional Reservoirs of Oil and Gas���������������������������������������������������� 38
2.2.3  Unconventional Reservoirs of Oil and Gas������������������������������������������������ 39
2.3  Problems Anticipated in Different Areas of Pipeline-Connected Systems������������� 44
2.3.1  Gathering Lines and Wastewater Lines����������������������������������������������������� 45
2.3.2  Surface and Subsurface Facilities�������������������������������������������������������������� 46
2.3.3  Water-System-Related Issues�������������������������������������������������������������������� 60
2.3.4  Summary of Problems in Surface/Subsurface Treatment Plants���������������� 62
2.3.5  Crude Oil Trunk and Gas Transmission Pipelines�������������������������������������� 63
2.4  Planning for Pipeline/Facility Reliability and Flow Assurance�������������������������������� 64
2.5  Summary and Lessons Learned��������������������������������������������������������������������������� 67
3  Corrosion Processes in Pipelines and Facilities����������������������������������������������������� 69
3.1  Fundamentals of Corrosion Chemistry������������������������������������������������������������������ 70
3.1.1  Effects of Pipeline Metallurgy on Corrosion Processes����������������������������� 72
3.1.2  Effect of Dissolved Salts and the Hydrocarbon Phases����������������������������� 76
3.1.3  Carbon Dioxide Corrosion�������������������������������������������������������������������������� 79

xvii

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 17 03/11/17 6:50 PM


xviii  Table of Contents

3.1.4  Hydrogen Sulfide Corrosion and Various Cracking Conditions������������������ 81


3.1.5  Oxygen Corrosion�������������������������������������������������������������������������������������� 84
3.1.6  Organic Acid Corrosion������������������������������������������������������������������������������ 85
3.1.7  Microbiologically Influenced Corrosion������������������������������������������������������� 85
3.1.8  Corrosion From Cleaning/Stimulation Fluids���������������������������������������������� 89
3.2  Consequences of Corrosion: Manifestations in Pipelines�������������������������������������� 91
3.2.1  Top-of-Line Corrosion��������������������������������������������������������������������������������� 91
3.2.2  Localized Corrosion����������������������������������������������������������������������������������� 95
3.2.3  Erosion/Corrosion and Impingement Damage������������������������������������������� 99
3.2.4 Summary of Pipeline Corrosion Locations
and Manifestation Types��������������������������������������������������������������������������� 102
3.3  Corrosion Processes in Facilities������������������������������������������������������������������������ 102
3.4  Corrosion and Inhibitor Testing���������������������������������������������������������������������������� 104
3.4.1  Common Methods for Corrosion Rate Determination������������������������������ 104
3.4.2 Specific Laboratory Test Methods for Studying Corrosion
and Inhibition�������������������������������������������������������������������������������������������� 112
3.4.3  Physical and Chemical Examination of Surfaces������������������������������������� 124
3.4.4  Comparison of Laboratory Tests�������������������������������������������������������������� 126
3.4.5  Field Monitoring of Corrosion Processes������������������������������������������������� 127
3.4.6 Placement of Probes and Coupons for
Maximum Effectiveness��������������������������������������������������������������������������� 135
3.5  Introduction to Integrity Management������������������������������������������������������������������ 136
3.5.1  Short History of Integrity Management Processes���������������������������������� 136
3.5.2  Preview and Nomenclature of Direct Assessment����������������������������������� 139
3.5.3  US State Regulations������������������������������������������������������������������������������� 139
3.5.4  US Federal Inspection Regulations���������������������������������������������������������� 142
3.5.5  Worldwide Pipeline Safety Regulations and Practices����������������������������� 144
3.5.6 Summary of Standards for Corrosion Assessments
and Safe Operation���������������������������������������������������������������������������������� 146
3.6  Corrosion Prediction and Assessment Processes���������������������������������������������� 147
3.6.1  Methods of Corrosion Prediction�������������������������������������������������������������� 147
3.6.2 Internal Corrosion Direct Assessment and Risk
Assessment Methods������������������������������������������������������������������������������� 152
3.7  Summary and Lessons Learned������������������������������������������������������������������������� 158
3.8  Best Practices and Case Histories for Corrosion Control������������������������������������ 159
3.8.1 Monitoring and Controlling Corrosion in an Aging
Sour-Gas-Gathering System: A Nine-Year Case History
(Nelson et al. 2007)���������������������������������������������������������������������������������� 159
3.8.2 Top-of-Line Corrosion in Multiphase Gas Lines:
A Case History (Gunaltun et al. 1999)������������������������������������������������������ 159
3.8.3 Special Issues Related to the Application of Direct
Assessment (Klechka 2009)�������������������������������������������������������������������� 159
3.8.4 Internal Corrosion Direct Assessment of Buried Steel
Pipeline in Chinese Oil Industry (Qimin and Guibai 2008)����������������������� 160
4  Chemistry of Product Flow Impairment in Pipelines and Facilities��������������������� 161
4.1  Why and Where Inorganic Scale and Organic Deposits Form���������������������������� 162
4.2  Inorganic Scale Formation���������������������������������������������������������������������������������� 163
4.2.1  Mineral Precipitation Scales��������������������������������������������������������������������� 163
4.2.2  Corrosion Product Scales������������������������������������������������������������������������ 172

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 18 03/11/17 6:50 PM


Table of Contents  xix

4.3  Organic Deposits������������������������������������������������������������������������������������������������� 173


4.3.1  Waxes (Paraffins)������������������������������������������������������������������������������������� 173
4.3.2 Asphaltenes��������������������������������������������������������������������������������������������� 174
4.3.3 Naphthenates������������������������������������������������������������������������������������������� 175
4.3.4  Clathrate Gas Hydrates���������������������������������������������������������������������������� 177
4.4 Mechanism of Formation of Mixed Deposits in Pipelines
and Facilities�������������������������������������������������������������������������������������������������������� 179
4.4.1  Overlapping Precipitation Regimes���������������������������������������������������������� 179
4.4.2  Pressure-Induced Deposits���������������������������������������������������������������������� 180
4.4.3  Alkaline Earth Scale and Organic Solids������������������������������������������������� 181
4.4.4  Corrosion-Triggered Mixed Deposits Including “Black Powder”���������������� 181
4.5  Deposits Found in Facilities��������������������������������������������������������������������������������� 187
4.6  Emulsions, Foams, and High-Viscosity Causes of Flow Problems��������������������� 188
4.6.1 Emulsions������������������������������������������������������������������������������������������������� 188
4.6.2 Foams������������������������������������������������������������������������������������������������������ 191
4.6.3  High Viscosity and Turbulent Flow������������������������������������������������������������ 191
4.7  Summary and Lessons Learned������������������������������������������������������������������������� 191
4.8 Best Practices and Case Studies of Flow Assurance Control
in Pipelines���������������������������������������������������������������������������������������������������������� 192
4.8.1 Review of Issues Associated with Inhibition of
Scale-Deposit-Covered Pipelines (Turgoose et al. 2006)������������������������� 192
4.8.2 A Case History of Heavy-Oil Separation in Northern
Alberta: A Singular Challenge of Demulsifier Optimization
and Application (Wylde et al. 2010)���������������������������������������������������������� 192
4.8.3 Best Practice for Organic Deposit Removal
(Montgomery et al. 1996)������������������������������������������������������������������������� 193
4.8.4 Multichemical Application Case History
(Shepherd et al. 2012)������������������������������������������������������������������������������ 194
5 Mechanical Methods of Assessment and Enhancement of
Pipeline Operations�������������������������������������������������������������������������������������������������� 195
5.1  Introduction to Pigs, Scrapers, and In-Line Inspection Devices�������������������������� 195
5.2  Types and Uses of In-Line Pipeline Devices�������������������������������������������������������� 199
5.2.1  Utility Pigs������������������������������������������������������������������������������������������������ 199
5.2.2  Pig Application Problems and Solutions�������������������������������������������������� 202
5.2.3  In-Line Operations Chemical Issues and Solutions��������������������������������� 207
5.2.4  Selection and Application of Mechanical Pigs������������������������������������������ 210
5.3  Hydrostatic Pressure Testing������������������������������������������������������������������������������� 214
5.3.1 General Procedures for Performing Hydrostatic
Pressure Tests������������������������������������������������������������������������������������������ 215
5.3.2  Water Quality and Wet Layup������������������������������������������������������������������� 216
5.3.3  Drying the Line and Dry Layup���������������������������������������������������������������� 217
5.4  In-Line Inspection Tools��������������������������������������������������������������������������������������� 218
5.4.1  Introduction to In-Line Inspection Technologies��������������������������������������� 218
5.4.2  Geometry and Mapping Devices�������������������������������������������������������������� 222
5.4.3  Magnetic Flux Leakage Technologies������������������������������������������������������ 224
5.4.4  Ultrasonic Technique Method������������������������������������������������������������������� 227
5.4.5  Eddy Current In-Line Inspection Methods������������������������������������������������ 230
5.4.6  Miniaturized In-Line Inspection Sensors�������������������������������������������������� 231
5.4.7  Cathodic Protection (CP) In-Line Tools���������������������������������������������������� 233

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 19 03/11/17 6:50 PM


xx  Table of Contents

5.4.8 Comparison/Selection of In-Line Inspection Methods


and Data Analysis������������������������������������������������������������������������������������ 235
5.4.9  Analyses of In-Line Inspection Data and Tool Performance��������������������� 238
5.4.10  Summary of In-Line Inspection Developments�������������������������������������� 243
5.5  In-Line Handling Equipment and Other Devices������������������������������������������������� 245
5.5.1  Traps (Receivers) and Launchers������������������������������������������������������������ 245
5.5.2  Pig Indicators and Tracking Devices��������������������������������������������������������� 247
5.5.3  Coiled Tubing for Conveying Tools Into Pipelines������������������������������������� 249
5.6  Summary and Lessons Learned������������������������������������������������������������������������� 251
5.7 Best Practices and Case Studies for Mechanical
Management of Pipelines������������������������������������������������������������������������������������ 251
5.7.1 Unpiggable Pipelines: What a Challenge for
In-Line Inspection! (Schmidt 2004)���������������������������������������������������������� 251
5.7.2 Case Study: Bayu Undan Pipeline and Darwin
Liquefied Natural Gas Project (Weatherford 2009)���������������������������������� 251
5.7.3 Single-Trip Pigging of Gas Lines During Late Field Life
(Mandke et al. 2002)�������������������������������������������������������������������������������� 251
5.7.4 In-Line Inspection on an Unprecedented Scale
(Brockhaus et al. 2015)���������������������������������������������������������������������������� 252
6 Chemical and Mechanical Treatments To Enhance/Maintain
Pipeline Operations�������������������������������������������������������������������������������������������������� 253
6.1  Corrosion Inhibitors and Inhibition Mechanisms�������������������������������������������������� 253
6.1.1 Review of Production and Pipeline Corrosion
Inhibitor Mechanisms������������������������������������������������������������������������������� 257
6.1.2 “Sweet” Corrosion Inhibitors��������������������������������������������������������������������� 258
6.1.3 Sour Brine Corrosion Inhibitors and Inhibition of
Microbiologically Influenced Corrosion����������������������������������������������������� 262
6.1.4 Inhibitors for Gas-Containing Lines and Multiphase Lines����������������������� 263
6.1.5  Selection of Corrosion Inhibitors for Use in Pipelines������������������������������ 269
6.1.6  Chemistries for Inhibition of Cleaning Fluid Acid Attacks������������������������� 271
6.2  Inorganic Scale Inhibitors and Inhibition Mechanisms���������������������������������������� 271
6.3  Organic Deposit Inhibition����������������������������������������������������������������������������������� 275
6.3.1  Paraffin Deposition and Pour Point Inhibitors������������������������������������������� 275
6.3.2  Asphaltene Inhibitors�������������������������������������������������������������������������������� 275
6.3.3  Gas Hydrate Inhibitors����������������������������������������������������������������������������� 277
6.3.4  Calcium/Sodium Naphthenate Inhibition�������������������������������������������������� 280
6.4  Flow Enhancement, Biocides, and Oxygen Scavengers������������������������������������� 281
6.4.1  Demulsifiers, Foaming Agents, and Defoamers��������������������������������������� 281
6.4.2  Flow Enhancers: Drag-Reducing Agents������������������������������������������������� 284
6.4.3 Biocides���������������������������������������������������������������������������������������������������� 288
6.4.4  Oxygen Scavengers��������������������������������������������������������������������������������� 290
6.5  Acid Gas Scavenger Chemistry�������������������������������������������������������������������������� 291
6.5.1  Oxidizing Agents�������������������������������������������������������������������������������������� 292
6.5.2 Aldehydes������������������������������������������������������������������������������������������������ 292
6.5.3  Nitrogen-Based Scavengers�������������������������������������������������������������������� 292
6.6  Chemistry for Producing Pipeline Gels���������������������������������������������������������������� 294
6.7  Fluid Additive Injection Methods and Equipment������������������������������������������������� 297
6.7.1  Continuous Treatments With Pipeline Chemicals������������������������������������� 299
6.7.2 New Formulations for Application of Production Corrosion
and Scale Inhibitors���������������������������������������������������������������������������������� 306

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 20 03/11/17 6:50 PM


Table of Contents  xxi

6.7.3 Applications of Treatment Chemicals in the


Batch Mode and Using Pigs��������������������������������������������������������������������� 307
6.8  Summary and Lessons Learned������������������������������������������������������������������������� 315
6.9 Best Practices and Case Histories for Maintenance
Treatments of Pipelines��������������������������������������������������������������������������������������� 316
6.9.1 Monitoring and Controlling Corrosion in an Aging
Sour-Gas-Gathering System: A Nine-Year Case History������������������������� 316
6.9.2 Black Powder in Pipeline: Cleaning Program
(Sirnes and Gundlach 2012)�������������������������������������������������������������������� 316
6.9.3 Beneficial Effects of Chemical Treatment and Maintenance
Pigging Programs in Returning an Offshore Pipeline to
Pre–Hurricane Ike Conditions Following a Breach and the
Ingress of Seawater and Sand, and the Effects of
Bacteria-Generated H2S (Powell et al. 2010;
Powell et al. 2012)������������������������������������������������������������������������������������ 316
6.9.4 Integrated Production Chemistry Management of the
Schoonebeek Heavy-Oil Redevelopment in the Netherlands:
From Project to Startup and Steady-State Production
(Shepherd et al. 2012)������������������������������������������������������������������������������ 316
7  Cleaning of Pipelines and Facilities������������������������������������������������������������������������ 319
7.1  Maintenance Pigging������������������������������������������������������������������������������������������� 319
7.2 Automated Methods and Mechanical Equipment for
Maintenance Cleaning����������������������������������������������������������������������������������������� 322
7.2.1  Horizontal Multiple-Pig Launchers����������������������������������������������������������� 323
7.2.2  Vertical Multiple-Pig Launchers���������������������������������������������������������������� 323
7.2.3  Automatic Sphere Launchers������������������������������������������������������������������� 325
7.3 Chemical Solvents and Mechanisms for Cleaning Pipelines
and Facilities�������������������������������������������������������������������������������������������������������� 326
7.3.1  Introduction to Solvent Chemistry������������������������������������������������������������ 327
7.3.2  Chemicals for Removing “Acid-Soluble” Inorganic Solids������������������������ 331
7.3.3  Solvents for Alkaline Earth Sulfates��������������������������������������������������������� 340
7.3.4  Corrosion Inhibitors for Inorganic Scale Removal Solvents��������������������� 343
7.3.5  Solvents for Organic Solids���������������������������������������������������������������������� 345
7.3.6  Solvents for Mixed Deposits��������������������������������������������������������������������� 356
7.4  Testing of Deposits To Develop Solvents������������������������������������������������������������� 357
7.4.1 Work Flow for Evaluation and Treatment of Inorganic,
Organic, or Mixed Deposit������������������������������������������������������������������������ 358
7.4.2  Chemical Identification of a Deposit��������������������������������������������������������� 359
7.4.3  Evaluation of Solvents for Removing Inorganic/Organic Deposits����������� 360
7.5  Application of Chemical Cleaning in Pipelines���������������������������������������������������� 365
7.5.1  Liquid-Phase Solvents����������������������������������������������������������������������������� 365
7.5.2  Examples of Treatment of Pipelines To Remove Mixed Deposits������������� 366
7.5.3  Uses of Gelly Pigs in Pipeline Cleaning��������������������������������������������������� 370
7.5.4  Use of Foamed Fluids in Pipeline Cleaning and Treatment��������������������� 373
7.6  Cleaning of Topside Facilities������������������������������������������������������������������������������ 377
7.6.1  Typical Topside Equipment Needing Cleaning����������������������������������������� 377
7.6.2  Methods for Cleaning Topside Facilities and Heat Exchangers��������������� 378
7.6.3  On-Site Cleaning Issues�������������������������������������������������������������������������� 381
7.6.4  Examples of Cleaning of Oil and Gas Facilities��������������������������������������� 382

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 21 03/11/17 6:50 PM


xxii  Table of Contents

7.7 A Company’s Best Practice for Maintaining Gas-Gathering and


Transmission Lines���������������������������������������������������������������������������������������������� 383
7.8  Summary and Lessons Learned������������������������������������������������������������������������� 383
7.9 Best Practices and Case Histories for Renovation/Remediation
of Pipelines and Facilities������������������������������������������������������������������������������������ 384
7.9.1 Pigging of Pipelines With High Wax Content (Tordal 2006)��������������������� 384
7.9.2 Recommissioning of Mothballed Pipelines Offshore California:
A Success Story of Cleaning, Pigging, Monitoring, and Integrity
Management (Wylde 2009)���������������������������������������������������������������������� 384
7.9.3 Available Methodologies Concerning the Treatment and
Removal of Sand From Pipelines, With Associated
Case Studies (Mackay 2013)������������������������������������������������������������������� 384
7.9.4 Innovative Methodology for Cleaning Pipes: Key to
Environmental Protection (Buzelin and Lima 2008)��������������������������������� 385
7.9.5  Cleaning the Valhall Offshore Oil Pipeline (Marshall 1990)���������������������� 385
7.9.6 Conclusions���������������������������������������������������������������������������������������������� 386
8 Pipeline/Facility Maintenance Health, Safety, and
Environmental Issues����������������������������������������������������������������������������������������������� 387
8.1 Introduction���������������������������������������������������������������������������������������������������������� 387
8.2 Health and Safety Considerations During Pipeline/Facility
Maintenance Operations������������������������������������������������������������������������������������� 387
8.2.1  General Considerations for All Oil/Gas Operations���������������������������������� 387
8.2.2  Specific Pipeline/Facility Considerations�������������������������������������������������� 393
8.3  Health, Safety, and Environmental Management������������������������������������������������ 395
8.3.1 Chemical Selection To Enhance Health, Safety,
and Environmental Management Compliance����������������������������������������� 395
8.3.2  Chemical Development Processes���������������������������������������������������������� 396
8.3.3 Chemical Handling Processes To Promote Health, Safety,
and Environmental Management Improvements�������������������������������������� 401
8.4 Handling, Reuse, and Disposal of Pipeline/Facility Treating
Fluids and Solids������������������������������������������������������������������������������������������������� 403
8.4.1 Planned Waste Disposal Options for Pipeline/Facility
Cleaning Fluids���������������������������������������������������������������������������������������� 403
8.4.2 Control and Remediation of Spills in Water Bodies
and on Land��������������������������������������������������������������������������������������������� 407
8.4.3 Use of Remediation Chemicals in Bodies of Water
and Proposed Mechanism of Action�������������������������������������������������������� 411
8.5  Summary and Lessons Learned (Chapter and Book)����������������������������������������� 415
8.5.1  Chapter 8 Lessons Learned��������������������������������������������������������������������� 415
8.5.2  General Lessons From This Book������������������������������������������������������������ 415
Appendix A: Glossary�������������������������������������������������������������������������������������������������� 417
Appendix B: Nomenclature������������������������������������������������������������������������������������������ 421
References�������������������������������������������������������������������������������������������������������������������� 423
Index������������������������������������������������������������������������������������������������������������������������������ 475

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 22 03/11/17 6:50 PM


Chapter 1

Introduction to the Technology of


Flow and Integrity Management

This chapter introduces the broad scope of this book. Here, you will find the concepts you need to
know to deal with issues you may encounter. These concepts will help you determine the specific
actions needed to reduce negative impacts on the delivery of products to a consumer. Important sub-
jects introduced here include the need for interventions and the economics of interventions. In addi-
tion, this chaper provides a review of physical and chemical principles that underlie the processes
described in this book.
Appendix A of this book contains a glossary of terms that are especially relevant to pipeline-
related usages, given that various names are used for the same cleaning and assessment tools and
other devices in different parts of an overall supply system, as well in different locations within the
petroleum industry worldwide. In addition, various hydrocarbon products are commonly known by
a variety of names. Some of the definitions come from the Schlumberger Glossary (Schlumberger
2012), PAPA (2013), and the Pipeline and Hazardous Materials Safety Administration (PHMSA)
(PHMSA 2016). These references (websites) can be accessed for definitions of important additional
pipeline or oil/gas terms. A list of symbols used in this document is in Appendix B. Acronyms are
defined the first time they occur in each chapter.
All the subjects introduced in this chapter are discussed in more detail in later chapters, of course.
Maintaining flow [also known as flow assurance (FA)] and ensuring the physical integrity of the
pipelines and facilities themselves are core necessities for providing useful products to the ultimate
consumers, as well as encouraging a healthy worldwide economy. The fluids are extremely vari-
able. They frequently consist of multiple phases and are subject to unpredictable changes in pres-
sure, temperature, and composition. Therefore, the formation of solids and emulsions is a constant
threat. Solids that result from equilibrium changes can decrease the effective diameter of the pipe,
completely block the pipe, or change the viscosity of the fluid. “Wet”-gas lines are frequently
impeded because of the amount of condensed low-molecular-weight hydrocarbons present as the
temperatures are decreased. Additional equilibrium changes can also cause corrosion or degrada-
tion of the pipe wall, leading to a leak, loss of product, and possibly major environmental damage
and loss of life.
Many chemical and mechanical processes are available and in current use to maintain the constant
flow of vital products and to assure the integrity of this important infrastructure. This book has been
written to provide an improved understanding of the role of chemical reactions and mechanical
devices for enhancing and maintaining the conveyance of oil, gas, and related products. Among the
many books that describe the thousands of chemicals used in the pipeline industry are those by Fink
(2003, 2012) and Kelland (2009, 2013). In addition, there are the works of Frenier and Ziauddin
(2014), Frenier and Ziauddin (2008), and Frenier et al. (2010), which describe many of the chemical

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 1 03/11/17 6:50 PM


2  Chemical and Mechanical Methods for Pipeline Integrity

reactions that occur in the pipeline environment. In this book, the focus is on the application of these
chemistries to the maintenance of a productive pipeline environment. This study will demonstrate
the synergism between these chemical and mechanical processes and will make the case for the
necessity of using mechanical devices in these applications.
Books that describe additional details of pipeline maintenance in general include Mohitpour
et al. (2010), which is part of the pipeline series from the American Society of Mechanical Engineers
(New York). A book by Mokhatab et al. (2006) covers many discussions of gas transmission tech-
nologies, including details of gas treatment chemistries and plants. The extensive “handbook” by
Revie (2015) contains 52 peer-reviewed chapters that cover various aspects of integrity and safety
of both production and transmission pipelines. Kennedy (1993) also describes aspects of pipelines,
including types of pipelines, pipe manufacture and coating, and fundamentals of pipeline design, as
well as types of pumps and compressors in use when his book was published.

1.1  Description of Pipelines and Operating Environments


API (2007) described several categories of petroleum/gas-related pipelines. Crude oil gathering
lines (approximately 2- to 8-in. outer diameter (OD) collect product from various wells, transmit it
to collection/treatment facilities, and then move it onward into “trunk” (transmission) lines. These
can be approximately 8- to 25-in. OD, and a few (such as the Alaskan pipeline) may be as much as
48-in. OD. The larger lines may cross state, territorial, and international borders. Most pipelines on
land are buried from several up to tens of feet in the ground. In specific areas such as the Alaskan
pipeline or for some river crossings, pipelines may also be placed above ground.
Subsea gathering lines and trunk lines are usually on the sea floor, but they may also be buried.
Risers then convey the well products to surface (or subsea) treatment facilities. There are more than
40,000 miles of crude oil gathering lines and 55,000 miles of trunk lines just in the US. Once at the
treatment facilities, the crude oil is refined into hundreds of different products.
Natural gas [and natural gas liquids (NGLs)] is collected in more than 200,000 miles of gathering
lines (PHMSA 2012) in the US and transmitted in more than 270,000 miles of larger (20- to 42-in.-
OD) transmission lines. For home and industrial use, the transmission lines connect to distribution
systems of local/regional gas companies (API 2007a). Note that gas is usually treated near the well-
head and oil fields to remove water, NGLs, and acid gases (Chapter 2 explains this).
Worldwide as of 2010, according to Mohitpour et al. (2010), there were more than 500,000 miles
of gas pipelines alone. PHMSA (2012) provided a map (Fig. 1.1) that shows the general location of
gas (red) and hazardous liquid (blue) pipelines in the US. The US Central Intelligence Agency (CIA
2008) reported that as of 2008, there were approximately 62,000 miles of lines in Canada, 160,000
in Russia, and 40,000 in China.
As a result of the increases in production from shale oil/gas wells, the need for new pipelines of
all types is expected to increase. INGAA (2011) predicted the need for new natural gas lines through
2035 and estimated new needed infrastructure (see Fig. 1.2). This increase in production from shale
formations is increasing the flow of liquid hydrocarbons that also will stimulate the construction of
thousands of miles of new lines to support that part of the oil/gas industry (Boyer et al. 2011; PI
2012). Note that this includes 16,000 miles/yr of new gathering lines that are especially susceptible
to blockages and corrosion (see Sections 2.3.1 and 3.2). In addition, Tubb (2013) estimated that as
of 2013, more than 116,000 miles of pipelines were planned or under construction worldwide.
Unless liquefied, compressed (packed), or stored in underground cavities or old wellfields, the gas
is used as produced and is more difficult to store than liquid hydrocarbons. In some mixed hydro-
carbon wellfields, the gas is reinjected to maintain reservoir pressure and has been flared on-site if
storage or transmission lines are not available (Fig. P-2). As noted in this book’s preface as well
as in Jacobson (2014), pipelines are one of the most economical and energy-efficient methods for
transporting any bulk product (see Fig. 1.3).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 2 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  3

Pipelines
Hazardous liquid
Gas transmission

Fig. 1.1—Gas and hazardous pipelines in the US (PHMSA 2012).

Added Gas infrastructure 2011-2020 2011-2035 Average


annual
Inter-regional pipeline capacity (Bcfd) 29 43 1.7
Miles of transmission mainlines (1,000s) 16.4 35.6 1.4
Miles of laterals to/from power plants, 6.6 13.9 0.6
storage fields and processing plants
(1,000s)
Miles of gathering lines(1,000s) 491 1,043 42
Average pipe size (in.) 30 29 30
Inch-miles of laterals to/from power 142 304 12
plants, storage fields, and processing
plants (1,000s)

Average pipe size (in.) 22 22 20


Inch-miles of gathering lines (1,000s) 592 1,518 61
Average pipe size (in.) 4 4 4
Compression of pipelines (1,000 HP) 3,039 4,946 197
Gas storage (Bcf) NA 589 24
Processing capacity (Bcfd) 18.1 32.5 1.3

Fig. 1.2—Added natural gas infrastructure to 2035 (INGAA 2011).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 3 03/11/17 6:50 PM


4  Chemical and Mechanical Methods for Pipeline Integrity

Fig. 1.3—Costs and energy of transportation of bulk products (Jacobson 2014).

Pipelines are highly regulated by US states, the US federal government, and international organi-
zations (see Section 3.5), with the safety and integrity of the lines a constant goal of all stakeholders.
General information on pipelines and approximately 30 references to pipeline systems are found in
Petrowiki (2014).

1.2  Need for Chemical and Mechanical Enhancements to Pipelines and Facilities
The drivers for needing chemical and mechanical interventions and enhancements are the overlap-
ping concepts of flow assurance (FA) and integrity management (IM).
Brown (2002) described FA as the “production operation that generates a reliable, manageable,
and profitable flow of fluids from the reservoir to the sales point.” FA is especially critical for deep-
water assets. Because of limited access to the seafloor infrastructure in deepwater areas, blockages
and corrosion in tubulars and lines from deposit formation or from other causes may lead to expen-
sive workovers (Brown 2002).
IM is specifically associated with the risks of a pipeline failure. Stephens and Playdon (1998)
noted that pipeline characteristics (or attributes) must first be evaluated to produce a line-specific
estimate of the failure probability for each segment within the system as a function of failure cause,
which might be

• Metal-loss corrosion
• Mechanical damage
• Ground movement
• Crack-like defects

Then, an estimate of the potential consequences of segment failure must be made in terms of three
distinct consequence components: lifetime line safety, environmental damage, and economic impact.
Cause-specific failure probability estimates are then multiplied by a global measure of the loss
potential associated with the different consequence components to produce a single measure of
operating risk for all failure causes associated with each segment. Segments can then be ranked by
failure cause and according to the estimated level of risk. The author of this book contends that IM

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 4 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  5

is inherently a part of FA because, for example, a failed, leaking pipeline segment cannot effectively
and efficiently move products to the consumers.
In many countries, including the US and Canada, various inspections are required by governmental
regulations (PHMSA 2012). Idem (2015) claimed that this was a response to a number of serious pipe-
line ruptures that had devastating environmental effects on land and water. Thus, the US Department of
Transportation PHMSA announced proposed regulations to require that all hazardous liquid pipelines
have a system for detecting leaks and establish a timeline for inspections of affected pipelines following
an extreme weather event, natural disaster, or operator negligence. President Barak Obama’s transpor-
tation secretary, Anthony Foxx, said, “Hazardous liquid pipelines crisscross the country and pipeline
failures can have profound impacts on local communities and the environment.” He declared, “This
proposed rule is an important step forward to enhance safety, and protect people and the environment.”
The mechanical and chemical enhancements and interventions described in this book are an inte-
gral part of both FA and IM by locating, anticipating, and assessing possible damages and interrup-
tions of flow conditions and then providing proactive solutions to prevent or remedy the situations.
Details of IM systems are provided in Section 3.5.

1.3  Economic and Market-Related Forces Affecting Pipeline Maintenance


This section includes a short review of the economic issues that affect most oil/gas treatment proj-
ects as well as concerns specific to the pipeline industry.

1.3.1  General Economic Issues for Oilfield Treatments. Chemical and mechanical treatments
are performed for many different reasons, and the general market for all oilfield-related chemicals
may exceed USD 15 billion (Freedoniagroup 2008). FA methods such as the use of corrosion, scale,
and organic deposit inhibitors, as well as the application of demulsifieres, are used to prevent prob-
lems from occurring. Many chemicals are also used in stimulation, in enhanced-oil-recovery (EOR)
activities (Kelland 2009; Frenier and Ziauddin 2014), and in pipeline infrastructure maintenance.
In the case of tight formations, hydraulic-fracturing methods are used initially to cause enough
conductivity to allow the well to produce economical amounts of hydrocarbons. Fracturing activi-
ties, along with directional drilling in shale oil/gas formations may produce enough hydrocarbons
to change the entire energy calculus of countries and regions of the world. Jaffe (2010) said that
because of these methods, there may be enough gas potentially available to change some geopoliti-
cal balances of power. Also, discoveries of retrievable liquid hydrocarbons associated with various
shale plays (RRCT 2011) may likewise greatly affect hydrocarbon supplies and economics and
pipeline activities (see Fig. 1.2).
For all the processes used, there is an economic element that also includes environmental and
political questions (Hess 2010). These economic issues must be taken into account in business
calculations before a decision is made to pump chemicals or to perform pipeline maintenance
activities. Initially, a calculation must determine whether to consider any enhancements. Then one
must decide which method should be used. Deciding which type of treatment to use may involve
more than economic issues, of course. Examples of other issues include the formation type, location,
equipment availability, time constraints, local availability of chemicals, and urgency of treating a
problem. If enough information is available to add up all the potential costs as well as the loss/gain
of products, then the economics equations described next can be used.
Economides and Boney (2001) discussed calculation methods and decisions that should be made
to decide on a project. They examined several indicators, shown here.
Payout Time. Payout time includes the total costs associated with the project, but not the value of
the money or the profit. This is a measure of liquidity for the project.
Net Present Value. Net present value (NPV), shown in Eq. 1.1, is the definition for cumulative
discounted cash flow (CDCF). The NPV is the maximum of this CDCF. NPV gives a dollar value
added to the property at present time. If it is positive, the investment is attractive; if it is negative,

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 5 03/11/17 6:50 PM


6  Chemical and Mechanical Methods for Pipeline Integrity

it means an undesirable investment. According to Economides and Boney (2001), NPV is the most
widely used indicator showing a dollar amount of net return. Here ∆USDn is the incremental revenue
(minus the incremental expenses and taxes that are a result of operations), n is the time period incre-
ments (e.g., years), and i is the interest rate.

n
∆USD n
NPV = ∑ − cost .���������������������������������������������������������������������������������������������������������� (1.1)
n =1 (1 + i )
n

Rate of Return. The rate of return (ROR) is a comparison with other investments and is deter-
mined by setting i = 0.
Return on Investment. The payback is the total amount of money earned from the investment in
the project. This is the return on investment (ROI). Investment relates to the amount of resources put
into generating the given payback.

( Payback − Investment )
ROI = • 100 .�������������������������������������������������������������������������������������� (1.2)
Investment
Corporate Goals and Risk. Corporate goals and risks are less tangible, but are important con-
siderations for judging projects. These also include matching or exceeding the capabilities of the
competition in some markets. Risks may include environmental or ecotox issues.

1.3.2  Pipeline-Specific Economic Considerations. Pipeline maintenance includes the use of chem-
icals and mechanical equipment for FA as well as for IM. Many projects will require consideration
of the general guidelines described in Section 1.3.1. However, calculations of conditions that lead to
a failure, especially calculations of corrosion susceptibilities, are mandated by governmental regula-
tions (Section 3.5) as well as by good business practices. Several specific examples are noted here.
NACE (1994; 2011) described two additional economic calculations. These include the dis-
counted payback period (DPBP) and benefit/cost ratios, explained next.
Discounted Payback Period. The DPBP method requires a calculation of discounted cash
inflow (DCI).

(Actual Cash Inflow)


DCI = ,���������������������������������������������������������������������������������������������������� (1.3)
(1 + i)n

where I is the discount rate and n is the time period in which the cash inflow happens. Then,

B
DPBP = A + . ������������������������������������������������������������������������������������������������������������������������ (1.4)
C

A is the last period with negative discounted cumulative cash flow, B the absolute value of dis-
counted cumulative cash flow at the end of period A, and C the discounted cash flow during the
period after A.
Benefit/Cost Ratios. The benefit/cost ratio is the ratio of the benefits of a project or proposal,
expressed in monetary terms, relative to its costs, also expressed in monetary terms. All benefits and
costs should be expressed in discounted present values (see Eq. 1.3 and 1.4; NACE 2011). Also note
that Appendix A and Appendix B in NACE (2011) show examples of these calculations made on the
basis of pipeline examples. This book (NACE 2011) has more than 100 references that may be of
value for these discussions.
Wei et al. (2009) described an integrated approach, made on the basis of corrosion modeling and
laboratory testing, to optimize the use of carbon steel in corrosive service for applications such as

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 6 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  7

downhole tubulars, pipelines, and facilities. This approach presents economic advantages, such as
reducing the use of expensive corrosion-resistant alloys while ensuring the operational integrity of
equipment and facilities. A key part of this integrated approach is to apply reliable corrosion models
underpinned with laboratory data (see Section 3.4). These authors claimed that to be most effective,
the models should account for the relevant chemistry and physics of the corrosion process, including
the effects of detailed water chemistry, liquid hydrocarbons, and the degree of protection from iron
carbonate or iron sulfide scales (these pipeline conditions and chemistries are described in Section 3.1
of this book). Thus, in their report Wei et al. (2009) claimed that ideally, models should account for
variations in conditions and flow characteristics along the length of a wellbore or pipeline. Addi-
tional information on modeling is in Section 3.6.
Dawotola et al. (2011) proposed a data-driven approach to find the optimal inspection interval (see
Section 3.5) for a petroleum pipeline that is subject to long-term corrosion failure. This approach is
claimed to account for the determination of both the failure frequency and consequences of failure.
Three forms of corrosion are studied for use in the model: uniform corrosion (Section 3.1), pitting
corrosion (Section 3.2.2), and stress corrosion and other forms of cracking (Section 3.2.2).
Failure frequency is estimated by fitting historical pipeline failure data into either a homogeneous
Poisson process (Wikipedia 2014b)—a stochastic program that counts the number of events and the
time points at which these events occur in a given time—or a power-law process. The consequences
of corrosion attack are calculated in terms of economic loss, environmental damage, and human
safety and are determined for small leaks, large leaks, and rupture of pipeline. Failure frequency and
consequences are both used to estimate total loss resulting from pipeline operation. A risk-based IM
optimization of the pipeline is obtained by minimizing the economic loss of pipeline, taking human
risk and maintenance budget as constraints.

1.4  Review of the Physics and Chemistry of Pipeline Interventions


The next sections (1.4.1 through 1.4.6) provide an introduction to the mechanical and chemical
principles that control and inform all the maintenance processes described in the subsequent chap-
ters of this book. For those readers skilled in the art and sciences of pipeline operations, this section
(or parts of it) is optional and may be skipped. The technologies reviewed here include pipeline
construction materials, simplified fluid dynamics, viscosity, rheology, simplified thermodynamic
principles, surface chemistry, and testing of well/pipeline fluids. These principles are applied in the
remainder of this book to explain and connect the various processes discussed.

1.4.1  Pipeline Materials for Construction and Pressure Requirements. Most oilfield pipe goods
are constructed using various carbon steels and must comply with ANSI/API SPEC 5L (2011) and/
or ANSI/NACE MR0175-2009 (2009) standards, depending on the expected product, temperatures,
and flow rates. High-strength carbon steel pipes (up to 80 ksi) are in use onshore and subsea. For
some subsea pipelines or facilities in sour or high carbon dioxide (CO2) service, 13% Cr martensitic
steel may be specified (see Section 3.1 for more details on metallurgy’s role in corrosion). NACE
SPO106-2006 (2006) provides design considerations that apply to pipelines made of steel used to
transport natural and manufactured gas, crude oil, and refined products for the control of internal
corrosion. A corrosion specialist should be consulted during pipeline design and construction. Two
major recommendations are to completely determine the compositions of the gases, liquids, and
other fluids and to design for flow velocities that mitigate corrosion and pipeline cracks. Chapter 4
of Heidersbach (2011) reviews the materials and metallurgy of oilfield equipment. He noted that
pipeline metals (mostly various steels) are specified in API SPEC 5L (2011). As with most oilfield
materials, specifications are made on the basis of performance standards, not chemical specifica-
tions such as SAE (2015) standards (e.g., for 1010 carbon steel) that have maximum or minimum
elemental limits. He also noted that most line piping is composed of “low-carbon steel” (less than
0.3% C). Other chemical elements may be added (Mn and small amounts of Cr, Ni, for example)

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 7 03/11/17 6:50 PM


8  Chemical and Mechanical Methods for Pipeline Integrity

to produce a “high-strength/low-alloy steel” if field specifications are to be met. Chapters 3, 4, and


5 in Revie (2015) discuss various aspects of pipeline design to address stress and fracture control.
Here, Section 3.1.1 provides additional details of the metallurgies of pipeline steels. Also, Petrowiki
(2014) provides a comprehensive list of pipeline metal grades, along with references.
Most pipeline segments are connected using various welding methods—these include electrical as
well as laser methods (see Komizo 2008)—and the welds require extensive inspections and testing.
Fig. 1.4 shows a drawing of the weld bead, the heat-affected zone (HAZ), and possible inclusions
(or voids) and cracks that could affect reliability. The weld bead and the HAZ are subjected to tem-
peratures that will change the metal’s microstructure (and the weld bead is essentially a casting),
so these differences can affect the corrosion and mechanical properties compared to the base metal.
Various heat treatments, including annealing, quenching, tempering, and normalizing, can be used
to reduce some effects of the welding and forming processes (see Heidersbach 2011).
In the US, many requirements for the allowed
pressure ratings of pipelines are driven partially by
Class Location F
Department of Transportation regulations (see Sec-
1 10 or less occupied buildings 0.72 tion 3.5). Wint (2011) describes the Title 49 CFR
2 Greater than 10, but less than 46 0.6 §192.105 (USDOT/OPS 2004) requirements for
3 More than 46 0.5 design of steel pipe in service in high-consequence
4 Where buildings of more than 4 0.4 areas (HCAs) such as urban environments. The most
stories are prevalent restrictive requirements (Table 1.1) have the lowest
value for the maximum-allowed design operating
Table 1.1—HCA location class.
pressure (MAOP). This value is calculated with

 2St 
MAOP =   ( E • F • T ) . �������������������������������������������������������������������������������������������������������� (1.5)
 D 

The values in Eq. 1.5 are

MAOP = maximum-allowed design operating pressure,


S = yield strength [specified minimum yield strength (SMYS)],
D = outside diameter,
t = wall thickness,
F = design factor HCA Class, §192.111,
E = longitudinal joint factor, §192.113,
T = temperature derating factor, §192.115.

Weld Bead

Heat-affected zone

Pipe wall

Defects
Cracks

Fig. 1.4—Pipeline weld and defects.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 8 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  9

There are various factors that can affect the thickness of the pipe (such as corrosion damage;
see Chapter 3) and may cause the pipe to be derated (also see DNV-RP-F101 2010; Leewis 2003).
Ashby (2013) noted that the classification of the HCA location classes may change in the future,
so more miles of piping will fall into the most restrictive class. Note that the complexities of the
regulations are beyond the scope of this book; furthermore, current regulations should always be
reviewed carefully when making these calculations. These restrictions and classification are covered
in Chapters 3 and 5.
Polymers (plastic-fiberglass) as well as polymer-coated steel pipe segments are being placed in
many sections of operating and transmission environments. The different general types of polymers
and coatings include thermoplastic resins that soften when heated and thermosetting resins that
undergo a chemical reaction during the formation process. Thermoplastic plastics include high-density
polyethylene, polyvinyl chloride (PVC), and fluoro polymers. Thermosetting plastics include epoxies,
polysilicones, and polyurethanes.
Elastomers are rubbery thermosetting plastics that may have less crosslinker and possibly other
additives that allow deformation for use in seals, packers, or other such materials. Also, mixtures of
resins with fiberglass or carbon fibers are present in some uses.
The benefits of plastic piping may include lower weight and better corrosion resistance, as well as
easier connections than otherwise possible. However, many FA issues such as scaling and emulsions
may affect flow in polymer/resin-coated lines. In addition, pressure limitations and cost consider-
ations may dictate the choice between steel and resins. Plastic piping, including fiberglass, is used in
some downhole applications and for transportation of oilfiled water and flowback fluids.
Pipeline systems also involve a number of ancillary devices that include pumping/compression
equipment, valves, taps, pig-launching and -retrieval equipment, and devices for introduction of
different treating chemicals. Fig. 1.5 is a diagram of a generic system of wells, gathering lines, and
flowlines that terminate in a separation/treating facility or a refinery (Idachaba 2016). This diagram,
while based on a design for a land-based oil/gas system, also applies to subsea systems. These sys-
tems are all designed for gas, crude oil, and mixed well fluids. Pigging problems associated with
valves and the like are described in Section 5.2.2.
If the pipeline is a subsea system (Albert et al. 2011), “risers” (or pipes) may convey the various
fluids to the surface. Some fluids may also go through subsea piping to injection wells. When all
the on-site processing is complete, the crude oil and gas enter pipeline systems (or other methods
of transport) that send the products to a “consumer.” The long-distance transition systems will also
have periodic pumping/compression stations, innumerable valves, and laterals and other devices for
measurement and maintenance. Some of these devices are described in subsequent chapters.

1.4.2  Fluid Mechanics and the Effects of Fluids and Phases on Pipeline Operations. The
mechanics of fluid flow in the multiple pipeline segments affects and is affected by the changing
nature of the fluid phases (details of the chemistries of the various phases, which include aqueous
liquid and hydrocarbon liquids, gases, and solids, are in Section 2.1). The pipelines themselves
are also affected by the production of scale, which is inorganic deposits (as noted in Chapter 4).
The fluid mechanics at any point in the pipeline system and the presence of deposits will affect
the causes of corrosion damage (see Chapter 3), and pressure and flow changes affect deposition
of scale. Thus, there is a feedback loop in which mechanical and chemical conditions influence
each other. All these conditions must consequently be understood and anticipated for successful
FA and IM.
The following subsections provide an introduction to single-phase and multiphase flow in pipes.
More details are in Brill (1987), Asante (2002), Shoham (2006), and Falcimaigne and Decarre
(2008).
Single-Phase Flow. Single-phase flow equations and concepts provide a starting point for the
much-more-complex multiphase flow conditions that actually exist in many pipeline segments.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 9 03/11/17 6:50 PM


10  Chemical and Mechanical Methods for Pipeline Integrity

a) Typical pipeline units (and key)


Export terminal/
Well location Flow station Small manifold Main manifold refinery

1 2 3 4 5
Flowlines Delivery lines Pipelines Trunk lines

Row = Pipeline b) Details of a complex well field and pipeline organization


right of way

Well locations 2 5
1

2
Row
Row
3 Small manifold
Row

Oil field Row 4

Flow station
Row 2 Row
4
Main manifold Row

Access
1 roads

Fig. 1.5—Generic layout of a pipeline system (Idachaba 2016).

The pressure drop over a distance, L, of a single-phase incompressible fluid can be obtained from
the mechanical-energy-balance equation (Economides et al. (1994) as

∆pL = ∆pPE + ∆pF + ∆pKE.���������������������������������������������������������������������������������������������������������� (1.6)

In an expanded form, it becomes

 g 2 f f ρν 2 l ρ
∆pL =   ρl sin θ + + ∆ν 2.�������������������������������������������������������������������������������� (1.7)
 gC  gC d 2 gC

Note that ∆pL is the difference between the upstream and the downstream pressure and ∆pPE is
the pressure drop because of potential-energy change. It is the hydrostatic head of the fluid and
accounts for the pressure change caused by the fluid column weight. Also note that r is the fluid
density and q is the pipe deviation from horizontal. Thus, the hydrostatic head for a horizontal pipe
(q = 0°) is zero. For a vertical tubing, q = 90° for upward flow and q = –90° for downward flow. For
fresh water, the potential-energy pressure drop per foot of vertical distance is 0.433 psi/ft. ∆pF is the
pressure drop resulting from pipe friction and is obtained from the Fanning equation (John Thomas
Fanning, 1837–1911):

2 f f ρν 2
∆pF = ,�������������������������������������������������������������������������������������������������������������������������� (1.8)
gc d

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 10 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  11

where v is the fluid velocity and d is the pipe diameter. For a pipe with a constant cross section, the
fluid velocity can be expressed in terms of flow rate, q, as

4q
ν= .������������������������������������������������������������������������������������������������������������������������������������ (1.9)
π d2

The Fanning friction factor, ff, for laminar flow is

16 4 π dµ
ff = = ,������������������������������������������������������������������������������������������������������������������ (1.10)
N Re qρ

where the Reynolds number of the flow is

4 qρ qd
N Re = = . ������������������������������������������������������������������������������������������������������������������ (1.11)
π dµ ν A

Substituting v and ff in the expression for ∆pF yields the Hagen-Poiseuille law for pressure drop in a
pipe (Hagen 1839; Poiseuille 1840):

128µ ql
∆p f = .������������������������������������������������������������������������������������������������������������������������ (1.12)
π gc d 4

From the foregoing definitions and equations, we can see that for a given flow rate, the pressure
drop is proportional to 1/d4 for laminar flow, and for turbulent flow in a smooth pipe, the pressure
drop is proportional to 1/d4.75. Hence, any reduction in effective pipe diameter from deposit buildup
can lead to a dramatic increase in pressure drop and hence to a similar decrease in the production
from the reservoir. For turbulent flow in a smooth pipe, ff = 0.079/NRe0.25.
In Eq. 1.6, ∆pKE is the pressure drop caused by change in kinetic energy between various positions
in the pipe. Generally, it is much smaller compared to ∆pPE and ∆pF. It is equal to zero if there is no
change in fluid velocity between the two points of measurement. For example, given an incompress-
ible fluid f flowing through a pipe of uniform cross-sectional area, the velocity does not change and
∆pKE is equal to zero. In cases of high gas volumes or high gas/oil ratios, a rapid change in velocity
may occur, but even then ∆pKE generally accounts for less than 10% of the pressure loss.
Elements in Eqs. 1.6 through 1.12 have the following values:

g = gravitational acceleration, m/s2,


gc = gravitational acceleration at the center of mass, m/s2,
ff = Fanning friction factor, dimensionless,
q = flow rate, m3/s,
l = length of pipe, m,
A = area of pipe, m2,
d = pipe inside diameter (for round pipe), m,
v = flow velocity, m/s,
r = density, kg/m2,
m = dynamic viscosity, kg/(m·s),
n = kinematic viscosity (m/r), m2/s.

Note that in the various segments of the pipeline industry, expressions used in pipeline models
differ; however, NRe is universally important because the transitions from laminar to turbulent flow
is generally described by this number with laminar flow conditions <2,000, turbulent flow >4,000,

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 11 03/11/17 6:50 PM


12  Chemical and Mechanical Methods for Pipeline Integrity

and a transition range between these values. The


Laminar Darcy friction factor (fD), also called the Darcy-
flow Weisbach friction factor, is used in some applica-
tions and equals 4(ff) or 64/NRe.
Clean pipe
In addition to velocity, pipe diameter, and vis-
cosity, other factors can affect the flow regime.
Shannon (2010) claims that the presence of debris
Slightly or scale in a pipe can change the flow patterns.
turbulent
flow The preceding equations assume a smooth pipe,
while Fig. 1.6 shows three possible conditions for
the same velocity and viscosities.
Dirty pipe
Thus, in this set of conditions, there is laminar
flow for a smooth pipe, slightly turbulent flow for
a “dirty” pipe, and fully turbulent flow when the
Turbulent diameter has constricted enough to cause this flow
flow
condition. The friction factor equation (Eq. 1.10)
will also change for a dirty, rough pipe. An equa-
Chocked-down pipe
tion to calculate the change is
Fig. 1.6—Effects of deposits on flow.

1  ε 2.51 
= −2 log10  +  .�������������������������������������������������������������������������������������������� (1.13)
f  3.7d N Re f 

In Eq. 1.13, e is the height of the roughness (mm). Although this can be measured for laboratory
measurements, it is usually calculated from the observed pressure drop in a line (see Eq. 1.14).
Multiphase Flow. Understanding the multiphase flow conditions in the lines is an important
aspect in predicting corrosion and solids accumulations, as well as distribution of inhibitors in pipe-
lines. Shoham (2006) stated that the hydrodynamics of single-phase flow is well-understood and
that pressure drop vs. flow rate and heat transfer can be calculated straightforwardly (see flow rate
equations in the previous subsection, Eqs. 1.6 through 1.13). However, adding a second (or third
phase) greatly complicates the analyses. This short section provides an introduction to the different
flow regimes found in piping and modeling methods to predict hydrodynamic problems. Shoham
(2006) and Asante (2002) are recommended for many more details of the methods for calculating
important variables such as the friction factor and the pressure drop during multiphase flow condi-
tions. Different multiphase flow combinations in piping include

• Hydrocarbon liquid/water
• Hydrocarbon gases/hydrocarbon liquids
• Hydrocarbon liquid/water/gases
• Combinations of phases

The gas phases may also include acid gases such as acetic acid (HAc), hydrogen sulfide (H2S),
and CO2, and temporally dispersed phases of water and liquid hydrocarbons will form. Because of
the density and viscosity differences between the different phases, they will have different mass and
volume velocities, and these differences cause the complex regimes to form.
Fig. 1.7, adapted from Shoham (2006) and Kee et al. (2015), depicts several conditions that may
occur given the rate of flow in pipelines experiencing horizontal multiphase conditions. This figure
shows water and oil/gas phases. As the flow rates increase, various different regimes form.
Some possible regimes from the papers by Shoham (2006) and Kee et al. (2015) are described
below with letters referencing Fig. 1.7. Note that the Fig. 1.7 diagram shows the continuous phases
as well as bubbles and droplets of liquids temporally dispersed. These conditions can affect corrosion

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 12 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  13

as well as the application of inhibitors. Two-phase flow follows a similar sequence, but without the
second liquid phase.

Flow direction

Key
Gas a) Stratified (ST)

Water

Oil
b) Elongated
Gas bubble bubble (EB)

Water droplet

Oil droplet
c) Slug (SL)

Increasing
flow velocities d) Wavy
annular (WA)

e) Annular mist
(AM)

Fig. 1.7—Horizontal multiphase (three-phase) flows in gas/liquid pipelines (after Shohan


2006 and Kee et al. 2015).

Stratified Flow (Fig. 1.7a). Stratified flow is characterized by the concurrent flow of liquid streams
at the bottom and a gas stream at the top of the pipe. The two liquid phases are often separated or
slightly dispersed at the oil/water interface. The gas/liquid interface may be smooth or show some
waviness caused by the drag of the gas passing over the liquid.
Elongated Bubble Flow (Fig. 1.7b). Elongated bubble flow is also called plug flow and is a form
of intermittent flow that occurs at low gas velocity.
Slug Flow (Fig. 1.7c). Slug flow occurs when the liquid bridges the entire pipe cross section, form-
ing a liquid slug, while the gas flows as a large bubble between the trains of liquid slugs. The large
gas bubble moves on top of a slower-moving stratified liquid layer characterized as the gas-bubble/
liquid-film zone. Smaller bubbles of gas as well as droplets of oil and water are dispersed in the vari-
ous phases.
Wavy Annular Flow (Fig. 1.7d). Wavy annular flow occurs at the transition between slug and
annular flow. The flow lacks the characteristic pressure fluctuation found in slug flow. The upper
wall is occasionally wetted by an unstable liquid film that keeps falling diagonally downward.
Annular Mist Flow (Fig. 1.7e). Annular mist flow occurs at very high gas velocity when gas flows
at the pipe core and liquid moves as an annular film enveloping the pipe wall. The turbulent gas
contributes to rough gas/liquid interfaces containing interfacial waves of varying amplitudes.
When liquid-containing lines, such as the current Alaskan pipeline (Abrams 2011), are not operat-
ing at capacity, the lines will not be full and three-phase flow may exist. Some of these lines could
contain solid particles as a result of changing conditions or the introduction of contaminants.
The flow regimes described in Fig. 1.7 may be present when lines are horizontal. Kesana (2013),
Thome (2012), and Brill (1987) describe these flow regimes. These authors describe plug flow as liquid

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 13 03/11/17 6:50 PM


14  Chemical and Mechanical Methods for Pipeline Integrity

a) Effects of terrain plugs separated by elongated


gas bubbles that are smaller
than the tube diameter.
Gas Gas Gas Because the bubbles are
more buoyant than the liquid,
Gas Gas there is a continuous film of
Condensate Condensate liquid in the tube bottom. As
H 2O H2O a line is inclined, changes
may occur as a result of flow
rate changes.
Fig. 1.8a depicts a flow
regime of liquids and a
b) Types of flow patterns in vertical pipes gas and shows the effects
of terrain. Here, the hills
and valleys may cause
liquids to remain in the
line bottom, especially at
low flow rates. This may
lead to corrosion in the
low points or to additional
backpressure on the line
caused by holdup of the
liquids. Fig 1.8b shows
flow regimes in verti-
cal pipes (see Brill 1987;
Thome 2012). In these
vertical orientations the
regimes can include bub-
ble flow, slug flow, churn
flow (a form of intermit-
tent flow in which peri-
Bubble Slug Churn Annular odical flow reversal of
flow flow flow flow
liquid film is observed),
and annular mist, depend-
Fig. 1.8—Effects of terrain (a) and vertical flow regimes (b) (Brill 1987
and Thome 2012).
ing on the flow velocities.
Kesana (2013) claimed
that slug flow, especially if it is carrying solids, can result in severe metal loss by erosion
(erosion and erosion/corrosion of metals are described in Section 3.2.3).
Wang et al. (2013) provided additional information on the variety of regimes that occur in mul-
tiphase flow in the piping. This group studied a range of flow conditions and pressure drops in a
loop described in this report. They also used a lower °API value (28.5 °API) and higher-viscosity
(dead-oil viscosity is 1.1 Pa·s at 15.6°C) oil phase than is usual (see Pan et al. 1995). In addition,
these investigators used natural gas (instead of a more inert gas) at a range of elevated pressures as
well as changing oil flow, gas flow, and water flow velocities. They also used a sapphire window
to observe the phases. A series of flow regimes were identified and are depicted in Fig. 1.9. Types
of flow in which the oil is dispersed in the water are depicted in Figs. 1.9a and 1.9b. Conditions of
water dispersed in oil are shown in Fig. 1.9c. In all cases, the tests had horizontal flows under inter-
mittent and slug flow conditions.
Modeling Multiphase Flow Conditions. Griffith (1984) reviewed qualitative aspects of pre-
dicting the pressure drop in pipelines with multiphase flow. He claimed that the most important

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 14 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  15

factor is the friction factor (see Eq. 1.10), but he recommended using Moody curves (Moody 1944),
assuming that the fluid is flowing at the velocity of the mixture. A Moody diagram showing the
Darcy friction factor (fD) plotted against Reynolds number for various roughness factors is shown as
Fig. 1.10 (from Wikipedia 2013b).

 d 2 
fD =    2  ∆p .���������������������������������������������������������������������������������������������������������������� (1.14)
 L   ρq 

Asante (2002) described some of the models used to calculate the consequences of multiphase
flow in gas pipelines. They parallel the regimes seen in in Fig. 1.7. Computations using single-phase
and multiphase models are in use to estimate important flow properties. Shoham (2006) explained
that the modeling methods can be divided into “empirical-physical models,” computational models
(including computational fluid dynamics), and mixed models. The methods describe next have used
experiments as well as some level of computational methods to describe the flow regimes occurring
at different points in a pipeline.
Single-Phase Approaches for Multiphase Flow Models. These models use calculation of
observed friction factors ( f in the Moody diagram shown in Fig. 1.10) and pressure drop (∆P) for a
tested fluid mixture and then use the single-phase equations (Eqs. 1.6. through 1.11) for predictions.

(a)
Gas phase with
thick oil film on
Oil and gas the wall INT(O/W-S&SOW-F)
dispersed Horizontal intermittent flow
in water Oil layer with O in W dispersion slug and
stratified oil and water film

Water layer

Slug body region Liquid film region

(b)
Gas phase with
thick oil film on INT(O/W-S&O/W-F)
Oil and gas the wall
dispersed Horizontal intermittent flow with
in water O in W dispersion slug and
oil and water dispersion film

Oil dispersed in
water layer

Slug body region Liquid film region

(c) Gas phase with


thick oil film on
the wall
INT(W/O-S&W/O-F)
Water and
Horizontal intermittent flow
gas dispersed
with W in O dispersion slug and
in oil
water and oil dispersion film
Water
dispersed in oil
layer

Slug body region Liquid film region

Fig. 1.9—Various flow patterns in three-phase flow with heavy oil and natural gas (Wang et al. 2013).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 15 03/11/17 6:50 PM


16  Chemical and Mechanical Methods for Pipeline Integrity

Moody Diagram
0.1
0.09
0.08 Transition region
0.07 0.05
0.04
0.06 0.03
0.05 0.02

Relative Pipe Roughness, ε/D


0.015
0.04
0.01
Friction Factor, f

0.03 0.005

Laminar flow
64 0.002
0.02 Re
0.001
Material ε(mm) 5×10–4
0.015 Concrete, course 0.25
Concrete, smooth 0.025 2×10–4
Drawn tubing 0.0025
Glass, plastic 0.0025 Complete turbulence 10–4
0.01
Iron, cast 0.15 5×10–5
Steel, mortar lined 0.1
Steel, rusted 0.5
Steel, forged 0.025 10–5
Sewers, old 3.0 5×10–6
Water mains, old 1.0 Smooth pipe
10–6
103 104 105 106 107 108
Reynolds Number, Re

Fig. 1.10—Moody diagram.

Shoham (2006) considered this to be part of the “experimental” approach but noted that it is appli-
cable only under the conditions that are close to those of the experiment.
Multiphase Approaches for Multiphase Flow Models. Asante (2002) claimed that the simplest
models, termed “homogeneous models,” would apply to the mist flow in Fig. 1.7 and use modifica-
tions of the single-phase equations (Eqs. 1.6 through 1.11) and adjust the friction factor ( f ), the
density (r), and the viscosity (m).
An example of this model is in Fancher and Brown (1963), who discussed homogeneous calcula-
tions using the correlation of Poettmann and Carpenter (1952):

dp 1   q2 M 2 
= ∆p = − ρ + f  5 
.���������������������������������������������������������������������� (1.15)
dl 144   7.413 × 10 ρ d  
10

In this expression, p = pressure, psi; L = length, ft; r = flowing density, lbm/ft3; f = Fanning fric-
tion factor; d = pipe diameter, ft; q = oil flow rate, B/D; and M = total mass of gas and liquid, lbm.
Fancher and Brown (1963) noted that as the gas/liquid ratio increases and the liquid rate decreases,
and pressure gradients calculated from Eq. 1.15 are influenced more by the second term of this equa-
tion than by the first.
Stratified flow and annular mist flow conditions (Figs. 1.7, 1.8, and 1.9) are frequently encoun-
tered and are relevant to many FA and IM conditions. They are highly complex, and so the reader
is directed to Asante (2002) and Shoham (2006). Wang et al. (2013) describe multiphase flow tests
and correlations with the multiphase models of Zhang and Sarica (2006), and the authors found that
the water flows show reasonable predictions, whereas the pressure drops and liquid holdups were
underpredicted.
A three-phase flow model based on experimental as well as computational data was described
by Karami et al. (2016). In the nomenclature of Shoham (2006), this would be a mixed model.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 16 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  17

To simulate oil, water, and gas flowing conditions, an experimental setup—a 6-in.-ID pipe flow loop
(see Fig. 1 in Karami et al. 2016)—used mineral oil, tap water, and air to simulate the gas phase.
The experiments were conducted under low-liquid-loading condition, which is commonly observed
in wet-gas pipelines. The analyzed flow characteristics included wave pattern, liquid holdup, water
holdup, pressure gradient, and wetted-wall fraction.
Videos of the wave patterns of the liquid phases were also recorded for analyses. The observed
wave patterns included stratified smooth and stratified wavy with 2D waves, 3D waves, roll waves,
and atomization flow. The transitions between the flow patterns vary as a function of water cut. The
trends of pressure gradient, liquid holdup, and water holdup with respect to gas velocity (vSg) and
liquid velocity (vSL), as well as water cut, were observed. Good correlation with several mathemati-
cal models was demonstrated.
Details of additional flow models that are in use are described in NACE SP0208-2008 (2008),
Appendices A and B. Because corrosion, scale, and use of control chemicals depend on the phase
that coats the surfaces, the complex models are critical both to prediction and to pipeline treatments.
Hilgefort (2014) as well as NACESP 0208-2008 (2008) provided information on multiphase flow
models and how the various conditions affect corrosion in pipeline segments (many more details
of corrosion and its manifestations are provided in Chapter 3). Hilgefort (2014) stated that flow
models help predict important conditions that may cause corrosion. Thus, a first step is determining
the potential for accumulation of water and debris if stratified flow exists, given that internal corro-
sion occurs where water (or hygroscopic solids) comes into contact with the pipe wall. Therefore,
if it is known which pipelines are likely to experience water or solids accumulation, the operator
knows which are susceptible to internal corrosion. If the operators know where, along the pipeline,
the water and debris accumulation will occur, they can perform inspections to check for internal
corrosion.
The most important factors in the models are the (1) critical water velocities and inclination
angles for water and/or solids and (2) critical inclination angles (q ), which are

 ∆(elevation) 
θ = arcsin  .������������������������������������������������������������������������������������������������������ (1.16)
 ∆(distance) 
These critical factors are then compared to the elevation and inclination profile of the pipeline (like
those in Fig. 1.7d) to find where water or solids may accumulate.
Solids moving in a pipeline, Hilgefort (2014) noted, are subject to gravitational forces, which
deposit the particles, and also to turbulent forces, which keep the particles in suspension.
At lower flow rates, particles tend to settle out and can sit at the pipeline bottom.

1.4.3  Viscosity and Rheology of Fluids. The viscosity of a fluid is an essential property for char-
acterization of flow behavior. This property is a measure of a fluid’s resistance to being deformed
by either shear stress or tensile stress. Viscosity describes a fluid’s internal resistance to flow and
may be thought of as a measure of fluid friction. Wikipedia (2009) claims that James Clerk Max-
well (1831–1879) called viscosity “fugitive elasticity” because of the analogy that elastic defor-
mation opposes shear stress in solids, whereas in viscous fluids, shear stress is opposed by rate of
deformation.
Viscosity is a primary determinent of the energy needed to move fluids in pipes. The Hagen- Poi-
seuille equation (Eq. 1.12) notes that pressure drop is directly proportional to viscosity, and there-
fore more pumps or larger ones will be required to move a more viscous fluid compared with a less
viscous material. Thus, knowledge of the viscosity as a function of the piping and temperatures is
critical information for all pipeline operations.
Viscosity is a function of the fluid’s chemical composition and physical state (usually tempera-
ture). Two different definitions of viscosity are in common use: absolute viscosity and kinematic

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 17 03/11/17 6:50 PM


18  Chemical and Mechanical Methods for Pipeline Integrity

viscosity. Absolute (intrinsic) viscosity (h) is a measure of the resistance to flow that a fluid offers
when it is subjected to shear stress, shown as Eq. 1.17. This equation relates the shear stress (s)
exerted on a fluid to the resultant strain rate (g ). Kinematic viscosity (υ) is defined as the ratio of
absolute viscosity to the density of the fluid at the same temperature. The apparent (measured)
viscosity (µa) is the value usually reported in oilfield application and may depend on the shear rate.

σ = ηγ.������������������������������������������������������������������������������������������������������������������������������������ (1.17)

The viscosity of the crude oil and the viscosities of various treating fluids described in this book
are critical values. In general, the viscosity of crude oil increases with its density—that is, the lower
the °API value of a crude oil (denser oil), the higher the viscosity. Typically, the viscosity of dead oil
is experimentally determined as a function of temperature, and the viscosity of live oil is determined
as a function of pressure at reservoir temperature (for reservoir engineering purposes) and at a lower
temperature (for facility design purposes). In addition, the viscosity of various petroleum fractions
is important in downstream applications.
The viscosity of petroleum fractions also increases with a decrease in the °API value; for
residues and heavy oils with °API value of less than 10 (specific gravity above unity), the vis-
cosity varies from several thousands to several million poises. Viscosity is a bulk property that
can be measured for all types of petroleum fractions in liquid form. Kinematic viscosity is a
useful characterization parameter for heavy fractions in which boiling point data are not avail-
able because of thermal decomposition during distillation. Viscosity is not only an important
physical property but also a parameter that can be used to estimate other physical properties
as well as composition and quality of undefined petroleum fractions (Riazi 2005). Generally,
the kinematic viscosities of petroleum fractions are measured at standard temperatures 37.8°C
(100°F) and 98.9°C (210°F).
As Section 4.6.1 will demonstrate, when the fluid being transported contains several phases, emul-
sions or dispersions may form. Kalra et al. (2012) showed that emulsions will generally increase
fluid viscosity, and this will require the resizing of pumps and other facilities to accommodate the
needed flow rates.
The viscosity of various treating fluids is also of great importance because it affects the chemical/
physical properties of fluids. The measurement of viscosity is accomplished by several techniques.
In general, either the fluid remains stationary and an object moves through it, or the object is station-
ary and the fluid moves past it. The drag caused by the relative motion of the fluid and a surface is a
measure of the viscosity. The flow conditions must also have a sufficiently small value of the Reyn-
olds (Re) number for there to be laminar flow (see the discussions of flow equations in Section 1.5).
For Newtonian liquids (i.e., viscosity that is independent of shear rate), viscosity can be measured
by capillary U-tube viscometers (Fig. 1.11). This devices measures the time it takes for the test liq-
uid to flow through a capillary of a known diameter of a certain factor between two marked points.
By multiplying the time taken for the fluid to flow times the factor of the viscometer, one obtains
the kinematic viscosity. Viscometers are usually placed in a constant-temperature water bath for this
measurement, and kinematic viscosity is measured at temperatures from 15 to 100°C (≈60–210°F).
The test method is described in more detail in ASTM D445-04e1 (2004).
In this method, repeatability and reproducibility are 0.35 and 0.7%, respectively (Denis
and Briant (1997). A large variety of “complex fluids” are encountered in oilfield treat-
ments, especially in reactive stimulation, hydraulic fracturing, and EOR. These are fluids
that behave neither like a liquid nor like a solid under flow, but show a mixed behavior. The
study of these types of fluids is “rheology.” The simplest definition of rheology is the study
of the flow of matter. For simple fluids such as water or many organic liquids, including some
very viscous fluids, the relationship of the shear stress ( s ) applied to the resultant shear strain
rate ( g ) is linear and the ratio of these quantities is the liquid viscosity ( u or h ) as defined

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 18 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  19

U-tube (Ostwald) Viscometer Rotational (Couette) Viscometer

Fig. 1.11—Viscometer types (Frenier et al. 2010).

in Eq. 1.17. These fluids are


termed “Newtonian fluids.” Shear thinning
Fig. 1.12 shows types of Bingham plastic
stress/strain responses for
Newtonian
several different types of
fluids that include Newto-
nian and non-Newtonian
materials.
Shearing Stress, t

Note that a Bingham plas-


tic has a yield value and does
not respond until that point;
however, it then acts as a
µ
Newtonian fluid. For many
types of complex fluids such 1
as emulsions, suspensions,
slurries, gels, foams, polymer
solutions, and other complex Shear thickening
mixtures of substances, the
shear-stress/shear-rate rela-
tionship cannot be charac-
terized by a single value of
viscosity (u) (i.e., at a fixed Rate of Shearing Strain, dµ/dy
temperature). Instead, the
viscosity is a function of the Fig. 1.12—Types of viscous responses.
operating conditions, such as
the shear rate. One task of
rheology is to establish the relationships between viscosity and stresses by adequate measurements
and modeling. Two examples of the non-Newtonian behavior are shear thinning and shear thicken-
ing fluids. Thixotropic fluids are even more complex because a scan of stress vs. strain does not have
the same values when the strain is reversed. Latex paints provide an example of a thixotropic fluid.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 19 03/11/17 6:50 PM


20  Chemical and Mechanical Methods for Pipeline Integrity

100 For polymer melts (and wax/oil


mixtures) and suspensions, gener-
60°F
80°F
ally, the viscosity decreases as the
100°F
125°F
shear rate increases. This type of
150°F
Viscosity (Pa·s)

behavior, shear thinning, is of con-


–1 siderable industrial significance.
10
For example, fracturing fluids are
Newtonian
region shear thinning. Some fluids exhibit
different types of behavior depend-
Power law ing on the shear field. Carbohy-
region
drate-based fracturing fluids [e.g.,
10–2 hydroxypropyl guar (HPG)] exhibit
10–2 10–1 100 101 102 103 104 such complex behavior when in
Shear rate (s–1) water. Fig. 1.13 shows plots of
0.48% HPG in water as a function
Fig. 1.13—Hydroxypropyl guar (HPG) rheology as a function of
of shear rate and temperature (Guil-
temperature (Constein et al. 2001).
lot and Dunand 1985). Two models
will be required to fit this type of data. Eq. 1.17 is useful in the linear (Newtonian) low-shear region,
but a power-law model gives an apparent viscosity (ma):

µa = K / γ (1− n ) .������������������������������������������������������������������������������������������������������������������������ (1.18)

Eq. 1.18 is required for the high-shear region. Here, K is the consistency index in lbf-sn/ft2 or
kPa·sn, and n is the flow behavior index (dimensionless). These relationships hold for most fractur-
ing fluids over the range of shear rates in which the fluid displays non-Newtonian behavior. A log-
log plot of s vs. g usually yields a straight line over a portion of the shear range. The slope of the
straight-line portion is equal to the behavior index n, and the value of t at g = 1.0 s–1 is equal to the
consistency index K. A log-log plot of ma vs. g has a straight-line slope of n – 1 when the power-law
model is applicable. The slope is zero for Newtonian behavior (Constein et al. 2001). To better pre-
dict the full range of fracturing fluid viscosity, a rheology model must use not only n and K but also
a zero-shear viscosity term. The Ellis model (Matsuhisa and Bird 1965) added zero-shear viscosity
at γo to the power-law model to improve viscosity prediction:

1 1 1
= + .�������������������������������������������������������������������������������������������������������������������� (1.19)
µ a µ o Kγ n −1

Here, n and K are defined from the high-shear data.


Shear thickening is a less frequently observed phenomenon whereby the material exhibits an
increasing viscosity with increasing shear rate. A property of some complex fluids (e.g., some types
of clay in water and latex paint, as noted earlier), thixotropy refers to time-dependent viscosity.
A thixotropic fluid displays a decreasing viscosity with time at a constant shear rate. Materials that
exhibit the opposite behavior— that is, increasing viscosity with time at a constant shear rate—are
rheopectic. A complex fluid may also exhibit a yield stress. Below a certain value of applied stress,
the yield stress, the fluid does not flow; but above this stress, the fluid flows. It may be noted that
waxy crude oils, for example, are usually both shear thinning and thixotropic, and they may exhibit
a yield stress at lower temperatures as well.
A rheometer is an instrument that is used to measure the rheology of fluids. In addition to applying
a constant shear stress or a constant shear rate, rheometers can also apply oscillatory motion. This is
called “dynamic oscillatory rheometry.” The response to the oscillatory motion can be used to deter-
mine the solid-like behavior and the liquid-like behavior of the fluid. Further, the stress in the direction

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 20 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  21

normal to the flow can Elastic solid


also be measured. Thus, Strain
the stress tensor can be Stress
related to the strain. The δ=0
measured response can Measured stress response
Viscous fluid
be factored into two com-
Strain
ponents: the in-phase and
Stress
the out-of-phase. The in-
phase response defines Viscoelastic fluid δ = π/2
the storage modulus, G′,
Strain
which gives informa- Applied strain deformation
tion on the elasticity; Stress
and the 90° out-of-phase
response, G″, is the loss 0 < δ < π/2
modulus that gives infor-
Fig. 1.14—Dynamic oscillation rheometry.
mation about the viscous
properties of the fluid. To
perform the experiment, a sinusoidal stress is applied at a frequency, v, and the strain (in phase and
out of phase) is plotted vs. the frequency (see Fig. 1.14).
Rheometers can be programmed to run temperature sweeps, shear sweeps, and creep tests to
determine various rheological properties of fluids. A study of this behavior can provide mechanistic
information about the arrangement and rate of change of the polymer molecules.
Another type of viscometer is a rotary viscometer, which is used for a wide range of shear rates,
especially for low shear rate and viscous fluids such as lubricants and heavy-petroleum fractions and
fracturing fluids. In these viscometers, fluid is placed between two surfaces: One is fixed, and the
other is rotating. Rotational viscometers use the idea that the torque required to turn an object in a
fluid is a function of the viscosity of that fluid. They measure the torque required to rotate a disc or
bob in a fluid at a known speed. “Cup” and “bob” viscometers work by defining the exact volume of a
sample that is to be sheared within a test cell; the torque required to achieve a certain rotational speed
is measured and plotted. There are two classical geometries in cup and bob viscometers, known as
either the “Couette” or the “Searle” system, distinguished by whether the cup or the bob rotates. The
rotating cup is preferred in some cases because it reduces the onset of Taylor vortices (i.e., axisym-
metric toroidal vortices), but it is more difficult to measure accurately. See Fig. 1.11 for diagrams of
the two types of viscometers. Also see Wikipedia (2009) for examples of other types of viscometers.
The above methods are used to measure viscosity of liquids at atmospheric pressure. To measure
the viscosity of the live fluid, the viscometers must operate under pressure. Typically, three types of
viscometers are used for these measurements: a rolling ball viscometer, a capillary flow viscometer,
and an electromagnetic viscometer. The reservoir fluid sample is transferred to the viscometer at an
elevated pressure to ensure monophasic transfer.
In the rolling ball viscometer, a ball is allowed to fall through the fluid, and the time required for
the ball to travel through the fluid is correlated to the fluid viscosity. In the capillary flow viscom-
eter, the live fluid is allowed to flow under pressure through a long capillary tube. The pressure drop
across the tube is measured, and the viscosity of the fluid is calculated on the basis of the Hagen-
Poiseuille equation for laminar flow (Eq. 1.12). Another device is an electromagnetic viscometer
(CVI 2011). In this method, a piston moves back and forth through the pressurized fluid, and the
drag on the piston is measured, thus inferring the viscosity.

1.4.4  Thermodynamics and Kinetics of Pipeline-Fouling Reactions. The concepts of aqueous


solution thermodynamics are crucial for understanding many processes in oilfield chemistry. The
topic is covered in great depth in textbooks on aqueous chemistry. For example, see Zemaitis et al.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 21 03/11/17 6:50 PM


22  Chemical and Mechanical Methods for Pipeline Integrity

(1986) and Langmuir (1997). Therefore,


only a simplified description of important
principles is presented here. The reader
Rate of Reaction

Forward reaction should consult the above-mentioned texts


r1 = r2
for a more in-depth discussion.
A system at chemical equilibrium rep-
resents a dynamic state in which two or
Reverse reaction more opposing reactions are taking place
at the same time and at the same rate. Fig.
1.15 shows the approach of a chemical
system to equilibrium. At early time, the
Time rate of forward reaction (r1)—that is, the
Fig. 1.15—A chemical system’s approach to equilibrium. formation of products from reactants—is
large. As the system progresses in time,
the rate of forward reaction decreases and the rate of reverse reaction (r2) increase. When the
rates are equal, the system is in equilibrium and the net rate (i.e., rate of formation of products
minus rate of formation of reactants) is equal to zero. Once a closed system reaches chemical
equilibrium, the chemical composition of the system is, by implication from that point on, inde-
pendent of time and previous history.
Equilibrium constants are used to relate the amounts of reactants and products at equilibrium.
Consider, for example, a simple ideal chemical system consisting of reactants A and B, and products
C and D, for which the reaction stoichiometry is given by

aA + bB ⇔ cC + dD .�������������������������������������������������������������������������������������������(1.20a)

This would imply that the forward and reverse reactions are

aA + bB → cC + dD (forward reaction)������������������������������������������������������������ (1.20b)

and

cC + dD → aA + bB (reverse reaction) .�������������������������������������������������������������� (1.20c)

If the system is ideal, then the equilibrium constant for the system can be expressed in terms of
concentrations as

[C ]c [ D]d
K eq = ,���������������������������������������������������������������������������������������������������������������������� (1.21)
[ A]a [ B]b

where Keq is the equilibrium constant and [ ] denotes concentration of the species. The equilibrium
constant can also be expressed in terms of the free-energy change for the reaction (DG°) as

 −∆G  
K eq = exp  .���������������������������������������������������������������������������������������������������������������� (1.22)
 RT 

Here, R is the gas constant and T is the temperature. The Keq is a function of temperature and
pressure only and does not depend on composition of the system. Therefore, if the composition of
the system changes as a result of subsequent reactions or addition of new reactants or products to

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 22 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  23

the system, the same value of the equilibrium constant(s) can be used to calculate the equilibrium
distribution in the new system provided that the temperature and pressure remain constant and the
equilibrium assumption is valid.
Equilibrium constants are not constants in the true sense because they do depend on temperature
and pressure. An increase in temperature may affect the forward and reverse reactions differently.
The reaction that absorbs the most heat will increase that rate to a larger extent than the other reac-
tion. A new equilibrium constant will now represent the new situation. The simple equation by van’t
Hoff (J. H. van’t Hoff, 1852–1911) can be used to compute the change in equilibrium constant
caused by a change in temperature. The equation can be expressed as

d ln K eq ∆H °
= ,���������������������������������������������������������������������������������������������������������������������� (1.23)
dT RT 2

where DH° is the standard enthalpy change of reaction, T is the temperature, and R is the gas con-
stant. If the reaction is exothermic (i.e., if DH° for the reaction is negative), then the equilibrium
constant decreases as temperature increases. Conversely, Keq increases with temperature for an endo-
thermic reaction.
If the standard enthalpy change of reaction is assumed independent of temperature, then integrat-
ing Eq. 1.23 gives an even simpler result:

K eq  
ln = − ∆H °  1 − 1  .�������������������������������������������������������������������������������������������������� (1.24)
K eq,ref R  T Tref 

Here, Keq,ref is the value of the equilibrium constant at the reference temperature Teq,ref. This approxi-
mate equation implies that a plot of ln Keq vs. the reciprocal temperature gives a straight line. This
equation is helpful in interpolating and extrapolating equilibrium constant data with reasonable
accuracy.
The effect of pressure on equilibrium constants is typically smaller than the effect of temperature.
However, for deep wells it can be significant and needs to be considered along with the change in
temperature. The pressure dependence of the equilibrium constant can be calculated from

 ∂ln K eq  ∆V °
 ∂ P  = RT , ���������������������������������������������������������������������������������������������������������������� (1.25)
T

where ∆V° is the molar volume change of the reaction with all reactants and products in their stan-
dard states (Langmuir 1997). If the molar volume change of the reaction is independent of pressure,
then the integration of Eq. 1.25 yields

K eq ∆V °( P − Pref )
ln =− ,������������������������������������������������������������������������������������������������������ (1.26)
K eq,ref RT

where Keq is the equilibrium constant at the desired pressure P. Keq,ref is the equilibrium constant at
Pref, which is typically 1 bar.
In very dilute aqueous solutions, the anions and cations behave in an “ideal” manner, in which
each ion will act as if it is independent of all other ions in the solution. In real solutions, especially
those in which there are high concentrations of other ions (such as in a produced brine), the ions
are affected by the other ions in solution, and so the fluid may not behave as if it has exactly the
same number of ions as described by the concentration. The equilibrium constants in such noni-
deal systems are then expressed in terms of species activity. If, for example, the chemical system

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 23 03/11/17 6:50 PM


24  Chemical and Mechanical Methods for Pipeline Integrity

considered previously of components A, B, C, and D is nonideal, then the equilibrium constant


is given by

{C }c { D}d  −∆G  
K eq = = exp  , ������������������������������������������������������������������������������������������ (1.27)
a
{ A} { B} b
 RT 

where { } denotes the activity of the species. The activity of the species can be thought of as an
effective concentration of the species in solution. Programs to calculate or estimate the activity coef-
ficients are beyond the scope of this discussion, but there are several general purpose geochemical
models available in the public domain that can be used to predict formation of oilfield scale. Most of
them are available at no or minimal charge through the Internet and have been extensively reviewed
in texts on aqueous chemistry. They include Mangold and Tsang (1991); Glynn et al. (1992); Wolery
(1992); van der Heijde and Elnawawy (1993); Langmuir (1997); Butler and Cogley (1998).
These concepts can be applied directly to inorganic scales. Organic deposits such as wax (paraf-
fin), asphaltenes, gas hydrates, and naphthenates also become supersaturated before they can start
to deposit. However, because of system complexity, equilibrium constants cannot usually be calcu-
lated. See Section 4.2.1 for discussions of the rates of scale dissolution or deposition of solids and
how they are influenced by the thermodynamic properties described in this section.

1.4.5  Surface Chemistry. The chemistry and physics of surfaces affect a great number of processes
and services performed in oil/gas transportation and apply to solids as well as liquids. The surface of
any solid or liquid is an interface between that medium and some other that could be a solid, liquid,
or a gas. The chemistry of the interface is affected by both surfaces. In one example (Fig. 1.16), a
drop of oil is placed on a solid surface in a jar of water. The forces at that surface—that is, inter-
facial surface tension (IST, or g )—are caused by the cohesion of the molecules in that surface and
depend on the molecules in both surfaces. Therefore, IST is not a property of the liquid alone but of
the liquid’s interface with another material. In Fig. 1.16, the surface (on the left) is covered with a
water film (it is hydrophilic), and the oil drop is repelled and the contact angle (q ) is close to 0°. In
the middle case, the surface is partially oil-wetting, and in the right figure the surface is completely
oil-wet and the oil droplet spreads out on the surface.
If a liquid is in a container, such as a pipe or tank, then there will be a liquid/air interface at its top surface
and also an interface between the liquid and the container walls. The IST between the liquid and the air
is usually different from its IST with the container walls. Where two surfaces meet, the consequence
must also be such that all forces are in balance. As noted, where the two surfaces meet (and there are
three phases), they form a contact angle, q, which is the angle that the tangent line of the liquid surface
makes with the solid surface. Fig. 1.17 shows an example of solid, liquid, and gas interfaces. Here,
tension forces ( f ) are shown for the liquid/air interface, liquid/solid interface, and solid/air interface.

θ
γow
θ
γso γsw

θ ~ 0° γso = γsw + γow Cos θ θ ~ 180°

Fig. 1.16—Wetting angles of water, oil, and mineral surface (Abdallah et al. 2007).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 24 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  25

The vertical and horizontal forces f1a


θ
must both cancel exactly at the con- fsa
tact point. The horizontal compo-
fA
nent of fla is canceled by the adhesive
force, fA. f1s

fA = fla sin θ .�������������������������� (1.28)

The other balance of forces is in the


Fig. 1.17—Forces at surfaces
vertical direction. The vertical com-
(Wikipedia 2010d).
ponent of fla must exactly cancel the
force, fls.
Liquid Solid Contact Angle

fls = fsa + fla cos θ .���������������� (1.29) Water


Ethanol
Because the forces are in direct Soda-lime glass
Diethyl ether
Lead glass 180°
proportion to their respective surface Carbon tetrachloride
Fused quartz
tensions, we also have Glycerol
Acetic acid
γ ls = γ sa + γ la cos θ , ���������������� (1.30) Paraffin wax 73°
Water Silver 90°
where gls is the liquid/solid IST, gsa is the
solid/air IST, and gla is the liquid/air IST. Table 1.2—Contact angle values for solid/liquid converted to
Eq. 1.30 is known as Young’s equation. oilfield convention (Sears and Zemanski 1955).
Table 1.2 lists contact angle data
for common fluids, and Table 1.3 shows IST values for a number of liquids with air. The values of
the contact angles and the IST control many of the properties of processes described in subsequent
chapters. Note in Table 1.3 that the IST of pure water is much higher than that of solutions of ethanol
and water. Many surface-active agents (i.e., surfactants) also lower the IST of water and thus affect
the contact of water with surfaces. The convention used in the oil/gas industry is that a near-180°
contact angle indicates that the liquid will spread on (“wet”) that surface. The reader should note
which convention is being used when comparing data, given that some conventions call wetting to
have a 0° contact angle.
The contact angle and IST can be measured in several ways. The sessile-drop method is illustrated
in Fig. 1.18a. Here, a drop of liquid is seen after it has been placed on a surface in air, and the three
IST vectors are seen. The contact angle (q) can be measured visually by taking a photograph of the
surface, usually using a microscopic device; see Howard et al. (2010) for a specific example. Figs.
1.18a and 1.18c illustrate the importance of the solid surface itself. In these examples, a water drop has
been placed on two surfaces in an oil medium. When bronze is the solid (Fig. 1.18b), the water does
not displace the oil, whereas on glass (Fig. 1.18c), the water does displace the oil and wets the surface.
Fig. 1.18 also illustrates a possible caution when interpreting literature information, because
either the acute or the obtuse angle could be reported. If water is used to test the “wetting” of a
surface, low numbers refer to a hydrophilic surface and high values refer to a hydrophobic surface.
Also see Tadmor (2004).
The IST of a liquid vs. the IST of air can be measured using several methods, among them the Du
Nouy ring, the pendant-drop test, and the Wilhemly plate. These are explained next.
Du Nouy Ring. A manual method apparatus, the Du Nouy ring is depicted in Fig. 1.19. This is the
traditional method used to measure surface or interfacial tension. Wetting properties of the surface
or interface have little influence on this measuring technique. Maximum pull exerted on the ring

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 25 03/11/17 6:50 PM


26  Chemical and Mechanical Methods for Pipeline Integrity

Liquid Temperature (°C) IST, g (dyne/cm)


Acetic acid 20 27.6
Acetic acid (40.1%) + water 30 40.68
Acetic acid (10.0%) + water 30 54.56
Acetone 20 23.7
Diethyl ether 20 17.0
Ethanol 20 22.27
Ethanol (40%) + water 25 29.63
Ethanol (11.1%) + water 25 46.03
Glycerol 20 63
n-Hexane 20 18.4
Hydrochloric acid 17.7M aqueous solution 20 65.95
2-proponal 20 21.7
Methanol 20 22.6
n-Octane 20 21.8
Sodium chloride 6.0M aqueous solution 20 82.55
Sucrose (55%) + water 20 76.45
Water  0 75.64
Water 25 71.97
Water 50 97.91

Table 1.3—IST of various liquids measured against air (Dean 1961).

γla

θ
γsl γsa

(a) Illustration of sessile drop method with liquid


droplet partially wetting a solid substrate

(b) Water droplet immersed in oil (c) Water droplet immersed in oil
resting on a brass surface resting on a glass surface

Fig. 1.18—Sessile-drop method.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 26 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  27

Fig. 1.19—Du Nouy ring (Wikimedia 2016).

by the surface is measured (PHYWE 2004). The most impor-


tant experimental issues are using very clean equipment and fire
cleaning the platinum ring used in the test.
Pendant-Drop Test. Used with oils, the pendant-drop test
depends on surface tension to suspend a drop of liquid from the
end of a tube, as shown in Fig. 1.20.
The force from the surface tension is proportional to the
length of the liquid/tube boundary, with the proportional-
ity constant usually denoted as Given that the length of this d
Fg
boundary is the circumference of the tube, the force attribut-
able to surface tension is given by α
m
Fγ = π dγ .������������������������������������������������������������������������ (1.31)

Note that Here d is the tube diameter. The mass m of the drop Fig. 1.20—Surface tension ensured
hanging from the end of the tube can be found by equating the by pendant-drop method.
force caused by gravity,

Fg = mg.���������������������������������������������������������������������������������������������������������������������������������� (1.32)

The component of the surface tension in the vertical direction is Fγ = sin α . Thus,

mg = π dγ sin α . ���������������������������������������������������������������������������������������������������������������������� (1.33)

Here, a is the angle of contact with the tube, and g is the acceleration caused by gravity. The limit
of this formula, as a goes to 90°, gives the maximum weight of a pendant drop for a liquid with a
given surface tension, g. Thus,

mg
γ= .������������������������������������������������������������������������������������������������������������������������������������ (1.34)
πd

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 27 03/11/17 6:50 PM


28  Chemical and Mechanical Methods for Pipeline Integrity

Wilhemly Plate Method. Another method especially


F suited to check surface tension over long time intervals
is the Wilhemly plate method, described Fig. 1.21. A
vertical plate of known perimeter is attached to a bal-
ance, and the force resulting from wetting is measured.
This method uses the interaction of a test plate with the
liquid being tested. Biolinscientific (2009) described one
instrument that uses this method. The calculations for
air l this technique are based on the geometry of a fully wet-
ted plate in contact with, but not submerged in, the liquid
liquid phase. In this method, the position of the probe relative
to the surface is significant. As the surface is brought
θ into contact with the probe, the instrument will notice
this event through the change in forces it experiences. It
will register the height at which this occurs as the “zero
Fig. 1.21—Wilhemly plate. depth of immersion.” The plate will then be wetted to a
set depth to ensure that there is indeed complete wetting
of the plate (i.e., zero contact angle). When the plate is later returned to the zero depth of immersion,
the force it registers can be used to calculate surface tension (see Eq. 1.35).

F
γ la = ; l = 2w + 2d . ������������������������������������������������������������������������������������������������������ (1.35)
2lcosθ

A number of important properties and uses of pipeline chemicals are related to wetting and capil-
lary forces during pumping in the formation. Howard et al. (2010) explained that the Laplace and
Washburn equations can both be of importance (Grattoni et al. 1995). The Laplace equation relates
capillary pressure—the difference between the phase pressures of a nonwetting (P2) and a wetting
phase (P1) such as water and oil, Pc—to surface tension, g, and contact angle, q. Here, r is the radius
of a capillary tube, L is the height of the fluid rise in the tube, r is the density of the fluid, and g is
the gravitational acceleration (980 cm/sec2).

( P2 − P1 ) = Pc = L ρ g = 2γ cos θ / r �������������������������������������������������������������������������������������������� (1.36)

To evaluate wetting of two fluids in the same geometry, the ratio of two Laplace equations can be
used. For example, using water and a solution with unknown properties leads to

Lw ρw g = 2γ w cos θ w / r ������������������������������������������������������������������������������������������������������������ (1.37)

and

Lu ρ u g = 2γ u cos θ u / r . ������������������������������������������������������������������������������������������������������������ (1.38)

Dividing one by the other and removing constants gives

Lu γ u cos θ u
= . ���������������������������������������������������������������������������������������������������������������������� (1.39)
Lw γ w cos θ w

1.4.6  Testing of Pipeline Fluids. Testing of the well/pipeline fluids provides the base data to pre-
dict conditions for FA and IM planning. These short subsections give an outline. Details are in
Frenier and Ziauddin (2008) and Frenier et al. (2010).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 28 03/11/17 6:50 PM


Introduction to the Technology of Flow and Integrity Management  29

Compositional Characterization of Crude Oil/Gas. Important ways for characterizing crude oil
include

• Boiling points from distillation or gas liquid chromatography


• Density distribution from the assay cuts
• SARA (saturates, aromatics, resins, and asphaltenes)
• PNA (paraffin/naphthene/aromatic)
• Water content

Physical Characterization. Common characterization processes include

• Pressure/volume/temperature relationships
• Density
• Viscosity
• Rheology
• Refractive index

Composition of the Aqueous Phases. The composition of the aqueous phases may greatly affect
scale potential, corrosivity, emulsion tendencies, and hydrate formation. Analysis of water chemis-
try can give important information on the scaling tendencies of the formation water or of water that
may become mixed with the formation water. There are a number of standard methods for analyses
of water, including those found in the volume from the American Public Health Association (Clesceri
et al. 1999). Most of the cations that are important for determination of scaling tendencies can be
determined using atomic absorption spectrophotometry or inductively coupled plasma optical emis-
sion spectrophotometry. There are also portable kits (Hach 2005) for analysis of water components
that require addition of specific reagents and examine the development of a color change that can
be read using a simple optical spectrophotometer. Analysis of surface water samples will give some
indication of the presence of scaling ions, but unless a sample can be collected at the formation
face and maintained at the bottomhole temperature and pressure, important scaling ions will not
be present.
The concentrations of acid gases and the resultant pH values are determined by methods described
in Davies and Scott (2006). A major complication with this analysis is that the water content of
a fluid and the chemical composition change in time and place in the pipeline train as a result of
physical changes and introduction or removal of components in the separation and treating stages.
Stimulation and EOR can likewise change the chemical composition of the fluids entering the sys-
tem at any point. Additional testing methods, including microbiological methods, which are used
especially for field samples, are discussed in Section 3.4.3.

1.5  Summary and Lessons Learned


This chapter has provided an introduction to the physical and chemical technologies required to
understand processes for maintaining and protecting the vital pipeline systems.

• The pipeline systems of the oil/gas industry extend from the wellhead to the final consumer
of the products, and the same types of chemical/physical processes will apply throughout the
system. There are various physical and chemical reactions that are common to all systems,
and these were reviewed in this chapter and referenced for further study. The most universal
principles are the varied reactions at surfaces (liquid/liquid, liquid/gas, liquid/solid, and gas/
solid) and the realities of multiphase flow conditions.
• Multiphase flow of some form is the norm for all line segments, though degree and conditions
may change dramatically in different portions of the system. The different flowing forms may
greatly affect corrosion, scale formation, and other FA issues.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 29 03/11/17 6:50 PM


30  Chemical and Mechanical Methods for Pipeline Integrity

• Pipe materials of construction are critical issues for maintaining integrity and long life and
must be addressed when planning a line segment and when maintenance is needed.
• Although an almost infinite variety of chemicals may transit a system, the division into aque-
ous liquids, hydrocarbon liquids, gases, and solids will help make sense of the mixtures.
• As much chemical information about each chemical class and phase as is possible is necessary
for making predictions about FA and IM issues.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 30 03/11/17 6:50 PM


Chapter 2

From the Well to the Consumer

Hydrocarbon-producing wells deliver a vast number of chemicals and chemical mixtures into the
flowlines, into subsurface and surface treating facilities, and then into the pipelines that deliver a
product to a customer. This chapter describes the continuum of maintenance and integrity concerns
that affect the connected pipeline and surface facilities and thus the steady flow of products from the
wellhead to the final user of hydrocarbon chemicals.
The consumer facilities may use a product, such as natural gas, to directly generate power and
heat, or the hydrocarbons may be subjected to processing, such as at a refinery or chemical plant. In
this book, the chemicals produced from the wells that then transit the pipeline systems are grouped
by phase as

• Aqueous liquids
• Hydrocarbon liquids
• Gases
• Solids

This chapter includes short descriptions of the fluids and some solid phases. In addition, here
are described the influence of the reservoir history and type and the effects various possible fluid
phases have on the different categories of pipeline problems found in the upstream and midstream
petroleum environments.

2.1  Description of Well Production Fluids Entering the Pipeline System


Sections 2.1.1 through 2.1.4 provide details of the composition of the fluids and importance to flow
assurance (FA) and integrity management (IM) of the different fluid phases.

2.1.1  Aqueous Phases. Most of the hydrocarbon streams coproduce vast quantities of salt solutions.
The coproduced water includes connate water associated with the native hydrocarbons, as well as
improved oil/gas recovery (IOR) water (including water for pressure maintenance). Additional water
sources include those produced by stimulation treatments. In addition, condensed water from dif-
ferent operations, such as facilities, forms as a consequence of changes in temperature and pressure.
Most important, all the aqueous phases may differ in composition as a result of times and locations.
Included in the stimulation category are aqueous streams from fracturing and acidizing treatments
(see Frenier and Ziauddin 2014), as well as any other aqueous fluids used during separation activi-
ties. The chemical compositions of these water streams vary significantly from almost potable to
very concentrated brines. Fig. 2.1 (USGS 2002) shows a diagram of salinities of produced water
from various parts of the US. Enhanced oil recovery (EOR) and stimulation, especially hydraulic
fracturing (HF), may significantly change the water quantity and quality that is produced at the end
of all the well treatment operations. EOR and pressure maintenance using injected water are major

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 31 03/11/17 6:50 PM


32  Chemical and Mechanical Methods for Pipeline Integrity

Fig. 2.1—Total dissolved solids in produced water, US land (USGS 2002).

reasons that the water cut usually increases as the well field matures. The aqueous phases present in
a specific section of a pipeline system may also change as a function of the well’s life cycle (POSC
2006; see also Section 2.2.1) and various production activities. Dissolved alkaline and alkaline earth
salts and various acids constitute the majority of the “inorganic”- type materials present. However,
almost any soluble element may be present in small quantities, and HF may introduce additional
chemical species (see Shen et al. 2012).
The amount and composition of the aqueous phase may affect or cause corrosion, inorganic scale,
gas hydrates, and emulsions as well as affecting the use of the final hydrocarbon product. Reports
by Ruegamer et al. (2013) and Wilson (2014) claim that the amount of water used in HF treatments
is changing as a result of new knowledge and the changing nature of the wells.
Jacobs (2016) has reviewed the problems of production and injection of water from the Mississip-
pian Lime formation in western Oklahoma, USA. The play is a carbonate formation, and unconven-
tional techniques, especially directional drilling, are used to recover oil and gas from it. Last year,
production in the Mississippian Lime was estimated to be approximately 100,000 B/D of oil, which
accounted for a quarter of the state’s overall oil production. The problem is that the water cuts were
not just higher than normal, they were often extraordinarily high. Newly completed oil wells have
been known to pump out as much as 98% water at a rate of thousands of barrels per day. According
to Jacobs (2016), the injection of so much water into a porous formation (the Arbuckle) has probably
contributed to induced seismic activity in Oklahoma. Therefore, the continued production from this
important resource is in doubt.
Definitions and examples of stimulation treatments are presented in Frenier and Ziauddin (2014).
Any of these treatments and the changes in them may affect water quality and volume.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 32 03/11/17 6:51 PM


From the Well to the Consumer  33

The failure of a pipeline carrying waste water can be as significant as the failure of a line carrying
oil or gas in some situations. The Associated Press (AP 2014) has reported that a leaking under-
ground pipeline near Mandaree, Montana, USA, spilled approximately 1 million gal of saltwater
(from fracturing activities in the Bakken plays) near Bear Den Bay, a tributary of the Missouri River,
which provides potable water to some communities. SPE recognizes that the availability and pro-
duction of water is one of the most important issues facing the hydrocarbon production and transport
industries (SPE 2011).
Water vapor will also be entrained with natural gas as well as the liquid hydrocarbons, and its
removal, termed “dehydration,” is described in Section 2.3.2.

2.1.2  Hydrocarbon Liquids. Hydrocarbon liquids well as the gaseous chemicals described in this
section are usually the primary goal of petroleum well production activities. The in-situ crude oil
being produced constitutes a continuum of soluble chemicals, from C1 to large molecules with
molecular weight in the 750-dalton range.
Fig. 2.2 shows some basic chemical structures that constitute crude oil. The chemicals range from
nonpolar chemicals (i.e., saturated hydrocarbons) to highly polar ones (i.e., aromatic asphaltenes and
nitrogen-substituted chemicals). The lower
hydrocarbons are considered in the gases Paraffins
described in the next section. However, it
should be understood that at undisturbed
n-octane
reservoir conditions there exists a single
hydrocarbon phase, and the gaseous hydro-
carbon phases do not emerge until the bub- n-pentadecane
blepoint (Bp) pressure has been reached. Napthenes
When and where the phase changes take
place may greatly affect the viscosity and
stability of the flowing crude oil and the
production of wax, asphaltenes, and gas
hydrates (these subjects are reviewed in
isooctylperhydrophenanthren
Chapter 4). Depending on the source of
the oil, it may consist of heavy fraction as
Aromatic
well as lightweight fractions. Some shale
N
formations produce a very light (and flam-
mable) product—termed “natural gas liq-
uid” (NGL) (see Appendix A)—that may
xylene
cause flow and gathering line problems quinoline
(these are described in Section 2.3.1). Resins and Asphaltenes

2.1.3 Gaseous Phases. Included in the


category of gaseous phases are hydro-
carbons from C1 to approximately C5
as well as CO2, H2S, N2 (Alvarado et al.
1998), water vapor, low-molecular-weight
organic acids, and possibly oxygen (O2).
HO
All these gases (except O2) may be dis-
An asphaltene
solved in the hydrocarbon or aqueous
phases at reservoir equilibrium and form a
separate gaseous phase at different times 6-dodecylnaphthalene
and places in the production and pipe-
line train. Both chemical reactions and Fig. 2.2—Chemical component families in crude oil.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 33 03/11/17 6:51 PM


34  Chemical and Mechanical Methods for Pipeline Integrity

mechanical processes (temperature, pressure, flow rates) affect the presences and quantities of the
gaseous phases. The dissolved gases also partially control the aqueous fluid pH and may thus deter-
mine the corrosivity of the fluids. Organic acids such as acetic acid can become volatile under some
conditions and may be one cause of top-of-line corrosion (TLC). O2 is always considered to be an
introduced contaminant, given that the undisturbed reservoir conditions are usually reducing and
any oxygen would quickly be reacted.
Production of natural gas (mostly methane, with various amounts of C2–C5 hydrocarbons) is the
goal of many oilfield operations, and this mixture will usually be separated or removed from the pro-
duction stream before conveyance to a customer. In some wells, the production of CO2 or rare gases
such as helium (He) may also be a primary or secondary goal (see Daly 2005). CO2 has become an
important commodity because this gas is a significant EOR chemical.

2.1.4 Solids. Solids are complex materials usually considered to be unwanted contaminants. Sand
and other formation fines can result from uncontrolled production methods or migration of HF
proppants. Corrosion products, organic solids (especially paraffins, asphaltenes, gas hydrates, and
naphthenates), and scale particles also could enter a flow system. The impingement of flowing solids
onto a surface can cause mechanical and corrosion/mechanical damage to many well system com-
ponents (see Section 3.2.3). A sufficient amount of solids can change the flow regime or actually
block the pipe.

2.1.5  Emulsions, Foams, and Solid Dispersions. A wide range of mixed phases can form in pipe-
lines as well as in facilities. Emulsions are a dispersion of one or more liquids in another liquid
(Kokal 2006). Foams are a dispersion of a gas in a liquid. Solids also may be dispersed in liquids
and gases. Important characteristics causing all dispersions in the oil/gas industry include these:

• More than one phase is present.


• Shear forces are needed to mix the phases.
• Surface-active molecules are present to stabilize dispersions.
• Mixtures are thermodynamically unstable.
• Mixtures have a finite lifetime before individual phases separate.

Because of the heterogeneous nature of pipeline-associated fluids, all these factors may present at
any time. More details of problems and applications of mixed phases are provided in Sections 4.6
and 7.5.4 and in Frenier and Ziauddin (2014).

2.2  Effects of the Life Cycle and Reservoir Type on Pipeline Maintenance
The requirements to combat FA and IM problems in pipelines as well as the solutions to them are
largely dictated by the reservoir fluids that are produced, as described briefly in Section 2.1. Conven-
tional and unconventional plays may produce different fluids and fluid volumes that can affect scale
types and locations, corrosion types and locations, and formation of organic solids. Sections 2.2.2 and
2.2.3 provide short descriptions of several conventional and unconventional reservoirs and how pro-
duction and completion methods may affect pipeline problems and solutions. The actual life cycle
of the well, as well as the various stages in the production phase, can also greatly affect the need for
corrosion, scale, and organic solids control in pipelines.

2.2.1  Life Cycle of a Hydrocarbon-Producing Reservoir. The phases of the life cycle of a hydro-
carbon-producing reservoir have been identified in the industry as exploration (discover), appraisal
(define), development (develop), production (deplete), and abandonment (dispose) (POSC 2006).
Except for the earliest phases of exploration, in which geologic and seismic methods are used to find

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 34 03/11/17 6:51 PM


From the Well to the Consumer  35

promising areas where hydrocarbons may be located, large volumes of chemicals are used to aid
production. They are applied during the drilling, completion, production, and abandonment phases.
However, even when additional chemicals are not used, chemistry is important in all the phases (see
Fig. 2.3). Consider, for example, the following applications.

• Exploration/appraisal: Elucidate formation geochemistry and chemistry of the fluids.


• Development: Apply drilling and cementing chemicals to form and complete the well.
• Production: Apply chemicals for stimulation, flow assurance, EOR in the mature field,
and characterization (i.e., tracers).
• Abandonment: Apply cementing chemicals to seal the well and perform monitoring.

Exploration/Appraisal. In the exploration/appraisal phase, geochemical analyses are performed


on the basis of seismic and well probe data and analyses of outcrops or cores. McCarthy et al. (2011)
described the tests that geochemists have used to determine the hydrocarbon-producing potential of
a formation from the rock samples collected. These include total organic carbon analysis to find the
maximum amount of carbon in the rock, as well as a pyrolosis process in which the rock is heated to
increasing temperatures and the effluents are analyzed by several methods described in the papers.
Fluids may be captured in test wells for evaluation, and chemical probes may be placed using wire-
line or coiled tubing. Short summaries of methods used to analyze the liquid samples are described
in Frenier et al. (2010, Chap.3).
Development. In the development phase, various water-based and oil-based fluids are used in
drilling of most wells. Complex oilfield cements are then used to stabilize the production tubing and
to isolate various zones from communication with the surface and from nonproducing formations.
In addition, completion fluids may be used to maintain and control the well’s pressure balance.
Residual effects of drilling and completion chemicals may necessitate the use of acids or HF to
remove damage (see Frenier and Ziauddin 2014).
Production. Various chemicals are applied during all the phases of production to
maintain, control, and, frequently, to enhance the flow of the oil, gas, and aqueous phases.

Chemistry in the Well Life Phases


Exploration/
Appraisal Development Production Abandonment

Stimulation
Drilling and fluids flow assurance in
Geochemical cementing mature field: EOR Cementing
analyses services characterization: Monitoring
tracers

Fig. 2.3—Chemistry in the phases of a well’s life.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 35 03/11/17 6:51 PM


36  Chemical and Mechanical Methods for Pipeline Integrity

Various types of pipelines are used during this phase, and the chemicals affect them in ways
that constitute major themes of this book.
Abandonment. When the well is abandoned, cements and other chemicals are employed to make
sure that the hydrocarbons or other fluids will not reach the surface or pollute aquifers or damage
property. As well, pipelines and gathering lines directly associated with production must be decom-
missioned and safely removed.
Production Subphases. The production activity involves subphases that may affect pipeline activ-
ities. See the review by Lindley (2001) on the production phases; an abstract follows.
Primary Phase. In this phase, the reservoir fluids flow mostly as a result of the initial and internal
pressure of the reservoir. Note that the production is controlled by the pressure differential between
the formation and the bottomhole well pressure. While the pressures may be sufficient for initial
production, stimulation using fracturing and/or reactive chemical treatments may be applied to some
wells to remove formation damage or to improve returns from tight formations (Frenier and Ziauddin
2014). Other production chemicals can also be used to maintain flow, including inhibitors and sur-
factants (these are described in Chapter 6; see also Frenier et al. 2010). For the most part, during
the primary phase, chemicals, including any injected water or gas, are not added to the reservoir
except near the wellbore, so the reservoir is not changed significantly from a chemical standpoint.
However, just by flowing the wells, important equilibrium conditions may be changed. At some
point (either early or very late in the production phase) pressure maintenance will be required. This
may be defined as the secondary phase. Some types of pipelines described in Section 2.3 are in use
during this phase.
Secondary Phase. The pressure to move the fluids through the formations (see Eqs. 1.6 and 1.7)
can be maintained or enhanced by adding a downhole pump to reduce the flowing pressure or by
injecting fluids into the formation. This latter action presents a radical change to the reservoir, given
that a large number of injection wells may be required. This will require a large increase in the num-
ber of pipelines and other piping to serve the expanded production area.
A typical arrangement is a “five-spot” pattern, in which four input or injection wells are located at
the corners of a square—the exact shape of the flood and number of wells depends on the reservoir
dimensions (Singh and Kiel 1982; Chang 2010)—and the production well is placed in the center of
the square. The injection fluid, which is normally water (brine), steam, or gas, is pumped or injected
simultaneously through the four injection wells to displace the oil toward the central production
well (Schlumberger 2010). Otott (2007) presented one description of this layout, and Fig. 2.4 shows
a number of different plans developed on the basis of the reservoir characteristics (Singh and Kiel
1982). The injected fluids act to drive the hydrocarbons to the production wells as well as to main-
tain the flowing pressure. However, these activities may introduce multiple problems, including an
increase in water cut, a change in water saturation and possibly a change in the wettability of the
formation, and introduction of scale-causing ions such as new cations and sulfate. In addition, the
water should be treated to remove dissolved oxygen and a biocide should be added. Not adequately
treating the water is a major cause of corrosion and the “souring” of the reservoir through introduc-
tion of sulfate-reducing bacteria.
Increasing the water cut as well as possibly changing ions in the water can change the corrosion
rates and locations of damage, as well as scaling (more details are in Chapter 3 and Section 4.2).
Because there may now be four injection wells for each production well, the number of connecting
pipes as well as treating facilities may increase and/or change.
Although steam injection may be considered an integral part of an EOR process (or of primary
production for very heavy oils), it can cause multiple problems, including the formation of mixed
deposits (see Frenier et al. 2010; also see Chapter 4). Many chemical treatments can be used in
this phase to maintain production, including inhibitor injections as well as reactive chemical and
propped-fracture treatments. These are described in more detail in Chapters 3, 4, and 5 of Frenier
and Ziauddin (2014).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 36 03/11/17 6:51 PM


From the Well to the Consumer  37

Corner Side
well well

Injection well

(a) Direct line drive


(d) Nine-spot

Four-spot
Seven-spot

(b) Staggered line drive

Smallest area of
flow symmetry

(e) Seven and


four spots
(c) Five-spot, special case of (b)
where d/a = 1/2

Fig. 2.4—Injection/producer patterns (Singh and Kiel 1982).

At some point in the life of many reservoirs, the removal of additional hydrocarbons is not pos-
sible or is economically unproductive because the oil is trapped in the pore spaces and so strongly
adsorbed onto the rock surfaces that injection of water, natural gas, or steam cannot remove economic
amounts. Because of uneven coverage of the reservoir resulting from permeability differences, oil
may also have been bypassed by the sweep fluids. At this point, the massive injection of external
chemicals may be planned and the well may be considered to be in the tertiary production phase.
Tertiary Phase. Frenier and Ziauddin (2014) define this production phase as the tertiary use of
EOR chemicals that may be part of an overall IOR process. As much as 2 × 1012 bbl of conventional
oil and as much as 5 × 1012 bbl of heavy oil will remain in the world’s reservoirs after the primary
and secondary production phases have reached their economic limits. This incremental production is
difficult and expensive but will remain as one of the methods for prolonging production from mature
fields. Note that steam injection may be used at earlier phases in some heavy-oil fields (Thomas
2008).Thus, the injection of large amounts of chemicals to remove some of this oil usually defines
the tertiary phase. Much of the current interest is driven by the high price of oil. EOR activities,
especially CO2 flooding (see Frenier and Ziauddin 2014, Chap. 5), will have a dramatic effect on
pipeline FA and IM activities and concerns. The amount of produced water will increase, and the
low pH can greatly affect corrosion potential as well as scaling patterns. Changes in inhibitors will
probably be required.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 37 03/11/17 6:51 PM


38  Chemical and Mechanical Methods for Pipeline Integrity

Different types of reservoirs may produce different chemicals and may involve changing ratios of
hydrocarbons to water, as discussed next in Sections 2.2.2 and 2.2.3.

2.2.2  Conventional Reservoirs of Oil and Gas. According to Frenier and Ziauddin (2014), con-
ventional reservoirs include those drilled, completed, and stimulated using the tools that have been
used and developed in the past 100 years. Many of the most important techniques reviewed in the
book were developed for use in these types of plays, which exist in carbonate and sandstone forma-
tions. The stimulation methods include acidizing as well as fracturing using proppants. By excep-
tion, they are not “unconventional” (described in Section 2.2.3). However conventional reservoirs
may also include consolidated and unconsolidated formations.
Consolidated Reservoirs. Usually a sandstone formation is cemented together to form a
mass that has substantial compressive strength (see Fig. 2.5a, which shows the constituents,
and Fig. 2.5b, presenting a Berea micrograph). Although carbonate formations—CaCO3 or
CaMg(CO3)2—are much more homogeneous than sandstones, they look much like a sandstone,
even under a microscope, except that most of the matrix is soluble in hydrochloric acid (HCl). As
a consequence, consolidated limestone and sandstone will have high compressive strengths and
can be used as primary building materials. These reservoirs can be stimulated by matrix acidiz-
ing and acid fracturing (carbonates only), as well as by proppant fracturing methods (using both
carbonates and sandstones). Filling a reservoir with a conductive proppant will provide enhanced
permeability and may also improve connections within the reservoir. Lack of connections either
areally or vertically is a major reason for bypassed hydrocarbons. At some point in the well’s life
cycle (POSC 2006), the formation will probably require waterflooding, CO2 flooding, or another

(a) Constituents of
Sandstones

(b) Micrograph of
Berea Sandstone

Fig. 2.5—Sandstone constituents (a) and micrograph of Berea sandstone (b).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 38 03/11/17 6:51 PM


From the Well to the Consumer  39

type of IOR if it is an oil-producing formation. These operations will change the hydrocarbon/
water ratio as well as the chemical constituents, which will alter the impact on the connecting
pipelines.
Unconsolidated Reservoirs. Soft, unconsolidated sand formations exist and frequently have high
permeabilities, but still may be improved through stimulation processes including acidizing or
hydraulic fracturing. Because they may produce sand, frac-pack designs may apply (Morales et al.
2003) to control the sand production. Sand-producing formations can cause significant damage to
piping through erosion/corrosion processes and impingement of the solids (see Section 3.2.3). As
the reservoirs mature, changes in oil, gas, and water ratios and amounts will change pipeline operat-
ing conditions.

2.2.3  Unconventional Reservoirs of Oil and Gas. According to Holditch et al. (2007), the term
“unconventional” usually applies to a low-permeability reservoir (< 0.1 md) that produces mainly
dry natural gas. Many of the low-permeability reservoirs developed in the past are sandstone, but
significant quantities of gas are also produced from low-permeability carbonates as well as shales
and coalbed deposits. Some shale formations have also yielded liquid hydrocarbons (RRCT 2011).
Fig. 2.6 describes the geographic distribution of unconventional original gas in place (Dong et al.
2011) for various sections of the Earth. Note that at the time of the assessment by Dong et al. (2011),
tight gas sands were the largest known source of future gas. However, development of shale gas may
change these estimates significantly.
Cramer (2008) reviewed the stimulation of the “unconventional reservoir,” and he noted that
this term has different meanings to different people. Certain reservoirs that are termed “uncon-
ventional” have a rock matrix consisting of interparticle pore networks with very small pore con-
nections imparting very poor fluid flow characteristics. The author claims that abundant volumes
of oil or gas can be stored in these rocks, and often the rock is high in organic content and is the
source of the hydrocarbon. However, because of marginal rock matrix quality, these reservoirs
generally require both natural and induced fracture networks to enable economic recovery of
the hydrocarbon. Rock types in this class include shale and coalbeds. Cramer (2008) noted that
the term “shale” is a catch-all for any rock consisting of extremely small framework particles
with minute pores charged with hydrocarbon and includes carbonate- and quartz-rich rocks. He
claimed that another type of unconventional reservoir is the stacked pay unit, which exhibits some-
what better pore characteristics than in the case just outlined but with the individual units tend-
ing to be lenticular in shape and having an extremely small size or volume. These two classes of
unconventional reservoirs are amenable to well stimulation.

Region Coalbed Methane (P50) Tight Sands Gas (P50) Shale Gas (P50) Total (TCF) (P50)
Austral-Asia 1,348 6,253 2,690 10,291
North America 1,629 10,784 5,905 18,318
Commonwealth of 859 28,604 15,880 45,343
Independent States
Latin America 13 3,366 3,742 7,122
Middle East 9 15,447 15,416 30,872
Europe 176 3,525 2,194 5,895
Africa 18 4,000 3,882 7,901
World 4,052 71,981 49,709 125,742

Fig. 2.6—Geographic distribution of unconventional original gas in place, in trillion cubic feet (TCF)
(Dong et al. 2011).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 39 03/11/17 6:51 PM


40  Chemical and Mechanical Methods for Pipeline Integrity

When the above rock types become commercially exploited, they are known as resource plays.
Once a low priority, the depletion of conventional reservoirs and improving price for oil and gas
has driven unconventional reservoirs to an important place in the oil/gas industry. In some regions
(i.e., Rocky Mountain Province in the US), unconventional reservoirs represent the primary target
of current activity and remaining hydrocarbon development. Cramer (2008) claimed that given their
unique petrophysical properties, each type of unconventional reservoir requires a unique approach
to well stimulation, with often differing objectives than exist with conventional reservoir types. The
paper reviews the characteristics of the basic unconventional reservoir types, lessons learned, and
successful stimulation practices developed in completing these reservoirs; it also suggests areas for
improvement in treatment and reservoir characterization and in treatment design.
Hydraulic fracturing has contributed greatly to the economic producibility of these reservoirs.
Natural gas production from shallow, fractured shale formations in the Appalachian and Michigan
basins in the US has been under way for decades. What changed the game, as it were, was the rec-
ognition that one could “create a permeable reservoir” and high rates of gas production by intensely
stimulating horizontal wells through multistage fracturing. Chapters 3 and 4 of Frenier and Ziauddin
(2014) describe some of these advances. Details of several important unconventional reservoirs are
presented next.
Shale Gas and Oil Reservoirs. A very important type of predominately silicate formation that
contains shale beds is now being frequently treated using proppant fracturing methods. Shale is the
most common sedimentary formation on Earth (Boyer et al. 2011), but it is quite different from
sandstones that contain sand grains mixed with clays and other minerals. Shale has been defined
as a fine-grained clastic sedimentary rock formed from a clay mud that was compacted over time
(Wikipedia 2010c). As such, the major minerals represented are largely low-permeability reservoir
kaolinite, montmorillonite, and illite. This report noted that clay minerals of Late Paleozoic mud-
stones contain expandable clays, whereas in older rocks, especially in Middle to Early Paleozoic
shales, illite clays predominate. The shales may be very dark, almost black in color, because of the
presence of unoxidized carbon compounds as well as iron oxides (Fig. 2.7). These are also called
“organic-rich shales.” Some shale formations, such as the Haynesville (Buller 2010), may also con-
tain calcite and dolomite and may have HCl solubility up to 15%.
Akrad et al. (2011) called these formations “prospective shales” because the clay content is usu-
ally less than 50%. They noted that the high carbonate content of some of these rocks makes them

(b) Shale sheets

(a) Organic-rich shale

Fig. 2.7—Shale rocks (USGOV.jpg) and shale sheets (Alexander et al. 2011; Boyer et al. 2011).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 40 03/11/17 6:51 PM


From the Well to the Consumer  41

“soft,” and thus working these formations may require different fracturing fluids and proppants for
stimulation.
Shale has long been considered a “source” rock (Cramer 2008; McCarthy et al. 2011) as well as a
barrier trap for migrating hydrocarbons. Fracturing of shale now makes it a viable producing forma-
tion, and these formations are now called “resource rock plays” (see Cramer 2008).
The trapped carbon is the source of the methane/liquids and is the object of the fracture stimulation
treatments. Shales can also be called slates. The major physical attributes are extremely thin lamella
or parallel bands less than 1 cm thick (see Fig. 2.7); these are known collectively as fissility. The
shale beds have very low permeabilities. Soeder (1988) examined a Marcellus shale that was free
of a mobile liquid phase and had a measured gas porosity of approximately 10% under stress with a
fairly strong “adsorption” component. Permeability to gas (k) was highly stress dependent, ranging
from approximately 20 microdarcies (md) at a net stress of 3,000 psi down to approximately 5 md at
a net stress of 6,000 psi (note that most oil/gas-producing sandstones have permeability values in the
1- to 1,000-md range). These properties make extraction of gas from shales extremely difficult unless
they have been fracture treated. Kaufman et al. (2008) report that shale gas formations also have
microfractures and cleats that provide access to the gas. In this characteristic, they share similarities
with coalbeds that can also be fractured to produce methane. Very important fields of gas-producing
shale include the Barnett Shale in central Texas and the Marcellus Group that underlies parts of New
York, Pennsylvania, Ohio, and West Virginia.
An important formation is the Eagle Ford Shale Group of Texas, which runs from the Mexican
border almost to Dallas in the US. According to the RRCT (2011), the lower section of the Eagle
Ford consists of organic-rich deposits and fossiliferous marine shales. Also, a small area of the Eagle
Ford consists of a thin unit between the shales, and this area is especially amenable to hydraulic frac-
turing. The wells in the deeper part of the play deliver a dry gas, but moving northeastward (and with
an updip), the wells produce more liquids (see the discussion in Boyer et al. 2011; McCarthy et al.
2011). One of the fields is actually an oil field (Eagleville, Eagle Ford). Even though the conditions
are severe, these are the fields’ reservoir characteristics:

• 6 to 10% porosity
• 200 to 600 md
• 7,000- to 10,000-psi bottomhole pressure
• 2.0 to 4.5 million psi Young’s modulus
• Bottom hole static temperature (BST) 270 to 300°F

More than 5 million bbl of oil have been produced in 2 years (RRCT 2011) from this reservoir.
The field ends in the vicinity of Dallas, where an outcrop of Austin Chalk over the shale can be seen
(Fig. 2.8). Thus, it runs for almost 400 miles (width approximately 50 miles) and with a maximum
thickness of approximately 250 ft (RRCT 2011). These dimensions are unusual, because the kero-
gen in most oil shale formations must be removed using heat. Fig. 2.8 also shows the layers of the
Marcellus Shale, which is a very important gas play that covers several eastern US states.
Fig. 2.9 shows the location of major shale oil and gas plays. Note that other shale formations
such as the Bakken in North Dakota, USA, are also producing significant volumes of liquid hydro-
carbons. A US Energy Information Administration report (EIA 2014) claimed that as of 2014, the
Bakken region is producing 1 million BOPD of liquids and 1 billion ft3/D of gas. The reader will
understand that this is an ever-changing value.
Table 2.1 describes the properties of the major shale basins in the US. Because these are active
drilling/exploration areas, the map is continually changing.
Boyer et al. (2011) and Alexander et al. (2011) also claimed that there are significant shale gas-
and oil-containing formations in Europe, Africa, Russia, China, and South Asia, but current produc-
tion is not as developed as in North America.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 41 03/11/17 6:51 PM


42  Chemical and Mechanical Methods for Pipeline Integrity

Austin chalk

Eagle Ford
shale

Marcellus shale

Fig. 2.8—Outcrops of Austin chalk, Eagle Ford shale (Wikipedia 2011d), and Marcellus shale
(Arthur et al. 2009).

Fig. 2.9—Shale gas and shale oil plays (Boyer et al. 2011).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 42 03/11/17 6:51 PM


From the Well to the Consumer  43

Gas Shale New


Basin Barnett Fayetteville Haynesville Marcellus Woodford Atrium Albany
Estimated
area, square 5,000 9,000 9,000 95,000 11,000 12,000 43,500
miles
6,500– 1,000– 10,500– 4,000– 6,000–
Depth, ft 600–2,200 500–2,000
8,500 7000 13,500 8,500 11,000
Net thickness, ft 100–600 20–200 200–300 50–200 120–220 70–120 50–100
Depth to base
of treatable ~1200 ~500 ~400 ~850 ~400 ~300 ~400
water, ft
Rock column
thickness
between
5,300– 10,100– 2,125- 5,600–
top of pay 500–6,500 300–1,900 100–1,600
7,300 13,000 7,650 10,600
and bottom
of treatable
water, ft
Total organic
4.5 4.0–9.8 0.5–4.0 3–12 1–14 1–20 1–25
carbon, %
Total porosity, % 4–5 2–8 8–9 10 3–9 9 10–14
Gas content,
3–350 60–220 100–300 60–100 200–300 40–100 40–80
scf/ton
Water
production, N/A N/A N/A N/A N/A 5–500 5–5,000
bbl/day
Well spacing,
60–160 80–160 40–560 40–160 640 40–160 80
acres
Original gas
327 52 717 1,500 23 76 160
in place, tcf
Technically
recoverable 44 41.6 215 262 11.4 20 19.2
resources, tcf
References to various values are in the cited paper.

Table 2.1—Properties of key shale gas basins in the US (Kell 2009).

Dry Pipeline Gas Bakken Gas


Composition, Composition,
Some of the properties of the oil Chemical wt% wt%
and gas produced from the shale Methane, CH4 92.2 55
plays, especially the Eagle Ford Ethane, C2H6 5.5 22
and ­Bakken, may cause significant Propane, C3H8 0.3 13
difficulties for pipeline operations.
­
Butane, C4H10 5
Table 2.2 shows gas compositions
for dry-gas lines (Wocken et al. Pentane, C5H12 1
2013), in addition to a Bakken gas, Hexane, C6H14 0.25
which is “wet” with higher hydro- Heptane, C7H16 0.1
carbons. As the fluids are cooled, Nitrogen, N2 1.6 3
NGLs will form and will make Carbon dioxide, CO2 0.4 0.5
transmission difficult, which may
Higher heating value, Btu.scf 1041 1495
necessitate interventions described
in Chapters 5 and 6. Table 2.2—Gas composition comparisons.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 43 03/11/17 6:51 PM


44  Chemical and Mechanical Methods for Pipeline Integrity

Directional drilling and hydraulic fracturing


are vital technologies that have allowed these
resources to be developed (see Arthur et al. 2009;
King 2010). A diagram showing the placement
of multiple fractures is in Fig. 2.10 (PackersPlus
2011). Additional physical descriptions of frac-
tures are noted in a report by Rassenfoss (2015).
Details of fluids and treatment methods spe-
cifically for shale are in in Frenier and Ziauddin
(2014). These formations may produce dry gas,
Multiple wet gas, and other liquid hydrocarbons, as well
fractures as large amounts of nonconnate water that may
affect and complicate pipeline operations.
Fig. 2.10—Stimulation of a shale bed with multiple
Tight Gas. Deep (hot), tight gas has been
fractures (PackersPlus 2011).
a major focus of fracturing and has led to the
development of many of the chemistries described in Frenier and Ziauddin (2014, Chap. 4). Some
tight gas carbonates can also be stimulated using acid fracturing, as explained in Frenier (2014,
Chap. 3) and Bustos et al. (2007).
Key to proppant fracturing design are the length of the fracture and temperature. Length is impor-
tant to provide contact with the formation. Also, because temperatures are usually hot (>200°), fluid
selection to accommodate the reactivity of the fluids is critical. Note that the conductivity of the
proppant is not important, given that it will always be much higher than that of the formation.
Because the hydrocarbon phase is dry gas (except for the fracturing water), the impact on the
piping and pipelines may be less severe compared to a continual wet-gas flow that is encountered in
shale gas completions.
Coalbed Methane. Coal also is a low-permeability solid (Wikipedia 2010a). Almost all the per-
meability of a coalbed is usually considered attributable to fractures, which in coal are in the form
of cleats. The permeability of the coal matrix is negligible by comparison. Coal cleats are of two
types: butt cleats and face cleats, which occur at nearly right angles. The face cleats are continu-
ous and provide paths of higher permeability, whereas butt cleats are noncontinuous and end at
face cleats. Wikipedia (2010a) claims that on a small scale, fluid flow through coalbed methane
reservoirs usually follows rectangular paths. The ratio of permeabilities in the face cleat direction
over the butt cleat direction may range from 1:1 to 17:1. Because of this anisotropic permeability,
drainage areas around coalbed methane wells are often elliptical in shape. Coalbeds are usually very
wet, and therefore water must be removed to release the gas. Fracturing treatments and the use of
chemical surfactants to drain the water are required. In addition, extremely large volumes of water
are associated with retrieving coalbed methane. The high water volumes must be accommodated by
appropriate scale and corrosion protection systems for the piping and lines.

2.3  Problems Anticipated in Different Areas of Pipeline-Connected Systems


Different types of maintenance problems can be anticipated in gathering lines, surface facilities, and
transmission/trunk pipelines. Many of these issues are, however, controlled by equilibrium changes
in temperature, pressure, flow rate, and product composition and thus have similar chemical/
physical origins. This section describes known FA and integrity problems in three segments in the
hydrocarbon supply train. This separation is somewhat arbitrary, given that piping and pipelines are
integral to every system, including the refineries and chemical plants. In addition, what happens
(including introduction of new chemicals) may affect the stability of subsequent sections because
there is some carryover even if separation of the various chemicals is accomplished.
Note that from a US regulatory standpoint (USDOT 2012), a hazardous pipeline includes “the
pipeline or pipeline system, which means all parts of a pipeline facility through which a hazardous

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 44 03/11/17 6:51 PM


From the Well to the Consumer  45

liquid or carbon dioxide moves in transportation, including, but not limited to, line pipe, valves, and
other appurtenances connected to line pipe, pumping units, fabricated assemblies associated with
pumping units, metering and delivery stations and fabricated assemblies therein, and breakout tanks.”
The primary federal regulatory agency in the US is the Pipeline and Hazardous Materials Safety
Administration (see PHMSA 2012).
Refineries and chemical plant facilities are not within the scope of this book because many of the
chemical reactions needed to refine the crude oil are more complex (and at much higher temperatures)
than those in the upstream/midstream sectors. However, the refineries and facilities are affected by the
chemicals applied in the upstream parts of the systems. In addition, the unit operations and types of
equipment in some refinery operations are similar to those in upstream treating facilities (see Section
2.3.2). For information on refinery cleaning processes, see Frenier (2001) and Frenier (2017b).
Various types of additional products not directly associated with oil/gas production are conveyed
using pipelines. Examples of the range of other materials include, but are not limited to,

• Coal and other solid slurries


• Alcohols and other liquid or gaseous chemicals
• Anhydrous ammonia
• Finished petroleum product (e.g., gasoline and diesel fuels)
• Potable water

All these lines may have FA and IM issues similar to those faced by oil/gas pipelines, but these
lines are excluded from the current book because of space and time limitations. The reader is encour-
aged to examine Mohitpour et al. (2010), a book that mentions finished products but is also primar-
ily associated with oil/gas transportation.
In the normal parlance of the oil/gas industry, gathering lines, transfer lines, and surface facilities are
considered part of the “upstream environment”; the transmission lines are described as “midstream”;
and refineries and chemical plants are “downstream.” Note, however, that as long as the conditions are
the same, chemistry and physics work the same regardless of the sector. So the lines between the various
oil/gas sectors are frequently blurred, and operators therefore need knowledge of the entire supply train.

2.3.1  Gathering Lines and Wastewater Lines. Gathering lines are the piping systems that connect
individual wellheads to larger collection and treatment facilities. These systems include onshore as
well as offshore well production systems. See Fig. 1.5 for one example of a well gathering line lay-
out. In this category are included lines that move partially treated fluids to further treatment before
the fluids enter a transmission system. Also included is piping that feeds water for flooding, EOR,
and disposal activities. The water is generally recycled water from other production wells. The
wastewater lines can be composed of polymers, which are not easily corroded, as well as large steel
lines (see Harris et al. 2010). The latter are subject to corrosion, scale, and microbiologically influ-
enced corrosion (MIC). The waste water from oil/gas operations is becoming a valuable commodity.
A report by Boschee (2015) describes the very complex system of water recovery and treatments
conducted by Chevron in California. In this case, the recovered water is used for agricultural treat-
ments. The article mentions an 8-mile internally coated pipeline constructed as part of this system.
In the shale oil/gas fields, very large volumes of water are needed for fracturing treatments; there,
the highly contaminated returned fluids are being reclaimed, and much of the water transits newly
constructed pipelines (see Locke and Kimball 2013; Rao 2015). Methods for cleaning recycled
water are described in Section 8.4.1.
Fig. 2.11 shows a small section of a complex onshore system with numerous connections, valves,
filters, and control devices that can become fouled and/or corroded. Ruptures of wastewater pipelines
have caused damage in various oil/gas fields, including several in the Bakken play (Scheyder 2015;
AP 2014). Details of wastewater treatment methods and spill cleanup are in Sections 8.4.1 and 8.4.2.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 45 03/11/17 6:51 PM


46  Chemical and Mechanical Methods for Pipeline Integrity

Gathering lines may initially


experience wellhead tempera-
tures and pressures as well as
multiphase and multichemical
mixtures. The temperature and
pressure changes that occur as the
fluids transit the gathering lines
may set up conditions for various
types of corrosion, including MIC;
formation of inorganic scales; and
organic deposits, including gas
hydrates and wax.
Fig. 2.11—Fluid injection system.
Producers with subsea wells
usually explicitly plan to deal with
wax and hydrate deposition if conditions favorable to their deposition exist at any time in the well
life cycle (Frenier et al. 2010; Kelland 2013; Borden 2014). Pressure changes and mixing of incom-
patible waters provide the conditions for forming calcite and sulfate scales. Asphaltene deposition
is less likely in this segment, but it is not unknown. Risers in subsea producing systems connect
the seafloor facilities (i.e., gathering systems and subsea separators) to the surface. These piping
systems may convey untreated well fluids, like gathering lines, or they may convey partially treated
gases or liquids. Under the categories of gathering lines are the various connections to/from the
wellhead for various activities, including injecting fluids into the well/reservoir systems.
The mixing of various well fluids at manifolds may also change the chemical compositions, initiating
scale, organic solids, and conditions for corrosion. Gathering lines are constantly subject to chemical/
physical upsets as wellfield conditions change. Thus, many kinds of problems may occur in a gathering
line. Wint (2013) claimed that there is a range of issues for gathering lines serving shale plays that pro-
duce “rich gas” (i.e., wet gas), which contains significant levels of liquefiable hydrocarbons (such as eth-
ane or propane and higher alkanes), along with methane gas. Liquids can accumulate at low-elevation
points along gathering systems where the high liquid concentrations in the gas streams cause significant
issues with slugging, high differential pressures (liquids loading), and corrosion (see Preface as well as
Section 2.2.3). These conditions may require almost constant pigging (see Sections 5.5.1 and 5.2.1).
In addition, crude oil containing high levels of paraffin and other flow-reducing contaminants (e.g.,
fracturing sand, chlorides, and spent chemicals) presents flow restriction issues in these upstream
pipeline systems. Many factors contribute to the overall performance and flow efficiency of pipeline
systems that may include the elevation profile, flow volumes, product quality, and temperature; all
pipelines must be evaluated on an individual basis.
The gathering lines and piping connections are subject to a wide range of possible corrosive
attacks (see Chapter 3). Because of the wide variety of chemicals and mixtures in the lines, corrosive
attack is difficult to eliminate, especially as the lines age (see discussion in Berger 2007).

2.3.2  Surface and Subsurface Facilities. Surface and subsurface facilities include primary and
secondary separation facilities where the aqueous, hydrocarbon liquid, and gaseous phases are
treated to produce pipeline-transmission-quality products. Mokhatab et al. (2006) noted that pro-
cessing is designed to separate natural gas, condensate, incondensable liquids (e.g., water and
hydrocarbons), and acid gases. These processes will allow the further processing and sale or reuse
of the valuable products. Several current engineering/chemical operations are in use to achieve
these goals. Fig. 2.12 shows a generic flow chart of the processes. In the diagram, the fluids are
usually subjected to a physical separation sequence to produce aqueous, liquid hydrocarbon, and
gaseous phases. The different phases may then be given additional treatment, depending on the
original composition and needs (details are in the 2.3.2 subsections).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 46 03/11/17 6:51 PM


From the Well to the Consumer  47

Acid gas removal

Well fluids
Physical separation Gas Compression

Solids Oil Water

Pipeline

Refinery Injection/recycle

Fig. 2.12—Well fluids treatment processes.

Treatment facilities range


in complexity from a single Flare stack
tank or tower at a well pro- Separation or treatment towers
duction site to very large inte-
grated plants. Walton (2013)
showed a photo (Fig. 2.13)
of a gas/liquids processing
plant (ONEOK Partners)
in the US state of North
Pipelines
Dakota. This type of sub-
stantial plant is necessary
in field locations to prepare
large volumes of well prod-
ucts for further transit. Such
plants may serve a part of a
Fig. 2.13—Gas treatment plant, ONEOK Partners Stateline Processing
well field, with the individual plant in North Dakota (Walton 2013).
well connected to units with
pipelines. Fig. 2.14 shows a
photo­­
graph of a land-based Piping Reactors/separators
gas treatment plant that is
slightly different from the
plant shown in Fig. 2.13.
Note the amount of piping
that connects to underground
lines. The complexity will
be determined by the fluid
chemical quality, volumes,
and other processing re­­
quired downstream to achieve
transport-quality products. Fig. 2.14—Land-based gas treatment plant.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 47 03/11/17 6:51 PM


48  Chemical and Mechanical Methods for Pipeline Integrity

Shell and Tube Heat Exchanger Outlet tube side


and valve
Inlet shell side
and valve

Inlet tube side


and valve

Head (open) Tube sheet

Inlet shell side


and valve

Fig. 2.15—Drawing of a shell and tube heat exchanger.

An example of a complex
integrated treatment plant was
described by Jarragh et al.
Reboiler
(2013). They noted that the
Kuwait Oil Company has facil-
ities that consist of 22 crude-
processing plants (or gathering
Hot process centers) as well as four gas
vapor and processing plants (or booster
liquid outletq
stations). Two effluent water
disposal plants, as well as a
Steam
inlet seawater treatment plant and
injection plants, are included
in a vast network of pipelines
carrying different products,
Tubes as well as early production
facilities. They also provided
several detailed diagrams that
show the large number of indi-
vidual treatment units, pumps,
and connecting piping that
must be protected from corro-
sion and flow problems. Note
that these diagrams are not
Hot condensate
outlet reproduced here because of
their complexity.
Jarragh et al. (2013)
Warm process described several important
liquid inlet types of processing units.
These include heat exchang-
ers (Fig. 2.15), reboilers
Fig. 2.16—Reboiler schematic. (Fig. 2.16), and various towers

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 48 03/11/17 6:51 PM


From the Well to the Consumer  49

used for hydrocarbon separation as well as for contacting a liquid with a gas to cause a chemical
or physical reaction. This will include dehydration or removal of an acidic gas. One example of a
bubble-cap tower is shown, though many other designs are possible. See Fig. 2.17 for details of the
gas contact in the countercurrent flow reactor (the cleaning of these units is covered in Section 7.6).
Numerous types of physical separation devices are shown in Figs. 2.18 and 2.19. A wide range
of organic, inorganic, and mixed deposits can foul these units. These solids come from the inlet
gases and liquids as well as from the reactions that take place in the units. See Engel and Sheilan
(2014) as well as examples in Section 4.5.

Liquid flow

Overflow
weir
Gas flow Downcommer

Acid gas removal


tower
Liquid flow

Gas flow

Details of flow

Fig. 2.17—Bubble-cap gas removal tower.

Gas to
Typical crude oil compression,
treatment system dehydration
(separation, heating,
dehydration, stabilization,
Multiphase meter storage, metering, pumping)
(Test separator) Oil dehydration
Water-to-water
Gas to Gas to treatment
Electrostatic treater,
compression, compression, heater treater
processing processing degasser, desalter

MP separator LP separator

Crude heater
From Manifold Water-to-water Water-to-water Crude/crude
wells treatment treatment exchanger

Crude oil Crude oil


storage tank storage tank

Crude oil To crude oil


metering skid export pipeline

Crude oil Crude oil


booster pumps export pumps

Fig. 2.18—Separator layout (Schlumberger 2010).

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 49 03/11/17 6:51 PM


50  Chemical and Mechanical Methods for Pipeline Integrity

Safety valve Coalescing plates Mist Gas line


extractor
Inlet breaker Deflector plate Foam breaker

Effluent
inlet

Oil-level
controller
Water-level Water line
Oil line and float
controller and float
Weir plate Vortex breaker

Fig. 2.19—Gravity settler (Sims 2010).

The production of oil and gas from the shale formations of North America has developed so rap-
idly that the industry has challenges keeping up with the demand. Boschee (2014d) reported that
the shorter time frame for construction of various facilities in the shale environment may reduce the
planning time. This means that the safety of the designs, especially pressure containment (and thus
the integrity of the piping), is a major consideration.
Physical Separation. The surface facilities are similar to some refinery operations in that chemi-
cal and mechanical methods are used, including heating and cooling of fluids. Most initial separa-
tions depend on gravity and density. This may take place in tanks as well as in more-complicated
equipment. The aqueous, liquid hydrocarbons and the gases may also undergo additional physical
and chemical reactions to dehydrate the hydrocarbons and remove CO2 and hydrogen sulfide (H2S).
Fig. 2.18 shows a detailed layout for initial surface separation of the different phases when oil,
water, and gases are present. Kokal (2006) and Kelland (2009) note that application of heat promotes
oil/water separation and accelerates the treating process. An increase in temperature has three sig-
nificant effects: It reduces the viscosity of the oil, increases the mobility of the water droplets, and
elevates the settling rate of the water droplets.
Fig. 2.19 describes a gravity separator in more detail. Some subsea separation equipment can pro-
vide a level of dehydration and liquid/gas separation (for more information, see Estefen et al. 2005).
Mokhatab et al. (2006, Chap. 5) examined phase separation theory and designs. Because of the great
variety of chemical processes that are used, separation facilities are frequently fouled with mixtures
of organic solids, inorganic solids, and corrosion products.
Bedwell et al. (2015) reviewed problems found in aging facilities and how both engineering and
chemistry can lengthen the life of these vital units. A drawing by these authors indicates some of
the problems, which include foam, emulsions, and scale. The planning “circle” (Fig. 2.20) indicates
process steps for reducing the impact of these issues, which are addressed more fully in Chapter 4.
In many cases, the separation of solids, as noted by Rawlins (2013), is a highly important process
step that must preceed some of the other treating steps. This author reviewed the literature on solids
prevention at the well using sand control methods described in the paper and by King et al. (2003).
Also reviewed was physical separation in the surface facilities. Rawlins further noted that the solids
can be natural substances from the reservoir (including sand, clay fines, and scale) and introduced
solids (including fracturing proppant). This latter category is increasingly present as oil/gas produc-
tion from shale plays includes more hydrocarbon production.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 50 03/11/17 6:51 PM


From the Well to the Consumer  51

Gas outlet Gas outlet


Foarm brkr.

Foarm brkr.

Foarm brkr.

Foarm brkr.
Mist elim.

Mist elim.
Inlet

1. Theoretical
modeling

Oil outlet Water outlet Water outlet

5. Optional 2. Fluids and


process solids sampling
dianostics and analysis

4. Chemistry
3. System
and flow
mapping
assurance

1,000 ppm 5% Basic


oil-in-water sediment
and water

Fig. 2.20—Flow assurance (FA) problems and planning for maintaining separation facilities (Bedwell
et al. 2015).

Rawlins (2013) claims that desanders can be incorporated into a surface facility (Fig. 2.21a), or
they can be standalone if sand/solids production is of a higher volume than can be handled by the
sand filter (Fig. 2.21b).
Dehydration. After the primary separation, a natural gas stream may require further dehydration
and removal of the acid gases (H2S, CO2, and organic acids such as acetic acid). These operations
are required to reduce corrosion to pipelines and to provide a method for controlling scale and gas
hydrate formations. Dehydration may follow acid gas removal or may be a standalone operation if
the gas does not need acid removal treatment (see the following subsection for details of acid gas
treatments).
Mokhatab et al. (2006) claimed that there are two fundamentally different processes for dehydra-
tion: refrigeration as well as removal of the water using liquid or solid desiccants. Chapter 7 of their
book has detailed information on these processes.
Several low-temperature processes, among them refrigeration, are able to effect some degree of
dehydration and have been used for many years (Records and Seely 1951). Records and Seely
(1951) described the technology for low-temperature dehydration of natural gas using the Joule-
Thomson (J-T) effect. The change of temperature with respect to a change of pressure in a J-T
process is the J-T coefficient (Kittel and Kroemer 1980):

µ JT = ∂T
∂P ( ) H
. ���������������������������������������������������������������������������������������������������������������������������� (2.1)

The value of mJT is typically expressed in °C/bar (SI units: K/Pa) and depends on the specific gas
as well as the temperature and pressure of the gas before expansion. The process works by reducing
the temperature of a water-saturated gaseous mixture by expansion through a restrictive orifice (i.e.,
J-T effect) or by other methods described in subsequent paragraphs. It also reduces the equilibrium
water content of the gaseous mixture as a direct function of temperature. Then, gravity difference
forms and separates hydrate particles from the gaseous mixture.

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 51 03/11/17 6:51 PM


52  Chemical and Mechanical Methods for Pipeline Integrity

Facilities Desander (a)


Wellhead Produced
water
desander
Choke

Well-stream
desander with
integral
accumulator

Well-stream Solids transport and disposal system


desander with
external
accumulator

Wellhead
Desander (b)

Fig. 2.21—Desander applications (Rawlins 2013).

Note that chilling a gas stream by J-T expansion can also lead to hydrate formation if the pressure
is high enough. Dehydration by a mechanical refrigeration using this or other methods (e.g., external
gas expansion chillers using gas J-T processes) with glycol injection to prevent hydrate formation
is in current use (Huffmaster 2004). Mokhatab et al. (2006) noted that propane is a commonly used
refrigerant in such units. These types of chillers can also be used to produce NGL.
Hubbard (1991) claimed that the removal of water from natural gas can also be accomplished in
several additional ways. Commercial methods currently used when this report was written included
absorption using glycol dehydration and adsorption using a dry desiccant. These authors noted that
the two methods use mass transfer of the water molecule into a liquid solvent (glycol solution) or a
crystalline structure (dry desiccant). The third method uses cooling to condense the water molecule
to the liquid phase, with the subsequent injection of inhibitor (glycol or methanol) to prevent hydrate
formation.
Glycol absorption and dehydration involve the use of a liquid desiccant to remove water vapor
from the gas. In this process, the water forms a stable solution with the complex alcohols. Although
many liquids possess the ability to absorb water from gas, the liquid that is most desirable to use for
commercial dehydration purposes should possess the following properties:

• High absorption efficiency


• Easy and economic regeneration
• Noncorrosive and nontoxic properties
• No operational problems when used in high concentrations
• No interaction with the hydrocarbon portion of the gas
• No contamination by acid gases

BK-SPE-CHEMICAL_AND_MECHANICAL-170274.indb 52 03/11/17 6:51 PM

You might also like