Applied Subsurface Mapping Geological
Applied Subsurface Mapping Geological
SUBSURFACE GEOLOGICAL
MAPPING
WITH STRUCTURAL METHODS
2nd Edition
Daniel J. Tearpock
and
Richard E. Bischke
Tearpock, Daniel J.
Applied subsurface geological mapping / by Daniel J. Tearpock and
Richard E. Bischke.
p. cm.
Includes bibliographical references and index.
ISBN 0–13–859315–9
1. Petroleum—Geology. 2. Geological mapping. I. Bischke,
Richard E. I. Title.
TN870.5.T38 1990
550′.22′3—dc20 90–35994
CIP
Editorial/production supervision
and interior design: Jacqueline A. Jeglinski
Cover design: Lundgren Graphics
Manufacturing buyers: Kelly Behr and Susan Brunke
ISBN 0130919489
Text printed in the United States at International Book Technologies in Troy, New York.
7th Printing September 2007
FOREWORD xxi
PREFACE xxiii
ACKNOWLEDGMENTS xxv
Reviewers xxv
Contributors xxv
Drafting xxvi
Support Personnel xxvi
Contributing Authors xxvii
Special Recognition from Daniel J. Tearpock xxviii
Special Recognition from Richard E. Bischke xxviii
Biography of Daniel J. Tearpock xxix
Biography of Richard E. Bischke xxx
v
vi Contents
TEXTBOOK OVERVIEW 1
INTRODUCTION 8
THREE-DIMENSIONAL PERSPECTIVE 9
RULES OF CONTOURING 12
INTRODUCTION 43
INTRODUCTION 60
General Log Measurement Terminology 60
UNCONFORMITIES 128
INTRODUCTION 175
INTRODUCTION 251
INTRODUCTION 332
CONCLUSION 505
INTRODUCTION 506
INTRODUCTION 584
INTRODUCTION 635
SUMMARY 679
CONCLUSIONS 680
INTRODUCTION 682
Expansion Index for Growth Faults 683
CONCLUSIONS 721
INTRODUCTION 723
APPENDIX 788
REFERENCES 792
INDEX 810
This page intentionally left blank
FOREWORD
xxi
xxii Foreword
Many textbooks cover numerous geologic subjects; however, since Margaret S. Bishop’s classic
textbook (1960), and the first edition of this textbook, no complete and detailed book on the sub-
ject of subsurface mapping and structural methods has been published.
Subsurface geological maps are the most important and widely used vehicle to explore for
and develop hydrocarbon reserves. Geologists, geophysicists, and engineers are expected to un-
derstand the many aspects of subsurface mapping and to be capable of preparing accurate sub-
surface maps. Yet, the subject of subsurface mapping is probably the least taught of all
petroleum-related subjects. Many colleges and universities do not teach applied subsurface map-
ping courses, and over the past decade, many company-sponsored training programs have been
curtailed or eliminated.
As we enter the new millennium, we must become more aware of our limitations. This in-
volves questioning our methods and thinking more about our interpretation techniques. We need
to consider the tools we have at our disposal to support our interpretations and to generate a bet-
ter quality product. Inaccurate procedures, unjustified shortcuts, and limited mapping and struc-
tural skills will result in a poor product. During the past decade, the petroleum industry has
experienced sweeping changes; new technologies have emerged requiring new skills.
In today’s petroleum exploration and development activities, a geoscientist spends a great
deal of his or her time in front of a workstation or computer correlating, interpreting and map-
ping with the ultimate goal of generating viable interpretations resulting in economic prospects.
The computers provide increased speed and efficiency in nearly every aspect of exploration and
development. However, problems have developed with the computer-based activities. What we
often see today is the computers, not the interpreters, driving interpretations and maps, with
blind acceptance of the results. Too often, this procedure results in contour maps that violate geo-
logic principles, interpretations that are unlikely in three dimensions, and prospects that depict
subsurface geology and geometry that are outright impossible.
xxiii
xxiv Preface
We cannot accept computers driving interpretations, nor can we blindly accept the resultant
maps (Tearpock and Brenneke 2001a, 2001b). In the hands of geoscientists properly educated in
the basics and fundamentals of geology, including field experience, the workstations and per-
sonal computers are powerful tools. This text contains many of the techniques and methods re-
quired to generate sound geologic interpretations and prospects. It can help you become a more
productive and successful geoscientist or engineer.
It has been estimated that 30 percent or more of future reserve additions will be found in
areas that are maturely developed. These reserves will come from redeveloping older oil and gas
fields. This future potential is to be found in proved producing reservoirs, reservoirs with proved
reserves behind pipe, various types of accumulations (attic, infill, and untested fault blocks), and
wildcatting in and around the fields and in deeper stratigraphic sections than the current limit
within the area.
The techniques presented in this book, from correctly mapping well data to structural bal-
ancing, are ideally suited for both exploration and development activities, but especially valuable
when conducting detailed geoscience projects involving the development or redevelopment of a
mature area.
We have expanded this second edition to include new methods or techniques developed dur-
ing the last 10 years, the understanding and applications of computer-based log correlation, cross
section construction and 3D seismic interpretation. We have organized a wealth of information
that exists in the literature, in addition to material that has never been published. This second
edition builds on the new advances and our experience in petroleum exploration and exploita-
tion, as well as our extensive experience in teaching subsurface exploration and development
mapping and structural geology courses to geologists, geophysicists, and engineers in the energy
industry (worldwide), and at the college and university level.
We present a variety of subsurface mapping and structural techniques applicable in the four
major petroleum-related tectonic settings: extensional, compressional, strike-slip, and diapiric.
The detailed techniques presented throughout the book are intended to expand your knowledge
and improve your skills in preparing geologic interpretations. The knowledge of the principles
and techniques presented, plus a burning desire to explore the unknown, will ensure your
success.
This textbook is specifically designed for geologists, geophysicists, and engineers who pre-
pare subsurface geological interpretations with accompanying maps. This book should also be
beneficial to supervisors, managers, technical assistants, investors, and other persons who have a
requirement for the use, preparation, or evaluation of subsurface interpretations and maps.
Good luck and good prospecting.
Daniel J. Tearpock
Richard E. Bischke
ACKNOWLEDGMENTS
xxv
xxvi Acknowledgments
solutions rapidly. We thank Randy for his contribution of the section on computer-based fault
seal analysis in Chapter 7.
Second Edition
I want to thank the great staff at Subsurface Consultants & Associates, LLC (SCA), and espe-
cially Hines Austin, who continuously provided encouragement and support throughout the
preparation of this second edition. A special thanks goes out to my daughters Nicole and
Danielle and my grandsons Justin, Jonathan, Tyler, and Jesse for their understanding and pa-
tience with their busy dad and “Pappy.” Finally, to my parents John (deceased) and Laura
Tearpock who sacrificed much in their lives to make sure that I received a college education,
which without this geoscience education, this book would not have become a reality. Thank you,
both.
Second Edition
I would like to thank John Suppe, Peter Verrall, Jeffrey Milnes, Joe Brewton and Dan Tearpock
for helping me to understand how the earth works. I also thank Larry Walker for his meticulous
review of this manuscript.
DANIEL J. TEARPOCK
(Subsurface Consultants & Associates, LLC.)
Daniel J. Tearpock is the Chairman and CEO of SUBSURFACE CONSULTANTS & AS-
SOCIATES, LLC. (SCA) headquartered in Houston, Texas. SCA is an international upstream
petroleum consultancy and training company. In addition to managing the overall company, he
conducts and supervises a variety of consulting projects in the areas of exploration and develop-
ment, acquisitions and divestitures, in addition to serving as project manager or advisor to vari-
ous oil companies. He also conducts certain training courses and seminars around the world.
Mr. Tearpock has authored or co-authored three books and over 30 technical articles in the
areas of project management, petroleum subsurface mapping, shared earth models, structural ge-
ology and geothermal energy. He has generated numerous exploration and development projects,
either as the sole generator or as part of an organized team. The largest development discovery
was 4,500,000 BO in the offshore Gulf of Mexico in 1980. The largest exploration discovery
was over 250 BCF of gas in the offshore Gulf of Mexico in 1988.
Some of Dan Tearpock’s recognitions and honors include: Honored Professional in Mar-
quis Who’s Who in Science and Engineering, 1998–1999 Edition; 1998 Recipient of the Distin-
guished Service Award of the Bloomsburg University Alumni Association; Entrepreneur of the
Year Finalist, 1996 and 1998 (Sponsored by USA Today, NASDAQ, Ernst & Young, LLP. and
The Kauffman Foundation); Honored Professional of National Directory of Who’s Who in Exec-
utives and Professionals, 1995–1996 Edition. Mr. Tearpock served as an adjunct associate pro-
fessor at Tulane University (1984–1985), teaching in the Master of Petroleum Engineering
program. He was responsible for teaching several geoscience courses. He was also a geology in-
structor at Montgomery College (1974–1975).
Mr. Tearpock received his Bachelor’s Degree in Earth Science from Bloomsburg University
(1970) and his Master’s Degree in Geology from Temple University (1977). He is a certified
Petroleum Geologist (4114) and a member of various organizations, including the American As-
sociation of Petroleum Geologists, American Geological Institute, American Petroleum Institute,
European Association of Geoscientists and Engineers, Geological Society of America, Houston
Geological Society, Indonesian Petroleum Association, Lafayette Geological Society, New
Orleans Geological Society, Society of Exploration Geophysicists, and Society of Petroleum
Engineers.
xxix
RICHARD E. BISCHKE
(Subsurface Consultants & Associates, LLC.)
Dr. Richard (Dick) Bischke is Chief Structural Geophysicist for Subsurface Consultants &
Associates, LLC. (SCA). He works primarily on worldwide projects for major oil companies and
large independents. He is a 3D-workstation geophysicist specializing in the interpretation of and
prospect generation within complexly deformed structures, including salt-related structures.
He has conducted extensive studies throughout the world, including North and South
America, North Africa, Pakistan, Madagascar, Europe and Southeast Asia. He assisted in the ex-
ploration and development of the super giant El Furial Trend and in the exploration of the off-
shore Orinoco Delta Complex in Venezuela, and has worked on other projects throughout
Venezuela and other regions of South America. Dr. Bischke has significant expertise in all the
tectonic provinces, including compressional, extensional, strike-slip, and diapiric. In the United
States he has conducted studies in the eastern and western overthrust belts, including the Ap-
palachians, the Northern and Southern Rockies, and in the Ouchita Fold Belt, as well as the
North Slope of Alaska, offshore and onshore California, shallow and deep water Gulf of Mexico,
and the Gulf Coast and Midland Basins.
He is also heavily involved in teaching courses in structural geology and balancing in com-
pressional, extensional, and strike-slip regimes. He has authored many papers and several books
on tectonics and earthquakes, on extensional, compressional and strike-slip structures, and on
subsurface mapping and prospecting.
Dr. Bischke was an Associate Professor of Geology and Geophysics at Temple University
(1973–1982) and an adjunct Professor of Engineering at Drexel University. During this time, he
became associated with the consulting firm of International Exploration (INTEX). He consulted
mainly with major oil companies and large independents, primarily in frontier basins, from
1975–1988. After participating in an extensive structural study in the Philippines on a basin eval-
uation project sponsored by the World Bank for the Philippine National Oil Company, he worked
with Dr. John Suppe on the prospect generation within several Philippine fold belts. He later
joined Dr. Suppe at Princeton University to conduct applied research on extensional and com-
pressional structures (1987–1991).
Dr. Bischke graduated from the University of Wisconsin-Milwaukee with a B.S. in Geology
and an M.S. in Structural Geology. He received a Ph.D. from Columbia University in Tectono-
physics and Rock Mechanics. He is a member of several societies, including the American Asso-
ciation of Petroleum Geologists, Houston Geological Society, and Lafayette Geophysical
Society. Dr. Bischke and Dan Tearpock, as co-authors, have won Best Paper Awards from the
New Orleans Geological Society and the Gulf Coast Association of Geological Societies.
xxx
CHAPTER 1
INTRODUCTION
TO SUBSURFACE
MAPPING
TEXTBOOK OVERVIEW
This textbook begins where many others end. It focuses on subsurface structural and mapping
methods and techniques and their application to the petroleum industry. These techniques are also
important and applicable to other fields of study, and geologists, geophysicists, engineers, and
students in related fields, such as mining, groundwater, environmental, or waste disposal, should
benefit from this text as well.
The objectives of subsurface petroleum geology are to find and develop oil and gas reserves.
These objectives are best achieved by the use and integration of all available data and the correct
application of these data. This textbook covers various aspects of geoscience interpretation and
the construction of subsurface maps and cross sections based upon data obtained from well logs,
seismic sections, and outcrops. It is concerned with correct structural interpretation and mapping
techniques and how to use them to generate the most reasonable subsurface interpretation that is
consistent with all the data.
Subsurface geological maps are perhaps the most important vehicle used to explore for
undiscovered hydrocarbons and to develop proven hydrocarbon reserves. However, the subject of
subsurface mapping is probably the least discussed, yet most important, aspect of petroleum
exploration and development. As a field is developed from its initial discovery, a large volume of
well, seismic, and production data are obtained. With these data, the accuracy of the subsurface
interpretation is improved through time. The most accurate interpretation for any specific oil or
gas field can be prepared only after the field has been extensively drilled and most of the hydro-
carbons have been depleted. However, accurate and reliable subsurface interpretations and maps
are required throughout all exploration and development activities.
1
2 Chap. 1 / Introduction to Subsurface Mapping
From regional exploration to a field discovery and through the life of a producing field, many
management decisions are based on the interpretations geoscientists present on subsurface maps.
These decisions involve investment capital to purchase leases, permit and drill wells, and work
over or recomplete wells, just to name a few. An exploration or development prospect generator
must employ the best and most accurate methods available to find and develop hydrocarbon
reserves at the lowest cost per net equivalent barrel. Therefore, when preparing subsurface inter-
pretations, it is essential to use all the available data, evaluate all possible alternate solutions, use
valid structural interpretation methods, and use the most accurate mapping techniques to arrive
at a finished product that is consistent with correct geologic models.
Subsurface geoscientists have the formidable task of mapping unseen structures that may
exist thousands of meters beneath the earth’s surface. In order to accurately interpret and map
these structures, the geoscientist must have a good understanding of the basic principles of struc-
tural geology, stratigraphy, sedimentation, and other related geological disciplines. The geosci-
entist must also be thoroughly familiar with the structural style of the region being worked. Since
all subsurface interpretations and the accompanying maps are based on limited data, the geolo-
gist, geophysicist, or engineer must use (1) his or her educational background, (2) field and work
experience, (3) imagination, (4) an understanding of local structures, (5) an ability to visualize in
three dimensions in order to evaluate the various possible alternate interpretations and decide on
the most reasonable, and (6) correct subsurface structural and mapping methods and techniques.
There are many textbooks on structural geology, tectonics, stratigraphy, sedimentology,
structural styles, petroleum geology, and other related geologic subjects. Since Bishop’s classic
work (1960), our original text published in 1991 was the only single-source textbook on applied
subsurface mapping techniques. Many new developments occurred during the 1990s. In particu-
lar, the field of structural geology advanced to the extent that structural methods and techniques
have enhanced subsurface interpretation and mapping. As we enter the 21st century, the objective
of this second edition of Applied Subsurface Geological Mapping is to present a variety of sub-
surface structural, mapping, and cross-section methods and techniques applicable in various geo-
logic settings, including extensional, compressional, strike-slip, and diapiric tectonic areas. The
detailed structural and mapping techniques illustrated throughout the book are intended to expand
your knowledge and improve your skills in preparing geological interpretations using a variety of
maps and cross sections.
All energy companies expect positive economic results through their exploration and devel-
opment efforts. Some companies are more successful than others. Many factors lead to success,
including advanced technology, aggressive management, experience, and serendipity. A signifi-
cant underlying cause of success that is often overlooked or taken for granted, however, is the
quality of subsurface structural and mapping methods. The application of the numerous tech-
niques presented in this book should improve the quality of any subsurface interpretation. This
improved quality should positively affect any company’s economic picture. This is accomplished
by:
1. Developing the most reasonable subsurface interpretation for the area being studied, even
in areas where the data are sparse or absent.
2. Generating more accurate and reliable exploration and exploitation prospects (thereby
reducing associated risk).
3. Correctly integrating geological, geophysical, and engineering data to establish the best
development plan for a new field discovery.
The Philosophical Doctrine of Accurate Subsurface Interpretation and Mapping 3
The following text provides more detail, defining the ten points of the Philosophical
Doctrine.
itself can be confusing with respect to true subsurface relationships. For example, cross sections
and seismic sections can misrepresent true three-dimensional subsurface relationships by the
simple nature of their orientations. All data (well log, seismic, production, paleontologic, etc.)
must be integrated into an interpretation if it is to be considered sound and viable.
Important thoughts to keep in mind while compiling and using subsurface data are the phys-
ical limits and accuracy of these data. A degree of confidence must be built into any final inter-
pretation, but the end result can be only as good as the data that are used. The reasons for failing
to recognize inaccuracies of data can range from blindly accepting auto-correlations to not inves-
tigating questionable directional surveys, mud logs, electric logs, core reports, velocity functions,
production data, and so on. Inclusion of incorrect data may be unavoidable due to inaccuracies
inherent in tools and procedures that generate data or to historical methods of record keeping.
Many times, however, it is possible to determine whether the data are out of date, incomplete, or
subject to error.
Questionable data may represent the only data that are currently available. In other cases,
more reliable data are available and should be sought. The physical limits and possible inaccura-
cies of the data should always be noted when presenting completed work. Always acknowledge
questionable findings and possible discrepancies in a final interpretation. Also, all data used must
be used correctly in order to ensure accuracy in any interpretation.
5. Accurate correlations (well log and seismic) are required for reliable geologic interpre-
tations. An interpretation that properly integrates all data, such as well log, seismic, and pro-
duction data, is always more accurate than an interpretation that ignores one of these sources
(Tearpock and Bischke 1991). Likewise, the correlations must be accurate because geologic
interpretations have their foundation in correct correlations. Consider that all aspects of subsur-
face interpretation and prospecting are based on correlations. These correlations are used to pre-
pare cross sections and fault, unconformity, salt, structure, and isochore/isopach maps.
Workstations have given geoscientists the ability to work immensely large databases in a rela-
tively short time. They have also provided, however, a method of auto-picking correlations.
These correlations are at times blindly accepted and therefore not verified by the interpreter
before plunging into map generation. This can be a blessing as well as a curse to the less knowl-
edgeable.
Eventually, a geoscientist’s correlations, right or wrong, are incorporated into the final inter-
pretation. Incorrect correlations can be costly; they can result in a dry hole, an unsuccessful
workover or recompletion, the purchase of an uneconomic property, or the sale of a producing
property that has significant, unrecognized potential.
6. The use of correct mapping techniques and methods is essential to generate reasonable
and correct subsurface interpretations. Maps and cross sections are in most cases the primary
vehicles used to organize, interpret, and present available subsurface information. The reliability
of any subsurface interpretation presented on maps and cross sections is directly related to the use
of accurate and correct mapping methods and techniques.
Geoscientists who have a good understanding of the mapping methods applicable in the area
of study prepare the most accurate geologic interpretations. There is no substitute for correct
mapping techniques. A poor understanding of mapping techniques can result in incorrect proce-
dures, unjustified short cuts, and costly inaccurate interpretations. We illustrate interpretive con-
touring as an example of the use of a correct mapping technique. Interpretive contouring is the
most acceptable method of contouring subsurface features (Bishop 1960). Unlike other contour-
ing methods, interpretive contouring allows the geoscientist to use knowledge of the structural
and depositional style in the tectonic setting being worked, the ability to think in three dimen-
6 Chap. 1 / Introduction to Subsurface Mapping
sions, field and work experience, imagination, and geologic license to generate an interpretation
that is geologically sound.
In today’s world of computer-contoured maps, can you, as a geoscientist, define which grid-
ding or triangulation procedure should be used to generate a map that honors the data and reflects
a mapped interpretation in a way similar to a hand-contoured map made by an experienced geo-
scientist? This is one example of a technique with which a geoscientist must be familiar in order
to generate a good subsurface interpretation.
7. All important and relevant geologic surfaces must be mapped and the maps integrated to
arrive at a reasonable and accurate subsurface picture. These include surfaces such as tops of for-
mations, stratigraphic markers, faults, unconformities, and salt. For example, in faulted areas it is
typically the faults that form the structures, such as rollovers, fault bend folds, and fault propa-
gation folds. To develop a good understanding of any faulted structure, one must analyze, inter-
pret, and map the faults. We cannot overemphasize the importance of mapping faults. In addition,
faults play an important role in both the migration and trapping of hydrocarbons. Therefore, a rea-
sonable structural interpretation in most faulted areas is dependent upon an accurate three-dimen-
sional understanding of the faults in the area.
The next step in the interpretation process is to integrate the faults with the structure to arrive
at an accurate interpretation. If you want to drill more than your share of dry holes, don’t map faults.
8. The mapping of multiple horizons is essential to develop reasonably correct three-dimen-
sional interpretations of complexly faulted areas. It allows the interpreter to establish a three-
dimensional structural framework prior to generating prospects. The mapping of at least three
horizons (shallow, intermediate, and deep) allows the geoscientist to determine if the interpreta-
tion is plausible and fits at all levels from shallow to deep. Remember, almost any set of fault and
structural data can be forced to fit on one horizon. The true test of an interpretation is to have the
data fit at all structural levels. Therefore, no prospect should be accepted for review without ver-
ification that multiple levels have been interpreted and mapped. The multiple horizon mapping
must show that the interpreted structural framework is geologically sound and that it conforms to
three-dimensional spatial and temporal relationships.
9. Balanced cross sections are required to prepare a reasonably correct interpretation of
complexly deformed structures. In complexly deformed areas the application of sound structural
methods and the generating of structurally balanced cross sections are necessary to prepare
admissible interpretations. Structural interpretations are not cast in stone. Any given structure was
not always as it is today. If an interpreted structure does not volume-balance or a cross section
does not area-balance, then the interpretation cannot be correct from a simple geometric point of
view. Thus, balancing can direct the interpreter toward the correct interpretation. Structural bal-
ancing provides a better understanding of the present configuration of structures, how structures
form, how, when and where fluids may have entered the structures, and where hydrocarbons may
presently exist. Structural balancing should be an integral part of geologic interpretations in com-
plex areas.
10. All work should be documented. Significant volumes of data are collected, evaluated,
used, and manipulated during a project. Good, accurate recording of data and a description of
procedures taken by an interpreter make everyone’s work go more smoothly and accurately. The
documentation of work is an integral part of that work. All subsurface projects will provide bet-
ter results if the data collected and generated are recorded in some format that can easily be ref-
erenced, used, or revised. These data include information on formation tops and bases, faults,
unconformities, net sand counts, correction factors, and more. All completed work needs to be
supported by all the raw data, whether obtained from commercial services or internally generated.
Types of Subsurface Maps and Cross Sections 7
Many persons, including the interpreter, supervisors, managers, other members of an organ-
ized study team, or persons inheriting the area of study, may at some time need to review the sub-
surface data. Numerous types of data sheets are available for documenting geological activities.
Everyone should become familiar with the forms used in their company and use them on a regu-
lar basis.
The Philosophical Doctrine presented here has been employed by many successful geoscientists.
Hydrocarbons are not found by luck, serendipity, or guesswork, but instead by solid scientific work.
Historical data show that those geoscientists who practice this philosophy on a regular basis are more
successful finders and developers of hydrocarbons than those who do not. Workstations and comput-
ers have enabled geoscientists to evaluate more data in a short amount of time and to apply the phi-
losophy and methods faster than by hand. However, we must remember that people, not workstations,
find oil and gas, and the philosophy and methods used have a great impact on success.
SUBSURFACE MAPS. Structure, structural shape, porosity top and base, fault surface,
unconformity, salt, net sand, net hydrocarbon, net oil, net gas, isochore, interval isopach,
facies, and palinspastic maps.
CROSS SECTIONS. Structure, stratigraphic, problem-solving, final illustration, balanced,
and correlation sections. Conventional and isometric fence diagrams and three-dimensional
models are also presented.
CHAPTER 2
CONTOURING
AND CONTOURING
TECHNIQUES
INTRODUCTION
A wide variety of subsurface maps are discussed in the following chapters. Each map presents a
specific type of subsurface data obtained from one or more sources. The purposes of these maps
are to present data in a form that can be understood and used to explore for, develop, or evaluate
energy resources such as oil and gas.
It might seem elementary to have a chapter on contouring since most geologists are taught
the basics of contouring in several introductory geology courses. However, there are two good
reasons for this chapter. First, part of our audience includes members of the geophysical and
petroleum engineering disciplines. They may have had little, if any, training in basic contouring
principles and methods. Second, because the understanding and correct application of contour-
ing and contouring techniques is of paramount importance in establishing a solid foundation in
subsurface mapping, a review of contouring is appropriate.
The majority of subsurface maps use the contour line as the vehicle to convey the various
types of subsurface data. By definition, a contour line is a line that connects points of equal value.
Usually this value is compared to some chosen reference, such as sea level in the case of struc-
ture contour maps. In preparing subsurface maps, we are dealing with data beneath the earth’s
surface, which cannot be seen or touched directly. Therefore, the preparation of a geologically
reasonable subsurface map requires an in-depth knowledge of geology, interpretation skills,
imagination, an understanding of three-dimensional geometry, and the use of correct mapping
techniques.
Any map that uses the contour line as its vehicle for illustration is called a contour map. A
contour map illustrates a three-dimensional surface or solid in two-dimensional plan (map) view.
Any set of data that can be expressed numerically can be contoured.
8
Three-Dimensional Perspective 9
The following list shows examples of contourable data and the associated contour map.
Data Type of Map
Elevation Structure, Fault, Salt
Thickness of sediments Interval Isopach
Percentage of sand Percent Sand
Feet or meters of pay Net Pay Isochore
Pressure Isobar
Temperature Isotherm
Lithology Isolith
If the same set of data points to be contoured is given to several interpreters, the individual-
ly contoured maps generated would likely be different. Differences in an interpretation are the
result of educational background, the amount of geological training, field and work experience,
imagination, and interpretive abilities (such as visualizing in three dimensions). Yet the use of all
the available data and an understanding and application of the basic principles and techniques of
contouring should be the same. These principles and actual techniques are fundamental to the
construction of a mechanically correct map.
In the first part of this chapter, the importance of visualizing in three dimensions and the
basic rules of contouring are discussed. In addition, various techniques for contouring by hand
are illustrated and certain important guidelines identified. Later, computer-based contouring is
discussed.
THREE-DIMENSIONAL PERSPECTIVE
In this section, we show how three-dimensional surfaces are represented by contours in map
view. A good understanding of the geometry within our subsurface geological world, tectonics,
and the principles of three-dimensional spatial relationships are essential to any attempt at con-
structing a picture of the subsurface. Some geoscientists have a stronger educational background
in the geological sciences than others. In addition, some geoscientists have an innate ability to
visualize in three dimensions, whereas others do not. One of the best ways to develop this abili-
ty is to practice perceiving objects in three dimensions.
In its simplest form, we can view a plane dipping in the subsurface with respect to the hor-
izontal. (The reference datum for all the examples shown is sea level.) Figure 2-1 shows an iso-
metric view of a plane dipping at an angle of 45 deg with respect to the horizontal and a projec-
tion of that dipping plane upward onto a horizontal surface to form a contour map. This dipping
plane intersects an infinite number of horizontal planes; but, for any contour map, only a finite
set of evenly spaced horizontal plane intersections can be used to construct the map. (Example:
For a subsurface structure contour map, the intersections used may be 50 ft or m, 100 ft or m, or
even 500 ft or m apart.) By choosing evenly spaced finite values, we have established the contour
interval for the map.
Next, it is important to choose values that are easy to use for the contour lines. For example,
if a 100-ft contour interval is chosen, then the contour line values selected to construct the map
should be in even increments of 100 ft, such as 7000 ft, 7100 ft, and 7200 ft. Any increment of
100 ft could be chosen, such as 7040 ft, 7140 ft, and 7240 ft. This approach, however, makes the
map more difficult to construct and harder to read and understand. In Fig. 2-1, a 100-ft contour
interval was chosen for the map (the minus sign in front of the depth value indicates the value is
below sea level). The intersection of each horizontal plane with the dipping plane results in a line
of intersection projected into map view on the contour map above the isometric view. This con-
10 Chap. 2 / Contouring and Contouring Techniques
Figure 2-2 Isometric view of a curved surface intersecting a finite number of evenly spaced hor-
izontal planes. (Modified from Appelbaum. Geological & Engineering Mapping of Subsurface: A
workshop course by Robert Appelbaum. Published by permission of Prentice-Hall, Inc.)
Three-Dimensional Perspective 11
Figure 2-3 The spacing of contour lines is a function of the shape and slope of the surface being contoured.
12 Chap. 2 / Contouring and Contouring Techniques
for gridding, distribution of data, and other factors (refer to section Computer-Based Contouring).
The computer is a tool, as are an engineer’s scale and ten-point spacing dividers. Yes, it is a pow-
erful tool, but nonetheless a tool. We cannot accept computers driving interpretations, nor can we
blindly accept the resultant maps. Our educational background, experience, and geologic princi-
ples should control any interpretation or generated map. The workstations and personal comput-
ers in wide usage today are powerful tools, but they are not artificial intelligence capable of
generating geologically and geometrically reasonable interpretations or maps.
Look at a three-dimensional subsurface formation that is similar in shape to a topographic
elongated anticline and the contour map representing that surface (Fig. 2-4). The contour map
graphically illustrates the subsurface formation in the same manner that a topographic map
depicts the surface of the earth. By using your ability to think in three dimensions, it is possible
to look at the contour map and visualize the formation in its true subsurface three-dimensional
form.
RULES OF CONTOURING
There are several rules that must be followed in order to draw mechanically correct contour maps.
This section lists these rules and discusses a few exceptions.
(a)
(b)
Figure 2-4 A three-dimensional view of an anticlinal structure and the contour map representing the structure.
Rules of Contouring 13
1. A contour line cannot cross itself or any other contour except under special
circumstances (see rule 2). Since a contour line connects points of equal value, it cannot
cross a line of the same value or lines of different values.
2. A contour line cannot merge with contours of the same value or different values. Contour
lines may appear to merge or even cross where there is an overhang, overturned fold, or
vertical surface (Fig. 2-5). With these exceptions, the key word is appear. Consider a
vertical cliff that is being mapped. In map view the contours appear to merge, but in
three-dimensional space these lines are above each other. For the sake of clarity, contours
should be dashed on the underside of an overhang or overturned fold.
3. A contour line must pass between points whose values are lower and higher than its own
value (Fig. 2-6).
4. A contour line of a given value is repeated to indicate reversal of slope direction. Figure
2-6 illustrates the application of this rule across a structural high (anticline) and a
structural low (syncline).
5. A contour line on a continuous surface must close within the mapped area or end at the
edge of the map. Geoscientists often break this rule by preparing what is commonly
Figure 2-5 To clearly illustrate a three-dimensional overhang or overturned fold, dash the
contours on the underside of the structure. (From Tearpock and Harris 1987. Published by
permission of Tenneco Oil Company.)
14 Chap. 2 / Contouring and Contouring Techniques
referred to as a “postage stamp” map. This is a map that covers a very small area when
compared to the areal extent of the structure.
These five contouring rules are simple. If they are followed during mapping, the result will
be a map that is mechanically correct. In addition to these rules, there are other guidelines to con-
touring that make a map easier to construct, read, and understand.
1. All contour maps should have a chosen reference to which the contour values are
compared. A structure contour map, as an example, typically uses mean sea level as the
chosen reference. Therefore, the elevations on the map can be referenced as being above
or below mean sea level. A negative sign in front of a depth value means the elevation is
below sea level (e.g., –7000 ft).
2. The contour interval on a map should be constant. The use of a constant contour interval
makes a map easier to read and visualize in three dimensions because the distance
between successive contour lines has a direct relationship to the steepness of slope.
Remember, steep slopes are represented by closely spaced contours and gentle slopes by
widely spaced contours (see Fig. 2-3). If for some reason the contour interval is changed
on a map, it should be clearly indicated. This can occur where a mapped surface contains
both very steep and gentle slopes, such as those seen in areas of salt diapirs. The choice
of a contour interval is an important decision. Several factors must be considered in
making such a choice. These factors include the density of data, the practical limits of
Figure 2-6 A contour line must be repeated to show reversal of slope direction. (From Tearpock and Harris
1987. Published by permission of Tenneco Oil Company.)
Rules of Contouring 15
data accuracy (i.e., directional surveys), the steepness of slope, the scale of the map, and
its purpose. If the contour interval chosen is too large, small closures with less relief than
the contour interval may be overlooked. If the contour interval is too small, however, the
map can become too cluttered and reflect inaccuracies of the basic data.
3. All maps should include a graphic scale (Fig. 2-7). Many people may eventually work
with or review a map. A graphic scale provides an exact reference and gives the
reviewer an idea of the areal extent of the map and the magnitude of the features shown.
Also, it is not uncommon for a map to be reproduced. During this process, the map may
be reduced or enlarged. Without a graphic scale, the values shown on the map may
become useless.
4. Every fifth contour should be thicker than the other contours, and it should be labeled with
the value of the contour. This fifth contour is referred to as an index contour. For
example, with a structure contour map using a 100-ft contour interval, it is customary to
thicken and label the contours every 500 ft. And at times it may be necessary to label
other contours for clarity (Fig. 2-7).
5. Hachured lines should be used to indicate closed depressions (Fig. 2-7).
6. Start contouring in areas with the maximum number of control points (Fig. 2-8).
7. Construct the contours in groups of several lines rather than one single contour at a time
(Fig. 2-8). This should save time and provide better visualization of the surface being
contoured.
8. Initially, choose the simplest contour solution that honors the control points and
provides a realistic subsurface interpretation.
9. Use a smooth rather than undulating style of contouring unless the data indicate
otherwise (Fig. 2-9).
10. Initially, a hand-drawn map should be contoured in pencil with the lines lightly drawn so
they can be erased as the map requires revision.
11. If possible, prepare hand-contoured maps on some type of transparent material such as
mylar or vellum. Often, several individual maps have to be overlaid one on top of the
other (see Chapter 8, Structure Maps). The use of transparent material makes this type
of work easier and faster.
Figure 2-8 Begin contouring in areas of maximum control using groups of contour lines.
Unlike topographic data, which are usually obtainable in whatever quantity needed to con-
struct very accurate contour maps, data from the subsurface is scarce. Therefore, any subsurface
map is subject to individual interpretation. The amount of data, the areal extent of that data, and
the purpose for which a map is being prepared may dictate the use of a specific method of con-
touring. There are four distinct methods of hand contouring. These methods are (l) mechanical,
(2) equal-spaced, (3) parallel, and (4) interpretive (Rettger 1929; Bishop 1960; and Dennison
1968).
1. Mechanical Contouring. By using this method of contouring, one may assume that the
slope or angle of dip of the surface being contoured is uniform between points of control and that
any change occurs at the control points. Figure 2-10 is an example of a mechanically contoured
map. With this approach, the spacing of the contours is mathematically (mechanically) propor-
tioned between adjacent control points. We use line A-A′ in Figs. 2-10 and 2-11 to illustrate this
method of contouring. Wells No. 2, 4, and 5 lie on line A-A′ with depths of 600 ft, 400 ft, and
200 ft respectively. The contour interval used for this map is 100 ft. First, we can see that the
Methods of Contouring By Hand 17
(a)
600-ft, 400-ft, and 200-ft contour lines pass through Wells No. 2, 4, and 5. Next, we need to deter-
mine the position of the 500-ft and 300-ft contour lines. Remembering that this method assumes
a uniform slope or dip between control points, we can use ten-point spacing dividers or an engi-
neer’s scale to interpolate the location of these two contour lines. The 500-ft contour line lies
midway between the 600-ft and 400-ft contour lines. Likewise, the 300-ft contour line is placed
midway between the 400-ft and 200-ft contour lines. When this procedure is repeated for all adja-
cent control points, the result is a mechanically contoured map that is geometrically accurate.
Mechanical contouring allows for little, if any, geologic interpretation. Even though the map
is mechanically correct, the result may be a map that is geologically unreasonable, especially in
areas of sparse control.
Although mechanical contouring is not recommended for most contour mapping, it does
have application in a few areas and may be a good first step when beginning work in a new geo-
graphic area. When there is a sufficient amount of seismic or well control, such as in a densely
drilled mature oil or gas field, this method may provide reasonable results, since there is little
18 Chap. 2 / Contouring and Contouring Techniques
Figure 2-10 Mechanical contouring method. (Modified from Bishop 1960. Published
by permission of author.)
room for interpretation. This method is at times employed in litigation, equity determinations,
and unitization because it supposedly minimizes individual bias in the contouring. However,
although individual bias may be minimized, the method does not allow for true geological inter-
pretation. The method is therefore not recommended for the activities listed here.
2. Parallel Contouring. With this method of contouring, the contour lines are drawn paral-
lel or nearly parallel to each other. This method does not assume uniformity of slope or angle of
dip as in the mechanical contouring method. Therefore, the spacing between contours may vary
(Fig. 2-12).
As with the previous method, if honored exactly, parallel contouring may yield an unrealis-
tic geologic picture. Figure 2-13 shows a map that has been contoured using this method. Notice
that the highs appear as bubble-shaped structures with the adjoining synclines represented as
sharp cusps. This map depicts an unreasonable geologic picture.
Although this method may yield an unrealistic map, it does have several advantages over
mechanical contouring. First, the method allows some geologic license to draw a more realistic
map than one constructed using the mechanical method, because there is no assumption of uni-
Figure 2-12 Parallel contouring method. (Modified from Bishop 1960. Published by
permission of author.)
form dip. Also, this method is not as conservative as true mechanical contouring. Therefore, it
may reveal features that would not be represented on a mechanically contoured map.
the result of maintaining a constant dip rate or slope. The advantage to this method, in the early
stages of mapping, is that it may indicate a maximum number of structural highs and lows expect-
ed in the study area. One assumption that must be made in using this method is that the data used
to establish the slope or rate of dip are not on opposite sides of a nose or on opposite flanks of a
fold. In Fig. 2-14, Wells No. 6 and 8 are on the northern flank of this structure; therefore, neither
well can be used with a well on the southern flank to establish the rate of dip. These two wells
can be and were used to establish the rate of dip for contouring the northern flank of this south-
east trending structural nose.
This method does have application on the flanks of structures that have a uniform dip, such
as kink band folds. For example, the back limb of a fault bend fold, which typically is parallel to
the dip of the thrust fault that formed the fold, may have a constant dip between axial surfaces
(Fig. 10-33). In such a case, the equal-spaced method of contouring may have some applicabili-
ty in contouring this back limb.
Interpretive contouring is the most acceptable and the most commonly used method of hand con-
touring.
As mentioned earlier, the specific method chosen for contouring may be dictated by such
factors as the number of control points, the areal extent of these points, and the purpose of the
map. It is essential to remember that no matter which method is used in making a subsurface map,
the map is not correct. No one can really develop a correct interpretation of the subsurface with
the same accuracy as that of a topographic map. What is important is to develop the most rea-
sonable and realistic interpretation of the subsurface with the available data, whether the maps
are constructed by hand or with the use of a computer.
As an exercise in contouring methods, use vellum to contour the data points in Fig. 2-16,
using all four methods, and compare the results. Which method results in the most optimistic pic-
ture? The most pessimistic?
Special guidelines are used in contouring fault, structure, and isochore maps. Additional
guidelines for these maps are discussed in the appropriate chapters. When using computers for
mapping, there are other guidelines that should be used or at least considered.
22 Chap. 2 / Contouring and Contouring Techniques
Figure 2-15 Interpretive contouring method. (Modified from Bishop 1960. Published
by permission of author.)
Figure 2-16 Data points to be contoured using all four methods of contouring.
(Reproduced from Analysis of Geologic Structures by John M. Dennison, by permission
of W. W. Norton & Company, Inc. Copyright 1968 by W. W. Norton & Company, Inc.)
Computer-Based Contouring Concepts and Applications 23
Surface Modeling
We start from a table of X, Y, and Z, where X and Y define the locations of our “samples” and Z
is the “height” of the surface at that location. The problem is to position the contour lines so as
to depict a reasonable geologic surface. Mathematically, this is an interpolation problem.
It can be said that interpolation is the mathematical art of estimating. An interpolation
process calculates the value (height) of a surface at locations where it is unknown, based upon
the values at locations where it is known. Interpolation is an art in the sense that there is no limit
to the number of mathematical formulae that may be conceived to make the estimates, and the
choice of formulae includes subjective and aesthetic criteria: Does the map look geologically rea-
sonable? Does it come close to how you would do it by hand? Is it pleasing?
The great mathematicians of the past, Newton, Sterling, Chebyshev, Gauss, etc., who dealt
with interpolation methods, did not address (or at least did not publish) satisfactory methods to
deal with three-dimensional, randomly distributed data. This fact, plus the subjective and aes-
thetic requirements of contour maps, explains the proliferation of computer contouring methods
that have evolved in the past few decades. Two approaches to deal with random data distribution
have emerged: indirect (gridded) and direct (nongridded).
Indirect Technique (Gridding). Computer contouring methods that use the gridding tech-
nique typically start out by using the primary (original) data points to generate a set of second-
ary (calculated) points. These secondary points, which are traditionally along an orderly geo-
metric pattern (grid), then replace the primary data in steps that generate the contour maps. All
subsequent calculations to locate the positions of the contour lines use the secondary data set (the
24 Chap. 2 / Contouring and Contouring Techniques
grid). The purpose of using this technique is to simplify the subsequent steps by making the
geometry more manageable.
It is easy to ensure that any contour line drawn through a data grid honors its given grid of
data points. However, contour lines that honor the data grid cannot be guaranteed to honor the
original (primary) data points.
Direct Technique (Triangulation). The triangulation technique is the most common of the
direct contouring techniques that interpolate values along a pattern which need not be regular but
which is derived from the pattern of the original data. The pattern includes the locations of the
original data, which are kept throughout the subsequent processing, thus providing the opportu-
nity that all contour lines will honor all the original data.
For both techniques, gridding and triangulation, many ways exist to solve the basic problem.
Our goal is to choose a technique that most nearly fulfills the needs of the user: geologist, geo-
physicist, or engineer. In general, the computer-contoured map is more acceptable to the user if
it is geologically reasonable and looks as if it has been contoured by hand.
These last two steps are where numerous schemes have been developed to make maps that
are aesthetically pleasing and that honor the data points as much as possible.
Selecting Neighbors and Estimating Values at Grid Nodes. We now describe some of
the methods used for selecting neighbors and estimating values at grid nodes. When selecting
neighbors, two criteria are important: (1) neighbors should be evenly distributed around the grid
node, and (2) only data points near the grid node should be considered neighbors. As we discuss
some of the methods for selecting neighbors, certain data distributions will make it difficult to
meet the above criteria. We now describe several (out of many) schemes to select neighbors.
1. Nearest “n” Neighbors. This is the simplest method of selecting neighbors (Fig. 2-17a).
Although this method gives acceptable results for evenly distributed data, it has problems
when applied to an uneven data set. For example, where data have been gathered along
a straight line (e.g., 2D seismic data, as in Fig. 2-17b), all data points may lie on one side
of the grid node. Linear data are unsuitable for surface fitting. To insure that data are
chosen on all sides of a grid node, the domain around the node may be broken into sectors,
or segments, such as quadrant, octant, or other pie segments. One method of ensuring a
better distribution of data around each grid node is by sector search. Eight sectors are
used in Fig. 2-18. In the sector search, a certain number, “n,” of nearest neighbors are
selected from each segment. Obviously, the neighborhood has to expand in order to find
some neighbors in each segment. This may mean ignoring some nearby control points in
order to satisfy the limitation that only “n” points be taken from each segment.
Computer-Based Contouring Concepts and Applications 25
(a)
(b)
Figure 2-17 (a) “n” nearest neighbors. This is the simplest method of selecting neighbors.
(b) “n” nearest neighbors may all be on one side of grid node for 2D seismic data.
(AAPG©1991, reprinted by permission of the AAPG whose permission is required for
further use.)
26 Chap. 2 / Contouring and Contouring Techniques
One obvious drawback to any of the sector search procedures is their lack of uniqueness; i.e.,
if the data are rotated, neighbors and neighborhoods change. It would be preferable to find a
method that is invariant under rotation.
2. Natural Neighbors. On a flat surface that contains more than two data points, any three
data points that lie on the circumference of a circle that contains no other data points
form what is called a Delauney triangle. These three points are also defined as natural
neighbors.
Estimating Values at Grid Nodes. Once we have selected neighbors of a grid node, we pro-
ceed to use the Z values of these neighbors to estimate a value (and perhaps a slope) at the grid
node. The following are some of the methods used:
Each of these methods (and other schemes) has its advocates and adversaries. Any of them
works well if the data points are well distributed and well behaved. Each method has problems
under certain circumstances.
Computer-Based Contouring Concepts and Applications 27
Most gridded contouring programs give their users an opportunity to select the method of
choosing neighbors and the method for estimating values at grid nodes. All gridded contouring
programs require users to select grid size.
1. Gridding can never guarantee maps that honor all of the data points. On the other hand,
the non-honoring of data may be acceptable if the data are noisy or if the calculated value
and the observed value at a data point location differ by an amount that is within the
accuracy of the data. In fact, it the data are particularly noisy, the maps may be more
pleasing if all of the data are not honored.
2. Sparse data sets that contain clusters of closely spaced data can be troublesome for
computer contouring systems, gridded or nongridded. An example of such clustered data
distribution is in oil and gas exploration areas, which include wildcat areas (sparse data)
and some oil and gas fields (clustered data).
3. Changing the grid size often produces a different map because the neighbors of grid nodes
change with changes in grid size.
4. The user must choose a method of selecting neighbors and a method of estimating values
at grid nodes (interpolation).
1. The X-Y data set containing “n” data points has been broken down into a set of (2n-m-2)
triangles, where “m” is the number of data points on the convex hull. The convex hull
is defined as the smallest convex polygon that encloses all of the data.
2. If the triangles are Delauney triangles, they are as nearly equilateral as possible for any
data distribution, and they are invariant under rotation.
3. In large data sets, each data point will have an average of six natural neighbors.
Figure 2-19 is a sample data set with Delauney triangles connecting natural neighbors. If no
interpolation were performed, and if the surface being contoured were treated as flat in each tri-
angle, the map would have the characteristics of mechanical contouring and would look like Fig.
2-20.
Mechanical contouring, however, tends to be very angular and unrealistic. It is usually nec-
essary to interpolate values within the Delauney triangles, i.e., create more and smaller triangles,
based on curved mathematical models in order to overcome the angular appearance. Figure 2-21
shows a set of the subtriangles formed when each leg of the original Delauney triangles is divid-
ed into 16 segments.
28 Chap. 2 / Contouring and Contouring Techniques
Figure 2-19 Delauney triangles and natural Figure 2-20 Mechanical contouring in basic
neighbors. (AAPG©1991, reprinted by permis- Delauney triangles. (Published by permission
sion of the AAPG whose permission is of Scientific Computer Applications, Inc.)
required for further use.)
1. Triangulation always honors every data point because the original data points always
remain in the data set being contoured.
2. The interpretation is essentially the same, regardless of the number of triangles or
smoothing.
3. The user does not have to worry about data distribution.
Sample Data Set. Figure 2-22 shows a small (13-point) data set representing a structural sur-
face. We encourage you to contour these data yourself before reviewing the computer-contoured
maps. These were the data used to generate Figs. 2-20 and 2-21. Figure 2-23a through f are mon-
tages of contour maps of these data generated using various gridded contouring programs, which
in turn used various methods of (1) selecting neighbors and (2) calculating values at grid nodes.
Figure 2-24 is a contour map of these same data generated using Delauney triangles, where a
value was interpolated at each vertex of the subtriangles shown in Fig. 2-21.
Reviewing all these gridded and triangulated maps of the same data set shows that vast dif-
ferences exist in map interpretation provided by various methods.
Figure 2-21 Subtriangles forming each side of Figure 2-22 Sample data set. (AAPG©1991,
Delauney triangles are divided into 16 segments. reprinted by permission of the AAPG whose per-
(AAPG©1991, reprinted by permission of the AAPG mission is required for further use.)
whose permission is required for further use.)
seismic wavelet. Most of these surfaces are related to each other, for they represent stratigraphic
units deposited progressively on a relatively flat sea floor. So we need to address multi-surface
contouring.
In nature, many associated surfaces resemble one another. When material is deposited on an
old surface, the resulting new surface tends to exhibit the same or similar features as the old sur-
face, although the new features tend to be attenuated. Forces may reshape the structure after dep-
osition. Deformation applied to suites of surfaces results in similar features for neighboring sur-
faces that reflect their common history. It is intuitively expected and empirically verifiable that
under these circumstances, the true vertical thickness (isochore) between adjacent surfaces tends
to be less complicated than the surfaces themselves. The process by which we take advantage of
these facts to get improved interpretation is called stacking, or conformable mapping. When done
by hand, the procedure is usually as follows:
1. Contour the shallowest structure, which is usually the one that has the most data. Designate
it as Surface 1.
2. Prepare an isochore contour map of the interval between Surface 1 and Surface 2, using
all the data points that penetrate Surfaces 1 and 2.
3. Create and contour estimated structural points on Surface 2. First, on a light table, overlay
the Surface 1 structure map and the interval isochore map. Then overlay the Surface 2 base
map on those maps. Using points where Surface 1 structure contours cross the interval
isochore contours, add the thickness to the depth of Surface 1 and plot the calculated depth
for each of those points on Surface 2. Then contour the structure on Surface 2.
30 Chap. 2 / Contouring and Contouring Techniques
(a) (b)
(c) (d)
Figure 2-23 (a) - (f) Six structural contour maps using various gridding methods for the same data set.
(Published by permission of Scientific Computer Applications, Inc.)
Computer-Based Contouring Concepts and Applications 31
(e) (f)
Figure 2-24 Contour map of sample data generated using Delauney triangles. (AAPG©1991, reprinted by
permission of the AAPG whose permission is required for further use.)
32 Chap. 2 / Contouring and Contouring Techniques
Figure 2-25 Cross section showing stacking process for two surfaces. (AAPG©1991, reprinted by
permission of the AAPG whose permission is required for further use.)
To illustrate the value of stacking, consider the case shown in Fig. 2-25, which shows in
cross section the stacking process for two surfaces. Of the 11 wells, only five penetrate both sur-
faces. These are designated by the lowercase letter b. Six wells penetrate only surface A and are
designated by the letter a. If data from the wells that penetrate Surface B were used alone, the
surface would tend to be interpreted as shown by the dotted cross-section line.
Instead, stacking recognizes that there is a true vertical thickness (isochore) value at each of
these b points, and the computer program uses them to calculate (interpolate or extrapolate) esti-
mated true vertical thickness values at all of the a points. These calculated thicknesses are sub-
tracted from the elevation of the known Surface A values to get a calculated elevation of Surface
B at all the a points.
A real case will show the benefit of stacking. Figures 2-26 through 2-28 represent a Top-of-
Unit structure map, a Unit isochore map and a Base-of-Unit structure map for the same 13-point
sample data set that was used earlier. These maps were made using multi-surface stacking. The
isochore map was contoured using the five wells that penetrate the Base-of-Unit, interpolating or
extrapolating as necessary to cover the entire map area. Then the elevations for the Base-of-Unit
at the eight other wells were derived by subtracting the isochore values from the Top-of-Unit ele-
vations. Notice that the Base-of-Unit mimics features of the Top-of-Unit. Highs are shifted in the
direction of thinning isochores.
Figure 2-29 shows the same Top-of-Unit and Base-of-Unit and Unit isochore contours all
plotted on the same map. It illustrates precisely how a geoscientist would generate a Base-of-Unit
map by hand, using the Top-of-Unit map and the Unit isochore map. Note the three-contour
crossing points (e.g., wherever the two surfaces, Top-of-Unit and Base-of-Unit, are 100 feet apart,
that is a point on the 100-foot isochore).
Computer-Based Contouring Concepts and Applications 33
Figure 2-26 Structure contour map on Top-of- Figure 2-27 Unit isochore map. (Published by
Unit. (Published by permission of Scientific permission of Scientific Computer Applica-
Computer Applications, Inc.) tions, Inc.)
Figure 2-28 Structure contour map on Base-of- Figure 2-29 Top-of-Unit and Base-of-Unit maps
Unit with stacking. (Published by permission of and Unit isochore map overlaid. (Published by
Scientific Computer Applications, Inc.) permission of Scientific Computer Applications,
Inc.)
34 Chap. 2 / Contouring and Contouring Techniques
Base-of-Unit has been penetrated by five wells only. If it were processed without the infor-
mation derived from the 13 wells that penetrate the Top-of-Unit and the knowledge that the sur-
faces are similar, i.e., if it were processed without stacking, we would get a structure contour map
as shown in Fig. 2-30. If we proceed to make a Unit Isochore by subtracting Fig. 2-30 from Fig.
2-26, it would be as shown in Fig. 2-31. Note that all Unit isochore values are honored, but the
interpretation is much different than that shown in Fig. 2-28 and is unreasonable if Top-of-Unit
and Base-of-Unit are conformable surfaces. The value of stacking is clearly seen when maps with
and without stacking are compared.
Figure 2-30 Structure contour map on Base-of- Figure 2-31 Unit isochore map without stacking.
Unit without stacking. (Published by permission (Published by permission of Scientific Computer
of Scientific Computer Applications, Inc.) Applications, Inc.)
Computer-Based Contouring Concepts and Applications 35
Figure 2-32 Example of trace fault. (Published Figure 2-33 Example of contouring using fault polygons.
by permission of Scientific Computer Applica- (Published by permission of Scientific Computer Applica-
tions, Inc.) tions, Inc.)
Note that fault vertical separation is implicit from the data and changes radically. Also note that
shape is not copied across the trace fault.
Another procedure that is used to allow computers to handle faulted surfaces is based on
fault polygons. In this procedure the faulted surface is divided into a series of polygons that
describe individual fault blocks. Data in each fault block are contoured separately, one surface at
a time. Fault vertical separation is implicit and is not treated as an explicit variable. Figure 2-33
is an example of a map contoured using fault polygons.
Another procedure for contouring faulted surfaces on the computer is known as the restored
surface method (fault/structure map integration), which is further discussed in Chapters 7 and 8.
It is based on contouring both the fault surfaces and their vertical separations. Hence vertical sep-
aration (missing or repeated section) is explicit rather than implicit. Figure 2-34 shows a faulted
structure map made using the restored surface method and the same data shown in Figs. 2-32 and
2-33. Figure 2-35 is contoured on the fault, which has a constant 100 ft of vertical separation.
In this restored surface method, faulted systems are treated as what they are: sets of three-
dimensional fault blocks containing mappable strata, which once were continuous surfaces. The
boundaries of these fault blocks are the fault surfaces, and they are contourable. The restored sur-
face method is essentially the procedure that is used when contouring faulted surfaces by hand.
In the restored surface method, faulted systems are processed in three steps designed to
honor continuity of shape across faults:
1. Move (i.e., restore palinspastically) the fault blocks to their pre-faulted positions,
together with the contained geologic horizons.
36 Chap. 2 / Contouring and Contouring Techniques
Figure 2-34 Faulted structure using restored sur- Figure 2-35 Fault surface for 100-ft fault. (Published
face method. (Published by permission of Scientific by permission of Scientific Computer Applications,
Computer Applications, Inc.) Inc.)
2. Having restored the “continuous surface” attribute to the geologic horizons, perform all the
stacking (discussed earlier) and interpolations needed to obtain a smooth map or cross
section.
3. Rebreak (i.e., reverse the first step) and return the fault blocks and their contents to their
faulted positions and display contour maps or cross sections.
Procedure for Contouring Faulted Surfaces. To accomplish the steps of the restored sur-
face method, certain data are needed. These data consist of:
The missing section, or vertical separation, to the computer, is just another mathematical surface,
which can vary over the mapped area and can be contoured. Vertical separation is designated as
positive for normal faults and negative for reverse faults.
It is valuable to make a mental picture of the faulting process. In order to restore the fault
blocks to their pre-faulted position successfully, it is helpful to have a reasonable hypothesis
about the order in time of the various faulting events, since the most successful restoration to
unfaulted positions needs to be done in steps in reverse order to the original faulting. The fault-
ed system normally is analyzed first by making contour maps and/or cross sections of all fault
surfaces in order to:
Computer-Based Contouring Concepts and Applications 37
1. Test the faults for reasonableness: Do observed fault cuts assigned to the same fault result
in a fault surface that makes geologic sense?
2. Infer and/or ascertain the hierarchy of the faults with respect to age. Where two faults meet
in space and do not cross one another, which one extends beyond that junction? Which one
is therefore older? The analysis lends itself to “what if” games regarding the faulting
sequence.
To perform its task, the computer can use a set of restore commands, each of which
describes a fault block and instructs the program to move it to its prefaulted or restored position.
Restore commands take us “backward in time.” The first restore reverses the most recent faulting
event; the last restore reverses the oldest faulting event. Geologic knowledge must be used to
make decisions regarding fault analysis.
We use as an example a sand unit offset by a bifurcating fault system, and we create structure
maps of the top and base of the unit. The system consists of two merging faults, Fault A and Fault
B, contoured in Figs. 2-36 and 2-37. The maps also show the line of bifurcation where Fault B
merges with Fault A. Data for these faults were obtained by correlating logs, picking horizon tops,
locating faults, and measuring missing sections (vertical separations). Fault A has observed cuts
on both sides of the line of bifurcation and Fault B does not, so we conclude that Fault A is the
older fault and Fault B is the younger fault. Hence Fault B will be restored first, and then Fault A
will be restored. Also, we note that vertical separation of Fault A east of the line of bifurcation is
the sum of the vertical separations of Faults A and B west of the line of bifurcation because Fault
B no longer exists. Vertical separation balance must be maintained around the line of bifurcation.
Next, we restore the fault blocks to their original unfaulted positions and then contour the
paleosurface. Figures 2-38 through 2-40 display conceptually the steps taken by the computer
program in restoring the faults. Notice the changes in well depths from figure to figure. Figure 2-
38a shows the depths of the Top-of-Unit in the three fault blocks, as separated by the approxi-
mate locations of Fault A and Fault B. Figure 2-38b is a north-south cross section before any
restoration has occurred. Figure 2-39a is a base map on the Top-of-Unit after Fault B has been
restored, and Fig. 2-39b is a cross section after Fault B has been restored.
Next, we restore the block moved during fault event A (the block downthrown to Fault A),
which incidentally contains restored Fault B. That restoration produces an unfaulted surface,
which is the paleosurface prior to all faulting. We then contour that surface. Figure 2-40a is a
structure map of the Top-of-Unit that was contoured after Fault A was restored. Figure 2-40b
shows the paleosurfaces in cross section.
Now that the faulted system has been restored to its pre-faulting configuration, the structur-
al surfaces are continuous and can be stacked, as described earlier in this chapter.
The final step in processing faulted surfaces is to rebreak and move geologic surfaces to their
true (post-faulted) positions. This step is the mathematical inverse of restoration. In the restored
surface method, the computer program generates hanging wall and footwall fault traces as the
intersection between structural horizons and faults. Rigorous vertical separation balance should
be maintained at all fault intersections. In our example, the structure contour maps for Top-of-
Unit and Base-of-Unit are completed and shown in Fig. 2-41a and b.
Limitations. Limitations of the restored surfaces method as implemented by the computer have
their origin in mathematical simplifications, the most important of which is the use of purely ver-
tical movement in the fault restoration process. This implies that the method becomes less appli-
cable when fault movement is essentially horizontal, for example, strike-slip or very low-angle
thrust faults.
38 Chap. 2 / Contouring and Contouring Techniques
Figure 2-36 Fault surface map of Fault A with vertical separation (missing section) posted. Line
of bifurcation with Fault B is shown. (Published by permission of Subsurface Consultants &
Associates, LLC.)
Figure 2-37 Fault surface map of Fault B with vertical separation (missing section) posted. Line
of bifurcation with Fault A is shown. (Published by permission of Subsurface Consultants &
Associates, LLC.)
Computer-Based Contouring Concepts and Applications 39
(a)
(b)
Figure 2-38 (a) Base map with Top-of-Unit elevations and approximate traces of Faults A
and B. (b) North-south cross section before restoration of faults. (Published by permission
of Subsurface Consultants & Associates, LLC.)
40 Chap. 2 / Contouring and Contouring Techniques
(a)
(b)
Figure 2-39 (a) Base map of Top-of-Unit after Fault B has been restored. (b) North-south
cross section after Fault B has been restored. (Published by permission of Subsurface
Consultants & Associates, LLC.)
Computer-Based Contouring Concepts and Applications 41
(a)
(b)
Figure 2-40 (a) Structure map of Top-of-Unit contoured after Fault A has been restored. This is
the paleosurface prior to faulting. (b) North-south cross section after Fault A has been restored.
(Published by permission of Subsurface Consultants & Associates, LLC.)
42 Chap. 2 / Contouring and Contouring Techniques
(a)
(b)
Figure 2-41 Completed structure maps of (a) Top-of-Unit, and (b) Base-of-Unit, after sur-
faces have been moved to their true post-faulted positions. (Published by permission of
Subsurface Consultants & Associates, LLC.)
CHAPTER 3
DIRECTIONALLY
DRILLED WELLS
AND DIRECTIONAL
SURVEYS
INTRODUCTION
A directionally drilled or deviated well is defined as a well drilled at an angle less than 90 deg to
the horizontal (Fig. 3-1). Wells are normally deviated intentionally in response to a predetermined
plan; however, straight holes often deviate from the vertical due to bit rotation and natural devi-
ation tendencies related to rock types and structure.
The technique of controlled directional drilling began in the late 1920s on the U.S. Pacific
coast (LeRoy and LeRoy 1977). Through the use of controlled directional drilling, a wellbore is
deviated along a preplanned course to intersect a subsurface target horizon(s) at a specific loca-
tion (Fig. 3-2). Our primary interest in directionally drilled wells in this textbook centers around
their application to subsurface mapping.
43
44 Chap. 3 / Directionally Drilled Wells and Directional Surveys
One very important safety application of a deviated well is the drilling of a relief well to kill
a well that has blown out (Fig. 3-3f). There are other applications of deviated wells, but they are
beyond the scope of this textbook and are not discussed.
General Terminology
The terms used to describe various aspects of a directionally drilled well are defined here and
illustrated in Fig. 3-1.
KOP Kick-off point. Depth of initial deviation from vertical measured as measured
depth (MD), true vertical depth (TVD), or subsea true vertical depth (SSTVD).
Build Rate Build angle. Rate at which the angle changes during deviation. It is usually
expressed in degrees per 100 ft drilled. Example: 2 deg per 100 ft.
Ramp Angle Hole angle, drift angle, angle of deviation. Angle from the vertical that a well
maintains from the end of the build through the ramp segment of the well.
BHL Bottom-hole location. Horizontal and vertical coordinates to the total depth
point usually measured from the surface location.
Drop Rate Rate at which the ramp angle changes in degrees per 100 ft. Measured in “S”
shape holes.
Vertical Point The depth where the well is back to vertical, measured as MD, TVD, or
SSTVD.
Horizontal Wells
Horizontal wells are typically considered to be wells with the borehole drilled within about 3 deg
of bed dip or wells drilled nearly horizontally. These include extended reach wells with long hor-
izontal displacements, as well as long and medium radius horizontal wells. Short radius horizon-
tal wells are often called drain hole wells.
Extended-reach wells can be similar to the “S” shape well, but with very high ramp angles
in the 80 deg range. Although they are nearly horizontal, they might not be considered true hor-
izontal wells. These wells are generally drilled when the surface location is necessarily a great
distance from the target. Long-radius horizontal wells have build rates in the 3 deg/100-ft range
and generally the horizontal part of the borehole is several thousand feet in length. Medium
radius horizontal wells have build rates in the 30 deg/100-ft range and are usually drilled for shal-
low objectives. Short radius horizontal wells are borehole segments drilled from a vertical bore-
hole that penetrated the objective interval, with the deviation from vertical to horizontal made
within a vertical interval of about 20 ft. These wells usually have horizontal segments of only a
Common Types of Directionally Drilled Wells 45
1000' KOP = 1000' MD, 1000' TVD 1000' KOP = 1000' MD, 1000' TVD
ANGLE MADE AT 2432' TVD, 2500' MD ANGLE MADE AT 2432' TVD, 2500' MD
4000' 4000'
DROP RATE 1-1/2o/100'
(b)
HORIZONTAL WELL
SURFACE
0'
4000'
(c)
Figure 3-1 Diagrammatic cross section illustrations of (a) a simple ramp or “L” shape well, (b) a more complicated “S”
shape well, and (c) a horizontal well. ([a] and [b] published by permission of Tenneco Oil Company, [c] published by
permission of J. Brewton.)
46 Chap. 3 / Directionally Drilled Wells and Directional Surveys
few hundred feet, but several horizontal segments may be drilled from the same vertical wellbore.
These drain hole wells are used in low permeability reservoirs and enhanced recovery projects.
The purpose of drilling most horizontal wells is to improve the economics of a project by
increasing production rates and shortening well life. For example, horizontal wells can improve
production rates from (1) reservoirs containing heavy oil, such as the Orinoco heavy oil belt
(Venezuela); (2) reservoirs with mostly fracture porosity, such as the Austin Chalk (onshore Gulf
of Mexico); and (3) reservoirs of low permeability. Horizontal wells can also be used to penetrate
and produce from multiple reservoirs that are laterally discontinuous. For example, separate flu-
vial channel sands that are in the same stratigraphic interval but are laterally discontinuous can
be penetrated by a single horizontal well to drain multiple sand bodies. Also, for multilobed reser-
voirs with attic reserves above the highest wells on the structure, a horizontal well can be
designed to encounter all the lobes, whereas a vertical well might miss some of the lobes that
truncate up-dip. Another application of horizontal wells is to drill for thin oil zones over water.
These zones are subject to coning of water into vertical wells. A horizontal well can significant-
ly reduce the problem of water coning. There are other applications of horizontal wells, but most
well plans are based on increasing production rates and shortening well life to improve project
economics.
Figure 3-3 Applications of directional drilling. (a) Multiple wells offshore or from artificial islands. (b) Shoreline
drilling. (c) Fault control. (d) Inaccessible surface location. (e) Stratigraphic trap. (f) Relief well control. (g)
Straightening hole and side tracking. (h, i, j) Saltdome drilling. (From LeRoy and LeRoy 1977. Published by permis-
sion of the Colorado School of Mines.)
Directional Well Plan 47
0o 0' EAST
PLANE OF PROPOSAL PLANE OF PROPOSAL
GRID
S 46o 25' W S 46o 25' W
100' 100' MD 8000' 7000' 6000' 5000' 4000' 3000' 2000' 1000' 0 1000'
900' MD S/L
1749' MD 100'MD
2808' MD
2714' MD 3709' MD 1000'
3802' MD 4621' MD
4000'
5000' MD 5462' MD
2000'
6318' MD 6420' MD
7685' MD
7405' MD 3000'
8771' MD
8494' MD 9996' MD
8000' 9599' MD 11,321' MD 4000'
10,644' MD 12,416' MD
11,719' MD 13,429' MD 5000'
14,514' MD
12,752' MD
13,849' MD 16,000' MD 6000'
12,000' 17,000' MD
14,992' MD
18,000' MD
7000'
19,000' MD
17,000' MD 19,484' MD
B.H.L. 8000'
10,872' 10,873' S-47o 12' 07" W
15,695' 19,484' MD
S - 7388.19' , W - 7976.55'
S 46o 25' W
T.V.D. -15,695'
M.D. -19,484'
0 4000' 8000' 12,000'
(a) (b)
Figure 3-4 (a) Vertical section plan for a directional well. (b) Horizontal plan for the same directional
well shown in Fig. 3-4a. (Published by permission of Gardes Directional Drilling.)
48 Chap. 3 / Directionally Drilled Wells and Directional Surveys
Figure 3-5 Vertical plan for a nearly horizontal well with a maximum deviation angle of 94 deg.
(Published by permission of Gardes Directional Drilling.)
Directional Tools Used for Measurements 49
Figure 3-6 Scaled chart for a build rate of 2 deg per 100 ft of hole drilled.
(Published by permission of Eastman Christensen.)
50 Chap. 3 / Directionally Drilled Wells and Directional Surveys
Drift angle and drift direction are measured by a survey tool conveyed in drill pipe or by wire-
line, and measured depth is determined by length of drill pipe or wireline. The various tools that
are used fall into two categories: magnetic and nonmagnetic.
Magnetic Surveys
Magnetic is a generic term for describing several survey tools that use a magnetic compass for
direction and therefore must be run inside a special nonmagnetic drill collar to negate the effects
of the drill pipe. An example of such a survey is the “single shot” magnetic survey. This device
records, on a heat-resistant film disc, the magnetic direction and inclination angle of the wellbore
at specific depth intervals. A “multishot” survey uses a filmstrip to record several readings of hole
angle and direction at different depth intervals. Newer wireline tools provide real-time survey
data at the surface. Measurement accuracy of magnetic survey tools has been improved within
the last 20 years. Modern magnetic tools are used for measurement-while-drilling (MWD) sur-
veying. These surveys provide real-time data for more efficient directional drilling.
Nonmagnetic Surveys
Nonmagnetic survey tools are of two types: those with no direction-finding device and those with
a gyroscopic mechanism for determining direction. A drift indicator tool (e.g., a Totco tool) meas-
ures only drift angle and is usually run in vertical wells or shallow vertical sections of deviated
wells where directional information is not required. This tool generally consists of a housing or
barrel, a motion indicator, a timer, a punch, and a printed paper disc. The unit is either run on a
wireline or dropped on a drill bit. When the motion sensor determines that the tool is no longer
moving, the timer is activated, and after a predetermined interval of time the punch is released.
The punch, which is allowed to swing freely and act as a plumb bob, drops vertically and punch-
es a hole in the paper disc, which is marked in degrees. Figure 3-7 shows a Totco disc scaled to
a maximum of 8 deg. The hole punched in the disc indicates an inclination angle of 4.5 deg, but
the drift direction is unknown.
Gyroscopic survey tools are widely used and are capable of providing more accurate data
than magnetic survey tools. Because the magnetic compass is replaced by a gyrocompass, the
system can be run in both cased and uncased holes and run where cased holes are nearby, as in a
platform well cluster. The gyro system can be set up as a single or multishot instrument.
Conventional tools, referred to as “free” gyros, are less accurate than the newer “rate” gyros,
which are of different construction. A gyro survey is sometimes run after an MWD magnetic sur-
vey to provide additional information for well path determination. Newer types of gyroscopic
tools are currently being tested, including an MWD gyro tool.
Figure 3-7 Example of a Totco survey. Well deviation angle is 4.5 deg.
you can reduce the wellbore position uncertainty due to calculation method by using the same
algorithm, such as minimum curvature to recalculate all surveys. Computer software for this is
readily available.
Three measurements go into the directional survey: (1) measured depth, (2) deviation angle,
and (3) direction of deviation. These measurements, taken at specific depth intervals, are used to
calculate the directional survey of a well. Figure 3-8 shows a portion of the directional survey
from a deviated well. The tabular printout for this directional survey has nine separate columns
of data for each survey point in the well.
Column Data
1 Subsea depth of wellbore in feet
2 Measured depth of wellbore in feet
3 True vertical depth of wellbore in feet
4 Angle of wellbore deviation (inclination angle)
5 Direction of wellbore (true bearing)
6 Distance in feet from the surface location along the proposed
directional path
7 True bearing and distance of each survey point from the surface
location in rectangular coordinates
8 True bearing and distance from surface location directly to each
survey point
9 Maximum change in hole angle in degrees per 100 ft
52 Chap. 3 / Directionally Drilled Wells and Directional Surveys
If the cased or surface section of the hole was surveyed with a nonmagnetic survey tool such
as a drift indicator, the angle of the cased portion of the hole will be displayed on the survey along
with an estimate of the maximum possible deviation of this portion of the hole, as shown in Fig.
3-9. The 146.75-ft maximum deviation, prominently shown on the survey, indicates that if the
wellbore drift in the cased portion of the hole were all in the same direction, the well at a depth
of 3513 ft could be as much as 146.75 ft from the surface location. Notice at the end of the sur-
vey that it indicates that the bottom of the hole lies within a circle of radius 146.75 ft with its cen-
ter located 203.94 ft south 63 deg 27 min west of the surface location. Such information may be
important in fault, structure, and isochore mapping.
Directional surveys are used to plot wellbores on base maps. The applications of these plots
are discussed later in this chapter.
Figure 3-8 Part of the directional survey for a deviated well. (Published by permission of Gardes Directional Drilling.)
Figure 3-9 A directional survey from a well in which the surface casing was surveyed with a Totco tool providing devia-
tion angle but not direction. Notice, at 3513 ft, that the maximum possible deviation of the cased hole is 146.75 ft. This
assumes that the wellbore deviation in this portion of the hole was in one direction.
Sea, circa 1970, for a variety of reasons. For example, a higher latitudinal position produces more
uncertainty in magnetic surveys. Also, the mathematical model does not apply to horizontal
wells, including extended-reach wells. Many companies have detailed proprietary information
about degree of error in particular types of survey tools, as used under certain operating condi-
tions in given areas.
Tenneco Oil Company conducted a detailed study of directional survey uncertainties in
1980. An important conclusion from the study indicates that there is a 90 percent certainty that
any directional well will have an error of 35 ft or less TVD and 140 ft or less in departure. This
conclusion is drawn irrespective of MD, hole angle, or survey type. This means that wells with
hydrocarbon contacts that vary up to 35 ft TVD may well be in the same reservoir, with the vari-
ations due merely to survey error rather than such geological factors as permeability barriers or
faults.
Today’s directional surveys are potentially more accurate than those described above, pro-
vided that the tools and operating procedures are of high quality. However, that does nothing to
mitigate the errors inherent in surveys of older wells. Always remember that data based on direc-
tional wells, including new wells, should not be taken at face value. Apparent discrepancies in
data used in mapping may be due to errors in depth and lateral position of points in wells. For
example, the depth of a mapped surface or a water contact in a deviated well may not match sat-
54 Chap. 3 / Directionally Drilled Wells and Directional Surveys
VERTICAL UNCERTAINTY
20
C
TI
UNCERTAINTY (± FT. / 1,000 FT. OF HOLE)
NE
10 AG YRO
R
M
FR EE G
OO P OOR
P
IC
5 ET
GN
MA
OD
GO
RO
2 GY
EE
FR O)
D G YR
OO TE
G (RA
TA
1 DA
RO
GY
Figure 3-10 Expected vertical
uncertainty in a deviated well consid-
0.5 ering various types of surveys.
(Modified from Wolff and de Wardt
0 10 20 30 40 50 60 70 80 90 1981. Published by permission of the
Journal of Petroleum Technology
INCLINATION, DEGREES and Gyrodata, Inc.)
isfactorily with depths at nearby points. The reason could be an error in the directional survey or
perhaps in the spotting of the wellbore. The same caution applies even to data discrepancy in a
vertical well, which may have a significant but unrecognized deviation from the vertical.
Survey errors can be corrected in some cases in fields that have hydrocarbon/water contacts.
Wells are often “corrected” to fit the contact. This is usually done by selecting a contact that fits
most of the wells and then adjusting the depth of the well(s) that does not fit. An example of how
to adjust a well is shown in Fig. 3-12. The water contact in Wells No. 1 and 2 is at a depth of
–9738 ft, whereas the contact in Well No. 3 is at a depth of –9748 ft. Since data from two wells
are in agreement with a water contact at –9738 ft, Well No. 3 is adjusted upward 10 ft to correct
the water level from –9748 ft to –9738 ft. Not only is the water level corrected, but the struc-
tural depths of the sand (top and base) are also corrected upward 10 ft. Therefore, the top of the
sand at a depth of –9720 ft becomes –9710 ft. An understanding of directional survey errors can
at times eliminate the need for a “production fault” or “permeability barrier” to explain discrep-
ancies in water levels.
LATERAL UNCERTAINTY
C
TI
100 NE RO
AG GY
M EE
OR FR
PO O R
PO
50
UNCERTAINTY (± FT. / 1,000 FT. OF HOLE)
IC
ET
20 GN
MA
OD
GO RO
GY
EE
FR
D
10
GOO
1
Figure 3-11 Expected lateral uncertainty in a
deviated well considering various types of sur-
vey tools. (Modified from Wolff and de Wardt
0.5
1981. Published by permission of the Journal
of Petroleum Technology and Gyrodata, Inc.)
0 10 20 30 40 50 60 70 80
INCLINATION, DEGREES
The simplified straight-line method of plotting a directional well is shown in Fig. 3-13a. The
only data required and plotted on the base map are the surface location and BHL, which may be
the only data available. The MD to TD may be recorded next to the BHL. A straight dashed line
is usually drawn between the surface location and BHL. This directional well plot provides
absolutely no information about the position or depth of the wellbore in the subsurface between
the surface and bottomhole locations. Such a plot is not helpful in the interpretation, construc-
tion, and evaluation of fault, structure, or isochore maps.
When directional survey data are actually plotted to provide detail as to the lateral position
and subsea depth of the wellbore, as shown in Fig. 3-13b, the plot has real value. Such a plot pro-
vides a visual guide (in map view) to the location and subsea depth of the wellbore anywhere
along its path. It saves time in preparing subsurface maps and is extremely helpful in the inter-
pretation, construction, and evaluation of fault, structure, and isochore maps. Later chapters
examine several important benefits to fault surface and structure mapping derived from plotting
on a base map the actual location and subsea depth, at fixed increments (usually 500 ft or 1000
ft), of all directional wells.
Figure 3-14 illustrates the cross-sectional view of a directionally drilled well and the detailed
map-view plot of the directional data in 500-ft increments of subsea TVD along the actual well
path. If subsea data are to be used in mapping, the directional well data are corrected to subsea
before being plotted on the base map.
Figure 3-15 shows the map-view well plot for the Cognac “A” Platform in the Mississippi
Canyon Block 194 in the U.S. Gulf of Mexico. The platform is the largest multiple-well platform
in the world. The platform is located in 1000 ft of water and has a total height of 1260 ft. Wells
were deviated with high angles, up to 75 deg. Horizontal displacements up to 11,500 ft result in
a well pattern that covers an area with a diameter of more than 4 mi. Total cost for the project
was over $1 billion.
Directional Wells Plots 57
A-2
A-1 A-5
A-6
A-4
0 500’
500' A-3 A-7
SCALE
(a)
0 500’
SCALE
(b)
Figure 3-13 (a) Straight-line method of plotting directional wells in map view. (b) Detailed plot of
directional survey data indicating the location and subsea depth of the wellbores along their
entire length. Compare this plot to that in Fig. 3-13a.
58 Chap. 3 / Directionally Drilled Wells and Directional Surveys
Figure 3-14 Cross-sectional view of a directional well and its detailed map-view plot
in increments of 500 ft.
Directional Wells Plots
Figure 3-15 Spider directional well plot for the Cognac “A” Platform, Offshore
Gulf of Mexico. (Published by permission of Gardes Directional Drilling.)
59
CHAPTER 4
LOG
CORRELATION
TECHNIQUES
INTRODUCTION
Correlation can be defined as the determination of structural or stratigraphic units that are
equivalent in time, age, or stratigraphic position. For the purpose of preparing subsurface inter-
pretations, including maps and cross sections, the two general sources of correlation data are
electric wireline logs and seismic sections. In this chapter, we discuss basic procedures for cor-
relating well logs, introduce plans for performing various phases of well log correlation, and
present fundamental concepts and techniques for correlating well logs from vertical as well as
directionally drilled wells.
Fundamentally, electric well log curves are used to delineate the boundaries of subsurface
units for the preparation of a variety of subsurface maps and cross sections (Doveton 1986).
These maps and cross sections are used to develop an interpretation of the subsurface for the pur-
pose of exploring for and exploiting natural resources, including hydrocarbons.
After the preparation of an accurate well and seismic base map, electric log and seismic cor-
relation work is the next step in the process of conducting a detailed geologic/geophysical study.
No geologic interpretation can be prepared without detailed correlations. Accurate correlations
are paramount for reliable geologic interpretations.
60
Electric Log Correlation Procedures and Guidelines 61
The SSTVD measurement is the only depth measurement from a common reference datum,
sea level. Therefore, SSTVD is the depth most often used for mapping. Logging depths measured
from a vertical or directionally drilled well for mapping are usually corrected to SSTVD. For ver-
tical wells the SSTVD = TVD – KB. The measurements for directionally drilled wells were dis-
cussed in Chapter 3.
ensure similar competence in other settings. In other words, someone who is an expert at corre-
lating well logs in the South American Andes Mountains Overthrust may not be equally compe-
tent when working in, for example, Offshore West Africa. Just as it took time to become profi-
cient at correlating logs in the South American Andes, so too will it take time and familiarity in
the new area to become proficient.
When geologists correlate one log to another, they are attempting to match the pattern of
curves on one log to the pattern of curves found on the second log. A variety of curves may be
represented on a log. For correlation work, it is best to correlate well logs that have the same type
of curves; however, this is not always possible. A geologist may be required to correlate logs that
have different curves. And at times, even if the logs have the same curves, the character or mag-
nitude of the fluctuations of the curves may be different from one log to the next. Therefore, the
correlation work must be independent of the magnitude of the fluctuations and the variety of
curves on the individual well logs. Figure 4-2 shows sections from two electric logs. The pattern
of curves on Well No. A-1 are very similar to the patterns on Well No. A-2. We can say that these
two logs have a high degree of correlation.
The data presented on a well log are representative of the subsurface formations found in the
wellbore. A correlated log provides information about the subsurface, such as stratigraphic mark-
ers, tops and bases of stratigraphic units, depth and amount of missing or repeated section result-
ing from faults, lithology, depth to and thickness of hydrocarbon-bearing zones, porosity and per-
meability of productive zones, and depth to unconformities. The information obtained from cor-
related logs is the raw data used to prepare subsurface interpretations and maps. The maps may
include fault, structure, stratigraphic, salt, unconformity, and a variety of isochore maps. These
data can also be used to prepare a variety of cross sections. Accurate correlation is the founda-
tion of reliable subsurface interpretations. Subsurface geological maps based on log correlation
are only as reliable as the correlations used in their construction. Eventually, a geologist’s corre-
lations, right or wrong, are incorporated into the construction of subsurface geological maps. An
incorrect correlation can be costly in terms of a dry hole or an unsuccessful workover or recom-
pletion; therefore, it is essential that extreme care be taken in correlating well logs.
In this section, we introduce you to a general correlation procedure and discuss some guide-
lines for electric log correlation. The process of correlating logs varies from one individual to the
next. As geologists gain experience they modify and eventually establish a correlation procedure
that works best for them. If you have no experience in log correlation or want to improve your
skills, you can begin by using the procedures and guidelines discussed in this section.
Electric logs can be correlated by hand or with the assistance of a computer. When correlat-
ed by hand, the electric logs are arranged on a worktable in one of two ways (Fig. 4-3). The
arrangement shown in Fig. 4-3a is preferred over that shown in Fig. 4-3b by most geologists
because more log section can be viewed at one time and the logs are easier to slide during corre-
lation. The manipulation of logs on a computer screen is described in the section Computer-
Based Log Correlation.
As a starting point, align the depth scale of the logs and look for correlation as shown in Fig.
4-2. If no correlation is evident, begin to slide one of the logs until a good correlation point is
found, and mark it. Continue this process over the entire length of each log until all recognized
correlations have been identified. This process may seem relatively easy, but it can be compli-
cated by such factors as stratigraphic thinning, bed dip, faulting, unconformities, lateral facies
changes, poor log quality, and directionally drilled wells. There are some basic, universally valid
guidelines, which are useful in the log correlation process. If followed, these guidelines should
improve your correlation efficiency and minimize correlation problems.
Electric Log Correlation Procedures and Guidelines 63
Figure 4-2 Portion of two electric logs illustrating methods of annotating recognizable correlation pat-
terns on well logs.
1. For initial quick-look correlation, review major sand or carbonate bodies using the SP
curve or gamma ray curve.
2. For detailed correlation work, first correlate shale sections.
3. Initially, use the amplified shallow resistivity curve (focused or short normal), which
usually provides the most reliable shale correlations.
4. Use colored pencils to identify specific correlation points.
5. Always begin correlation at the top of the log, not the middle.
6. Correlate the entire log.
7. Do not force a correlation.
8. In highly faulted areas, first correlate down the log and then correlate up the log.
64 Chap. 4 / Log Correlation Techniques
(a) (b)
Figure 4-3 (a) Preferred method of arranging logs for correlation. (b) Alternate method
of arranging logs for correlation.
After an initial quick look using the SP curve or gamma ray curve to identify the major sand
or carbonate bodies, concentrate your correlation work on shale sections. There are three good
reasons for this. First, the clay and silt particles that make up shales are deposited in low-energy
regimes. These low-energy environments responsible for shale deposition commonly cover large
geographic areas. Therefore, the log curves (sometimes referred to as log signatures) in shales are
often highly correlatable from well to well and can be recognized over long distances. Second,
prominent sand bodies are often not good correlation markers because they commonly exhibit
significant variation in thickness and character from well to well and are often laterally discon-
tinuous. Finally, the resistivity curves for the same sand on two well logs being correlated may
be different. Variations in fluid content in a sand bed may cause pronounced resistivity differ-
ences (e.g., water versus gas).
Individual shale beds exhibit distinctive resistivity characteristics over large areas.
Therefore, when all log curves are considered, the amplified shallow-investigation resistivity
curve provides the most reliable shale correlations. Although all log curves should be used for
correlation work, the amplified resistivity curve is five times more sensitive than the unamplified
curve and exhibits patterns that are easier to recognize and correlate from well to well. The ampli-
fied curve should be the initial curve used for correlation (Fig. 4-2).
The liberal use of colored pencils is an excellent way to identify and mark correlation pat-
terns on well logs. The correlation patterns might be peaks, valleys, or groups of wiggles that are
recognizable in many or all of the well logs being correlated (Fig. 4-2). The colored pencils
should be erasable in the event that correlations are changed. Do not mark on original logs. A
blueline or blackline copy of the original logs should be used for marking during correlation.
In general, structures become less complicated toward the surface because of several factors.
Many faults tend to die upward toward the surface and are either small or nonexistent in the upper
part of the logs. This makes for easier correlations. Also, in many geologic provinces, especially
in soft rock basins, the structural dip, both local and regional, decreases upward. Therefore,
beginning correlation at the top of a log is usually easier.
Correlations are not always straightforward and everyone runs into correlation problems
from time to time. Often there is a tendency to force a correlation rather than bypass the problem
area until further work is done. This is not good practice. Correlation problems are commonly
due to the presence of faults, high bed dips, unconformities, and facies changes. It is best to pass
the problem area and continue the correlation work on the remaining section of the log. Later,
Correlation Type Log 65
when the remainder of the problem log and other logs have been correlated, the questionable cor-
relations can be reviewed again with this new information.
In highly faulted areas it is advantageous to approach a recognized fault from two directions.
First, correlate down the log to the fault and then correlate up the log to the fault. By taking this
approach, determination of the amount of missing or repeated section and depth of the fault in
the correlated well will be more accurate (Figs. 4-2 and 4-10). This method is discussed in detail
later in this chapter.
Figure 4-4 Composite stratigraphic type log. (Published by permission of Texaco USA.)
Correlation Type Log 67
TYPE LOG
( COMPOSITE )
GOOD HOPE FIELD
Figure 4-5 Composite show log from Good Hope Field, St. Charles Parish, Louisiana. (Published by
permission of the New Orleans Geological Society.)
68 Chap. 4 / Log Correlation Techniques
Figure 4-6 A cross section through a complex diapiric salt structure, penetrated by four vertical wells.
For this particular example, the correlation type log must be a composite log including sec-
tions from Wells No. 2 and 3. The best type log consists of Well No. 2 from the surface down to
a correlation marker just below the 11,000-ft Sand and Well No. 3 from the same marker to TD
in diapiric shale. This composite type log meets all the requirements in the definition shown ear-
lier and is an excellent standard for all other well log correlation work on this
structure.
Electric Log Correlation – Vertical Wells 69
Normally, faults are not included on a correlation type log. However, if a major decollement,
such as a thrust or listric growth fault, serves as the deepest limit of prospective section, it is
advisable to place the fault on the type log.
Step 1. First, prepare a correlation type log. Remember that a correlation type log must show
a complete unfaulted interval of sediments representative of the thickest and deepest sedi-
mentary section in the field. For the structure in Fig. 4-7, the wells furthest off structure, such
as Wells No. 5 or 7 or a composite of the two, are good candidates for a type log. These wells,
positioned off the crest of the structure, should show the thickest and most complete sedi-
mentary section.
Step 2. A good correlation plan involves the correlation of each well with a minimum of two
other wells. To ensure good correlation efficiency over the entire area, the plan should be
established to correlate by means of closed loops. The correlation plan in Fig. 4-7 illustrates
a sequence of closed loops. The recommended order of correlation is represented by “billiard
ball” type numbers for correlation sequence. Using this procedure, the log correlation work
within a loop begins and ends with the same log, eliminating correlation mis-ties and reduc-
ing the chance of other correlation errors.
Step 3. First correlate wells expected to exhibit the most complete and thickest stratigraph-
ic section. On a structure such as the one shown in Fig. 4-7, the structurally lowest wells usu-
ally have the thickest section. These include wells represented by the “billiard ball” type cor-
relation sequence number 1.
Step 4. Continue log correlation, progressing from wells in a down-structure position to
wells in an up-structure position (see the billiard ball type correlation sequence numbers 2,
3, 4, and 5). These numbers indicate the recommended correlation sequence for the
example.
70 Chap. 4 / Log Correlation Techniques
Figure 4-7 An example of a log correlation plan for vertical wells. Notice that there
is a hierarchy in the correlation sequence and that all the wells are correlated in
closed loops.
Electric Log Correlation – Vertical Wells 71
Step 5. Generally, correlate wells located nearest each other. In most cases, closely spaced
wells should have a similar stratigraphic section and so correlation is usually easier.
Step 6. In many geologic areas, rapid changes in stratigraphy (particularly changes in thick-
ness) occur over short distances. Where possible, correlate wells anticipated to have a similar
stratigraphic interval thickness. In extensional or diapiric tectonic areas, wells at or near the
same structural position often exhibit similar stratigraphy. For areas involving plunging folds,
similar stratigraphy may be exhibited along the plunge of the fold.
Also, if wells that exist at or near the same structural bend are connected together, as in the
first order correlation as shown in Fig. 4-7, the dip rate of the structure should be similar in these
various wells, thus eliminating correlation problems related to changing structural dip.
In geologic settings containing syndepositional faults (commonly referred to as growth
faults), some special considerations must be given in preparing a log correlation plan. For our
purposes, we define a growth normal fault as a syndepositional fault resulting in an expanded
stratigraphic section in the hanging wall fault block, with missing section on the growth fault
changing with depth.
If a growth fault is present in the area of study, restrict the correlation to wells within one
fault block of the growth fault. Keeping in mind that the hanging wall block of a major growth
fault has an expanded stratigraphic section, which can increase the difficulty in correlation, start
the correlation in the footwall block using the plan just outlined. Once the correlations are com-
pleted in the footwall fault block of the major growth fault, carry the correlations, if possible, into
the hanging wall fault block. Initially, check the correlation in wells located in down-structure
positions.
If a significant amount of growth has occurred on the fault, the thickness of the sediments
can be so great in the hanging wall block that correlation of the section from hanging wall to
footwall blocks may be difficult, if not impossible. In such a case, the best correlations can be
achieved by preparing a separate type log for the hanging wall block and correlating this fault
block independently from the footwall block.
We note here that other terminology is at times used in place of hanging wall and footwall.
For a normal fault, the hanging wall is also referred to as the downthrown fault block. As well,
the footwall block can be referred to as the upthrown fault block.
(a)
(b)
Figure 4-8 (a) Correlation of two vertical wells using the major sands as the primary vehicle
for correlation. (b) Detailed correlation of the two vertical wells shown in Fig. 4-8a using all the
reliable shale and sand correlation markers. (SRM = shale resistivity marker.)
Electric Log Correlation – Vertical Wells 73
relating Wells No. A-1 and 3 in Fig. 4-8a. By correlating the sands, we see that the interval from
the top of the 10,000-ft Sand to the top of the 10,300-ft Sand is about 325 ft thick in Well No. A-
1 and 480 ft in Well No. 3. The interval in Well No. A-1 is short between the two sand tops by
155 ft. This short interval, based on the sand correlations, suggests the possibility of a 155-ft nor-
mal fault in Well No. A-1.
Now we correlate these same two logs, shown in Fig. 4-8b, using the guidelines outlined ear-
lier. The guidelines recommend that detailed correlations be conducted in the shale sections using
all the electric log curves with an initial emphasis on the amplified resistivity curve. This curve
provides the most reliable shale correlations.
Through detailed correlations of the shale sections and the sands, a number of correlation
markers are identified on the two logs. These include a series of shale resistivity markers (SRMs)
labeled SRM No. 1 through SRM No. 4, with certain diagnostic resistivity correlation patterns
highlighted on the resistivity side of the logs, in addition to the two major sands. All these corre-
lation markers indicate that both log segments have a high degree of correlation and that no fault
is present in Well No. A-1.
It appears that the stratigraphic section in Well No. A-1 is uniformly thin relative to Well No.
3. The thickness ratio for the intervals between each of the four shale markers shows a consis-
tency in the stratigraphic thinning in Well No. A-1 when compared to Well No. 3. This uniform
thinning supports the idea that although Well No. A-1 is short to No. 3 as a result of stratigraph-
ic thinning, the two logs exhibit correlation.
Fault Determinations: Depth and Missing Section. Now that we have a basic under-
standing of how shale markers are used to aid in log correlation, look at the log segment from
two other electric logs run in vertical wells (Wells No. 1 and 3 in Fig. 4-9). Some geologists pre-
fer to correlate major sands. If we do that and use the 8600-ft and 9000-ft Sands as the principal
correlations, we see that the section in Well No. 3 between the two major sands is 80 ft short and
a fault appears possible in the well. With the limited correlation data, the missing section and
depth of the fault is uncertain. Also, the correlation of the top of the 9000-ft Sand in Well No. 3
is questionable. Is there a fault in Well No. 3 and is the fault (1) within the shale interval between
the base of the 8600-ft Sand and the top of the 9000-ft Sand, (2) at the top of the 9000-ft Sand,
or (3) is part of the 9000-ft Sand faulted out? If the fault is above or at the top of the 9000-ft Sand,
then the interval from the 8600-ft Sand to the top of the 9000-ft Sand is 80 ft short. If part of the
top of the 9000-ft Sand is faulted out, then the interval is short by some amount greater than 80
ft. With the major sand correlation methodology, the nature of the short section in Well No. 3 is
not apparent and so we have a correlation problem.
74 Chap. 4 / Log Correlation Techniques
Figure 4-9 Correlation of the major sands in two vertical wells may provide insufficient
correlation data to accurately determine the depth and size of a suspected fault.
Now we will follow the recommended correlation procedures illustrated in Fig. 4-10. These
procedures provide a number of correlation markers, including shale resistivity markers 1
through 7 and specific resistivity characteristics highlighted on the resistivity curve of each log.
These detailed correlation markers show that the interval between each correlation marker is
comparable in both wells except for the short section identified between SRM 5 and SRM 7.
Notice that SRM 6 is missing in Well No. 3, as is the lower segment of the resistivity character
highlighted through SRM 6 in Well No. 1. Finally, these detailed shale correlations and the cor-
relation data contained within the 9000-ft Sand indicate that the upper portion of the 9000-ft Sand
is also missing in Well No. 3. We have isolated the short section in Well No. 3 to a specific inter-
val 135 ft thick. The isolation of this short section to one particular location indicates that
the short section is the result of a fault rather than a variation in stratigraphy. The location
of the short section provides the depth of the fault in this well. By measuring the amount of sec-
tion in Well No. 1 that is missing in Well No. 3, we determine the missing section due to the fault
(135 ft) by correlation with Well No. 1. The missing section is highlighted in Fig. 4-10. In order
to ensure confidence in the fault, Well No. 3 should be correlated with at least one more nearby
well.
For each fault recognized in a well there are three important pieces of data that must be
obtained for documentation and later use in mapping: (1) the amount of missing section, (2) the
Electric Log Correlation – Vertical Wells 75
Figure 4-10 Detailed correlation of the two vertical wells shown in Fig. 4-9 using all recogniz-
able correlation markers to determine the depth and missing section for a fault in Well No. 3.
Notice that the top of the 9000-ft Sand and SRM 6 are faulted out of Well No. 3.
depth of the fault in the log, and (3) the well or wells correlated to identify the fault. The fault
data (135 ft/8957 ft/Well No. 1) and information regarding the faulted out (F/O) top of the 9000-
ft Sand are annotated next to the fault symbol on the log. Refer to Fig. 4-10 again for an exam-
ple of how these data are annotated on a log. For convenience, in most of our examples we use
measured well depths for faults and other points. For your mapping purposes, we recommend that
you annotate subsea depths.
The accuracy of identifying the depth of a fault in a well and determining the amount of
missing section is directly related to (1) the detail to which the logs are correlated, (2) the num-
ber of logs used for correlation, and (3) variations in stratigraphic thickness seen in the wells.
Obviously, the smaller the interval between established correlation markers, the more precise the
correlation in pinpointing the depth of the fault in the well and the amount of missing section.
Missing section can be incorrectly estimated if the reference well is of different thickness than
the faulted well. A thickness ratio can be used to appropriately adjust the amount of missing sec-
tion that is measured in the reference well log (see following section, Stratigraphic Variations).
The correlation detail and accuracy required are often dictated by the type of geologic study
being conducted. For example, if you are involved in a regional geologic study, pinpointing the
depth of a fault within several hundred feet on a well log may be sufficient. Also, you may only
76 Chap. 4 / Log Correlation Techniques
be interested in the larger faults (i.e., faults greater than 100 ft). If the study is to be detailed for
field development or enhanced recovery, however, it may be necessary to locate the depth of all
recognizable faults to within ± 20 ft. The same variation in accuracy applies to missing section
as well.
Stratigraphic Variations. Figure 4-11 shows a log segment from Wells No. A-1 and 3. In this
section, we use the correlation procedures to establish specific correlation markers to recognize
stratigraphic variations so that such thickness changes are not mistaken for faults.
In Fig. 4-11, two correlation markers are identified in each well. Based on these markers,
Well No. A-1 is 155 ft short to Well No. 3. Is the short section in Well No. A-1 the result of a fault
or variations in stratigraphy? With the limited correlation data shown in the figure, it is impossi-
ble to determine why the section in Well No. A-1 is short. We could use the major sands in each
well to aid the correlation work, but this added information provides little help in determining the
nature of the short section.
So far, we have shown that it is important to identify as many correlation markers as possi-
ble, especially in questionable log intervals. Closely spaced correlations generally improve the
accuracy of the correlation, help differentiate faults from stratigraphic variations, and improve
the estimate for the amount of missing section and depth of identified faults. Therefore, in order
to accurately correlate Wells No. A-1 and 3, additional correlation markers are required.
Figure 4-11 Correlation of Wells No. A-1 and 3 using limited correlation markers. Is
there a fault in Well No. A-1?
Electric Log Correlation – Vertical Wells 77
Figure 4-12 shows the same two logs with additional correlation markers identified. The cor-
relation process is improved with these additional markers. Notice that the shortening in Well No.
A-1 is not isolated to one specific interval, but is present in all the intervals in the well between
SRM 1 and SRM 4. This evidence strongly suggests that the thickness variations in Well No.
A-1 are stratigraphic and not the result of faulting. If necessary, interval thickness ratios can be
calculated between correlation markers to provide further evidence to support the conclusion.
Ts
Tr = (4-1)
Tl
Where
Tr = Thickness ratio
Ts = Interval thickness in short log
Tl = Interval thickness in long log
Figure 4-12 Correlation of Wells No. A-1 and 3 using all recognizable correlation
markers indicates that there is no fault in Well No. A-1. The short log section is the
result of variations in stratigraphy.
78 Chap. 4 / Log Correlation Techniques
(a)
(b)
Figure 4-13 (a) Electric log stratigraphic section laid out perpendicular to the strike of the 8500-ft Sand (par-
allel to dip). (b) Electric log section showing the relationship of TST to TVT with changing bed dip. The TVT
is that thickness seen in a vertical or straight hole.
Electric Log Correlation – Vertical Wells 79
where
Logs cannot be correlated in a vacuum. The correlation plan shown earlier illustrated the
need to know the structural relationship of logs being correlated. This can be accomplished by
having a well-log base map available during correlation that shows the general structure and the
location and structural position of each well being correlated.
Finally, let’s consider a situation as shown in Fig. 4-14. In this case, the stratigraphic section
identified in the two wells has a decreasing TST in the up-structure direction. We can say that the
structure was actively growing during the time of deposition of the section, resulting in strati-
graphic thinning toward the crest of the structure. However, by log correlation, the vertical log
thickness of the stratigraphic section in each well is exactly the same. If you recognized the same
interval thickness in each well irrespective of structural position, you might make the incorrect
80 Chap. 4 / Log Correlation Techniques
Figure 4-14 The cross section shows the effect of changing bed dip on the true
vertical thickness of a stratigraphic section.
assumption that since the thicknesses are equal, the structure was not active during deposition, or
that the TST of the stratigraphic section as seen in both wells is the same. A review of the wells
in cross section in Fig. 4-14 shows that the stratigraphic thickness of the section actually decreas-
es in the up-dip direction such that the stratigraphic thickness in Well No. 1 is only one-half the
thickness found in Well No. 4. It is the changing bed dip that causes the vertical log thickness to
be the same in each well. Equation (4-2) can be used to calculate the TST in each well to devel-
op a better understanding of the stratigraphic thicknesses as seen in each well.
These examples show that log thickness in vertical wells varies with changes in structural
dip and is equal to the TST only when the dip of the unit is zero (Fig. 4-14, Well No. 4). They
also emphasize that log correlation is not an isolated process. Structural and stratigraphic rela-
tionships and geologic knowledge of the area of study must always be kept in mind during cor-
relation. Errors in correlation and incorrect assumptions on such aspects as the local growth his-
tory of a structure may be incorporated into the geologic work. Such pitfalls can be prevented if
the structural and stratigraphic framework of the area is considered during correlation.
Electric Log Correlation – Directionally Drilled Wells 81
Step 1. Construct a correlation type log. Refer to the section on correlation type logs for the
complete definition of a type log. Do not use a deviated well in the construction of a type log
because a log from a directionally drilled well does not represent the true vertical strati-
graphic section. Wells farthest off structure serve as good type log candidates.
Step 2. Correlate all the vertical wells before correlating the deviated wells, since the verti-
cal wells are usually easier to correlate. For the vertical wells, use the same plan outlined in
Fig. 4-7.
Step 3. Once the vertical wells have been correlated, begin correlating the deviated wells. To
begin directional well log correlation, first organize the wells according to their direction of
deviation with respect to structural strike. Deviated wells are classified into one of three
groups: (1) wells drilled down-dip, (2) wells drilled along strike, and (3) wells drilled up-dip.
Step 4. Begin correlation of these three groups with the wells drilled generally down-dip.
First correlate the wells with the least amount of deviation, and where possible, correlate in
closed loops with each well log correlated with a minimum of two other wells. The wells with
the least amount of deviation will have a log section thickness closer to that seen in a verti-
cal well than other wells drilled down-dip. Looking at the wells drilled from Platform B in
Fig. 4-16, the first directional wells correlated are those represented by a billiard ball type cor-
82 Chap. 4 / Log Correlation Techniques
SURFACE SURFACE
0' 0'
1000' KOP = 1000' MD, 1000' TVD 1000' KOP = 1000' MD, 1000' TVD
ANGLE MADE AT 2432' TVD, 2500' MD ANGLE MADE AT 2432' TVD, 2500' MD
4000' 4000'
DROP RATE 1-1/2O/100'
HORIZONTAL WELL
SURFACE
0'
4000' (c)
Figure 4-15 Diagrammatic cross sections illustrating (a) a simple ramp or “L” shape well; (b) a more
complicated “S” shape well; (c) a horizontal well. [(a) and (b) published by permission of Tenneco Oil
Company, (c) published by permission of J. Brewton.]
Electric Log Correlation – Directionally Drilled Wells 83
relation sequence number 1. There are two wells drilled with a minimum down-dip deviation
(Wells No. B-5 and B-6). These wells can be correlated to each other and then with the
straight hole, Well No. B-1, drilled as a vertical well from the platform.
Step 5. Continue correlating wells with increased deviation in the down-dip direction. For
this example, these are Wells No. B-2 and B-3, indicated by correlation sequence number 2.
These two highly deviated wells can be correlated with each other and then with Wells No.
B-5 and B-6. Also, the vertical Well No. 3 may be used to correlate B-2 and B-3, since it is
an off-structure well exhibiting a thick stratigraphic section.
Step 6. When all wells classified as being deviated down-dip are correlated, the next group
to correlate are those wells deviated along structural strike. From Platform B, Wells No. B-7
and B-9 fall into this category. These wells can be correlated to each other and then with
straight hole B-1 to close the loop. When correlating wells drilled along strike, the effect of
bed dip is removed from the representative thickness of the directionally drilled wells. This
can often simplify correlation.
Step 7. Finally, correlate the wells deviated up-dip. Those wells drilled closest to the crest
of the structure usually are complicated by stratigraphic thinning, faulting, and unconformi-
ties. The correlation of these wells can be most difficult; therefore, they are normally corre-
lated last when all other correlation information is available and you can recognize the best
correlation markers. Wells No. B-4 and B-8 drilled from the B Platform fall into this catego-
ry. They are labeled as correlation sequence number 4. Wells drilled in an up-dip direction
can have variable log section thickness due to the geometric relationship between a wellbore
and structural dip. A log section from a well drilled in an up-dip direction can be thicker, thin-
ner, or equal to the thickness of a log section from a nearby vertical well drilled through the
same stratigraphic section. This potential complexity can add to the difficulty of correlating
wells drilled in an up-dip direction. Because of these complexities, we recommend that these
wells be correlated last, after significant knowledge is gained from other correlation work.
Step 8. Generally, it is best to correlate wells located nearest each other, especially in areas
where significant changes in stratigraphic thickness are probable. Wells nearest each other
and approximately in the same structural position usually are expected to have the most com-
parable interval thicknesses.
Step 9. After correlating the wells from one platform, begin correlation of wells on any addi-
tional platforms in the area. In Fig. 4-16, the A Platform wells in the northwest portion of the
field should be correlated next. It is not necessary, however, to isolate correlation to a single
platform. Often, wells from one platform are drilled in a direction toward another platform.
If wells from separate platforms are in close proximity to one another, they should be corre-
lated to each other. Notice that correlation sequence number 5 illustrates the correlation of
B-4 with A-5, and B-8 with A-4. Wells No. A-1 and B-1 are straight holes drilled from sep-
arate platforms, but since they are located in a similar structural position, they can also be
correlated to each other.
The primary focus of this correlation plan is to provide a logical method for correlating all
vertical and deviated wells in an area of study. The plan outlined is by no means the only one that
can be used. The important point is to have a plan. Without one, log correlation becomes a ran-
dom process, often resulting in some type of correlation problem or in miscorrelations.
84 Chap. 4 / Log Correlation Techniques
Figure 4-16 An example of a log correlation plan for directionally drilled wells. The
plan shows the hierarchy of the log correlation sequence and illustrates how to cor-
relate deviated wells in closed loops.
(sand and shale sections) for both wells indicates that they have penetrated the same stratigraph-
ic section. Although both wells have a high degree of correlation (see SRM 1 through SRM 4),
the stratigraphic section in Well No. A-2 is much thicker than the same section seen in Well No.
A-1. The log section in Well No. A-1 from SRM 1 to SRM 4 is 490 ft thick. The same section in
Well No. A-2 is 735 ft. Earlier in the chapter, in the discussion on vertical wells, we mentioned
that a short section in one well with respect to another might be the result of stratigraphic changes
or a fault. If the short section is isolated to one particular location, the short section is most like-
ly the result of a fault rather than variations in stratigraphy. Conversely, if the short section is uni-
formly distributed over a series of intervals, the short section is probably due to stratigraphic vari-
ations rather than a fault.
Based on correlation criteria, the thinner section in Well No. A-1 appears to be the result of
stratigraphic thinning rather than a fault. In this example, however, we introduce another possi-
ble explanation for the shortening. Since Well No. A-2 is directionally drilled, the thickness seen
in the well with respect to Well No. A-1 may be completely the result of the wellbore deviation.
Figure 4-18 shows vertical Well No. A-1 and deviated Well No. A-2 in its true orientation with
respect to the vertical. Well No. A-2 is drilled due west at a deviation angle of 48 deg (48 deg
from the vertical). The correlation markers in each well show that the strata are horizontal and
the thick section seen in Well No. A-2 is solely the result of wellbore deviation. We have now
Figure 4-17 Portion of an electric log from a vertical well (A-1) and a directionally
drilled well (A-2). The electric log sections show detailed correlations.
86 Chap. 4 / Log Correlation Techniques
introduced another complexity in correlation that must be considered when both vertical and
deviated wells are present in the area of study.
Here are several procedures that can be used to help correlate a vertical well with a direc-
tionally drilled well.
1. Mark the angle of deviation for the directional well on the log at least every 1000 ft. This
provides a reminder that the well is deviated and indicates the angle of deviation at 1000-
ft intervals on the actual log.
2. To compare interval thicknesses, slide the vertical well log as you correlate from marker
to marker. This allows you to compensate during correlation for the expanded or reduced
section in the directional well as a result of its deviation.
3. Calculate a thickness ratio for certain correlation intervals of interest to help evaluate
whether any short section is the result of faulting, stratigraphic thinning, or just wellbore
deviation (Fig. 4-18).
4. If a copy machine with a reduction mode is available, calculate the correction factor
required to convert the deviated (stretched) log section to a vertical log section, and then
reduce the log by the appropriate reduction factor. Use the reduced log for correlation.
5. In areas of horizontal beds or low relief, the MD log from a deviated well can be
corrected for wellbore deviation and converted into a TVD (true vertical depth) log
to use for correlation.
6. In areas with bed dips greater than 5 to 10 deg, if dip data are available from a dipmeter
log or previously constructed structure maps, these data can be used to convert the
deviated log to a TVT (true vertical thickness) log. A TVT log is one in which the
measured thickness has been corrected for wellbore deviation and bed dip to the thickness
represented in a vertical well. In areas of dip, a TVD log provides little aid, if any, in
correlation and can actually cause correlation problems (see section on MLT, TVDT, TVT,
and TST).
Horizontal Beds. We begin with a fault in the deviated Well No. A-2 (Fig. 4-l9a). By correla-
tion with Well No. A-1, this well cuts a fault near the 10,000-ft Sand level. To determine the depth
and missing section, we correlate the logs in the same manner as previously outlined in this chap-
ter. First, correlate down the logs starting with SRM 1. We can say that correlation is lost at points
A in both wells. Mark this location on the two logs. Next, find a correlation point below this sec-
tion on the logs, such as SRM 4, and correlate up the logs. We now lose correlation in the wells
at points B. By detailed correlation of the shale markers and sands, we have determined that Well
No. A-2 is faulted, and the section in Well No. A-1 that is stratigraphically equivalent to the miss-
ing section in Well No. A-2 is highlighted in Fig. 4-19a.
The faulted out or missing section in Well No. A-2 is equal to 150 ft by correlation with Well
No. A-1. Notice that the base of the 10,000-ft Sand is faulted out of Well No. A-2. This informa-
tion is annotated on the log along with the amount of missing section, the depth of the fault, and
the well(s) used to correlate the fault.
Electric Log Correlation – Directionally Drilled Wells
Figure 4-18 Vertical Well No. A-1 and deviated Well No. A-2 (shown in its true orientation with respect to verti-
cal). The correlation markers show that the thicker section in Well No. A-2 is a direct result of its deviation from the
vertical.
87
88
Chap. 4 / Log Correlation Techniques
(a)
Figure 4-19a Detailed correlation of a deviated well with a vertical well to locate the depth and the missing sec-
tion for a fault in the deviated well. The base of the 10,000-ft Sand is faulted out.
Electric Log Correlation – Directionally Drilled Wells
Figure 4-19b The simplified stratigraphic section through Wells No. A-1 and A-2 illustrates that the missing sec-
tion in Well No. A-2 is equivalent to the vertical section highlighted in Well No. A-1. No thickness correction factor
is required in this example.
89
90 Chap. 4 / Log Correlation Techniques
The missing section in directional Well No. A-2 is determined by correlation with Well No.
A-1, which is a vertical well. In a vertical well, the log thickness and vertical thickness are the
same. Since missing section is expressed as the vertical thickness of the stratigraphic interval
faulted out of a well, the vertical thickness of the missing section in Well No. A-2 is 150 ft. The
150 ft represents the missing section for the fault. This information will be used in future fault
and structure mapping.
Figure 4-19b is a simplified stratigraphic section showing Wells No. A-1 and A-2 positioned
in their true orientation with respect to the vertical. Well No. A-2, which is deviated at 48 deg
from the vertical, is pulled apart at the fault to show the restoration of the faulted-out section. This
cross section clearly illustrates that the missing section in Well No. A-2 is equal to the 150 ft of
vertical section highlighted in Well No. A-1.
Now consider a fault in vertical Well No. A-1 correlated with deviated Well No. A-2 (Fig.
4-20a). Well No. A-1 has a fault near the base of the 10,000-ft Sand. Detailed correlation, as
shown in the figure, identifies a 225-ft section in deviated Well No. A-2 that is faulted out of Well
No. A-1. The faulted-out section is highlighted in the figure. Since the missing section for the
fault is determined as the TVT of the stratigraphic interval faulted out of the well, the estimate of
225 ft of missing section based on the deviated log thickness must be corrected to express the
missing section in terms of TVT.
Figure 4-20b is a stratigraphic section showing Wells No. A-1 and A-2 positioned in their
true orientation relative to vertical. The log section of Well No. A-1 is pulled apart at the fault to
show the restoration of the faulted out section. Since we are working in an area with horizontal
beds, the correction of the measured log thickness in Well No. A-2 to TVT is determined by the
simple trigonometric solution of a right triangle. The insert in the center of the figure shows that
the TVT of the missing section is equivalent to the vertical side of a right triangle whose
hypotenuse is equal to the log thickness of the missing section in deviated Well No. A-2.
Where
Therefore,
The actual (corrected) missing section for the fault in Well A-1 determined by correlation
with deviated Well No. A-2 is 150 ft.
Electric Log Correlation – Directionally Drilled Wells 91
(a)
Figure 4-20a Detailed correlation of a vertical well with a deviated well to locate the depth and
determine the missing section for a fault in the vertical well.
Dipping Beds. The procedure for correlating deviated wells in an area of significant dip is, in
effect, the same as that presented thus far in this chapter. The primary difference occurs in esti-
mating the actual missing section resulting from a normal fault. Since the missing section due to
a fault is defined as the TVT of the stratigraphic interval faulted out of a well, any missing sec-
tion value determined by correlation with a deviated well exhibiting a measured log thickness
must be converted to TVT. In the last section, we defined a simple trigonometric relationship for
calculating the correction factor applicable in areas with horizontal beds. Where dipping beds are
present, the mathematical correction factor becomes somewhat more complex.
There are several equations available for calculating a correction factor to convert MLT in a
deviated well to TVT. We present two separate methods for computing the correction factor.
92
Chap. 4 / Log Correlation Techniques
(b)
Figure 4-20b A simplified stratigraphic section illustrating the relationship of the missing section in vertical
Well No. A-1 to the exaggerated sections seen in deviated Well No. A-2. The exaggerated section in Well No.
A-2 must be corrected for wellbore deviation to determine the true amount of missing section.
Electric Log Correlation – Directionally Drilled Wells 93
(a)
(b)
Type 1 Equation. The derivation of the Type 1 equation (well deviated up-dip) is shown here
and illustrated in Fig. 4-21b.
cosρ = MC / MT (1a) cos φ a = VE / VT (2a)
MC = MT cos ρ (1b) VE = VT cos φ a (2b)
MC = VE
94 Chap. 4 / Log Correlation Techniques
MT cos ρ = VT cos φ a
Rearranging,
MT cos ρ
VT =
cos φ a
cos ( ψ 1 + φ a ) (4-4)
TVT = MLT
cos φ a
Type 2 Equation. The derivation of Type 2 (well deviated down-dip) is shown here and illus-
trated in Fig. 4-21c.
MC = ST
Rearranging,
cos γ
TVT = MLT
cos φ a
Substituting γ = ψ 2 + φ a
cos ( ψ 2 + φ a )
TVT = MLT (4-5)
cos φ a
With Eqs. (4-4) or (4-5), the data required to calculate the correction factor are (1) ψ = well-
bore deviation from the vertical, (2) φa = apparent bed dip (bed dip in the direction of wellbore
deviation), and (3) MLT = measured log thickness in the deviated well. The apparent bed dip is
the most difficult data to obtain for these equations. The only source of apparent bed dip is from
an existing structure map.
Setchell (1958) and has been used successfully for over 40 years. We consider this three-dimen-
sional correction factor equation preferable because this equation can be used to calculate the cor-
rection factor regardless of the direction of wellbore deviation, and the true dip of the beds is used
instead of the apparent dip, which is used in the two-dimensional equations.
To derive the general three-dimensional equation, we introduce a three-dimensional spheri-
cal coordinate system (Fig. 4-22) with one horizontal axis in the direction of dip (called the x-
axis), one axis perpendicular to the first and also horizontal (called the y-axis), and finally, one
axis perpendicular to the other two (called the z-axis). The origin is the point (T) at which the
wellbore first penetrates the bed.
Where
φ ψ
φ
Figure 4-22 A three-dimensional spherical coordinate system with one horizontal axis in the direction
of bed dip (x-axis), one axis perpendicular to the first and also horizontal (y-axis), and finally, one axis
perpendicular to the other two (z-axis). (Published by permission of D. Tearpock).
96 Chap. 4 / Log Correlation Techniques
Substituting the spherical coordinate definition for Z, X, and the alternate trigonometric def-
inition of Z easily yields Eq. (4-6).
Z = MLT cos ψ
X tan φ + TVT = MLT cos ψ
TVT = MLT cos ψ – X tan φ
TVT = MLT cos ψ – MLT sin ψ cos α tan φ
or
Figure 4-23 One vertical and two deviated wells penetrate Bed A. In Well No. A-1, Bed A is faulted out. The cross sec-
tion shows the relationship of the missing section in Well No. A-1 to the MLT of the equivalent interval in deviated Wells
No. A-2 and A-3. The equation shown is used to correct MLT to TVT.
Electric Log Correlation – Directionally Drilled Wells 97
Application of Methods 1 and 2. Now apply these methods using the data shown in Fig.
4-23. Wells No. A-1, A-2, and A-3 are drilled from a single location and penetrate a section that
includes Bed A, which is of constant thickness and dips at 35 deg due east. Well No. A-1, a ver-
tical well, cuts a fault that completely faults out Bed A. Well No. A-2 is drilled in a down-dip
direction with an average deviation angle of 36 deg through Bed A. Well No. A-3 is drilled direct-
ly up-dip with a deviation angle of 35 deg through Bed A. For simplicity, the wells are all
assumed to be drilled in a vertical plane parallel to the dip of Bed A.
Through detailed correlation of the three wells, a fault is identified in Well No. A-1. Based
on the correlations, Bed A is completely faulted out of the well. Assuming that Wells No. A-2 and
A-3 are the only wells available for correlation, the amount of missing section for the fault in Well
No. A-1 must be estimated from these deviated wells.
Considering again that missing section is defined as the true vertical thickness of the strati-
graphic interval faulted out of a well, and the fault in Well No. A-1 completely faults out Bed A,
the missing section resulting from the fault in Well No. A-1 is equal to the true vertical thickness
of Bed A. A review of Fig. 4-23 shows that Bed A has a true vertical thickness of 200 ft. When
we correlate Well No. A-1 with the two deviated wells, however, we obtain a missing section in
terms of deviated log thickness rather than true vertical thickness. Therefore, these deviated log
thicknesses must be converted to true vertical thickness to estimate the amount of missing sec-
tion. The missing section (Bed A) in Well No. A-1 has a logged thickness in Well No. A-2 of 504
ft and a logged thickness in Well No. A-3 of 164 ft. The logged thickness of 504 ft in Well No.
A-2 is over two and one-half times greater than the TVT. The logged thickness of 164 ft in Well
No. A-3 is less than the TVT (0.82).
By log correlation, the amount of missing section for the fault in Well No. A-1 ranges from
164 ft, based on correlation with Well No. A-3, to 504 ft, by correlation with Well No. A-2. What
is the actual missing section for the fault? In order to determine the actual missing section, the
log thickness measured in Wells No. A-2 and A-3 must be corrected to true vertical thickness.
Data:
ψ1 = 35 deg
φa = 35 deg
MLT = 164 ft
cos ( ψ 1 + φ a )
TVT = MLT
cos φ a
cos 0°
= 164'
cos 35°
1
= 164'
0.8192
98 Chap. 4 / Log Correlation Techniques
= 164' (1.2207)
TVT = 200 ft
Type 2: Well Drilled Down-dip. In order to use Eq. (4-5), the same parameters used in Eq. (4-4)
are required.
Data:
ψ1 = 36 deg
φa = 35 deg
MLT = 504 ft
cos ( ψ 1 + φ a )
TVT = MLT
cos φ a
cos (36° + 35° )
= 504'
cos 35°
cos 71°
= 504'
cos 35°
0.3256
= 504'
0.8192
= 504' (0.3975)
TVT = 200 ft
Since Wells No. A-2 and A-3 are drilled directly down-dip and up-dip respectively, the value
for bed dip in Eqs. (4-4) and (4-5) is equal to the true bed dip (φ). This can be obtained from a
structure map or a dipmeter if available. If the direction of wellbore deviation is not parallel to
bed dip, however, then an apparent bed dip in the direction of wellbore deviation must be deter-
mined. This apparent bed dip cannot come from a dipmeter log since this log calculates true bed
dip. This is one of the main drawbacks to the two-dimensional equations.
ψ = 36 deg
αw = 90 deg
Electric Log Correlation – Directionally Drilled Wells 99
φ = 35 deg
αa = 90 deg
MLT = 504 ft
α = 0 deg
TVT = MLT [cos ψ – (sin ψ cos α tan φ)]
[
= 504' cos 36° – (sin 36° cos 0° tan 35° ) ]
= 504' [0.809 – (0.5878) (1) (0.70)]
= 504' [0.809 – 0.412]
TVT = 504' [0.397]
TVT = 200 ft
TVT for Well No. A-3: The data for this calculation are exactly the same as for Well No. A-2, with
two exceptions. The azimuth (αw) for Well No. A-3 is due west or 270 deg, and therefore the ∆
azimuth (α) is 180 deg and ψ is 35 deg.
[
= 164' cos 35° – (sin 35° cos 180° tan 35° ) ]
= 164' [0.8192 – (0.5736) (–1) (0.7002)]
= 164' [0.8192 – (–0.4016)]
= 164' [0.8192 + 0.4016]
TVT = 164' [1.2208]
TVT = 200 ft
Through the use of the two-dimensional and three-dimensional equations, we have success-
fully calculated the amount of missing section for the fault in Well No. A-1. Notice the close
agreement between the different equations. The missing section estimated at 200 ft can now be
used for all future fault surface mapping and structure map integration as well as other geologi-
cal work, such as cross-fault drainage analysis. The actual procedure for integrating a fault and
structure map is detailed in Chapter 8. We have now defined two methods for obtaining the cor-
rect values for missing section when correlating with deviated wells. Anytime you are correlat-
ing in an area with deviated wells and dipping beds, missing and repeated section are defined in
terms of TVT. Significant errors in structure maps, net pay maps, and even proposed well loca-
tions can occur if these corrections are not made or if missing section is not understood in terms
of vertical separation (see Chapter 7).
To
p Deviated Well
of
St
ra TS
ta 12 T
3' To
p of
Ba Deviated Well St
se TVT ra
ta TVD Thickness
of 150'
Me
St 357'
ra
a su
ta 7'
re
TVDT
12 35o
d
82' Ba
Lo
=
LT TVT = 150' se
gT
TVT
M MLT = 466' of
hic
150' St
35o TVDT = 357'
kn
ra
t
es
TST = 123' a
s=
TS
47
12 T TVDT = MLT COS Wellbore Angle
100'
6'
3'
TST = TVT COS Bed Dip
50o
100'
MLT = 127'
0 100' TVDT = 82'
TVT = 150'
0 100' 40o
TST = 123'
(a) (b)
VERTICAL
WELL
Deviated Well
Top of Strata
M
TVDT = LT
TVT=150'=TST =
150' 21
2'
100' Figure 4-24 (a) The MLT in a well deviated up-dip is
Base of Strata compared to TVDT, TVT, and TST. (b) The MLT in a
well deviated down-dip is compared to TVDT, TVT,
and TST. (c) The TVDT, TVT, and TST calculated from
0 100'
(c) a deviated well have the same value when the beds
are horizontal.
is normally not equal to TVT due to the wellbore deviation and to bed dip in areas of dipping
beds. However, remember that MLT will approximate TVT in the near-vertical part(s) of a direc-
tional well.
True vertical depth thickness (TVDT) is defined as the MLT between two specific points in
a deviated well, corrected only for wellbore deviation. The true vertical thickness (TVT) is
defined as the thickness of an interval measured in the vertical direction. It is the thickness seen
in a vertical well. For a directionally drilled well, the TVT can be calculated using the equations
introduced in the previous section. The true stratigraphic thickness (TST) is defined as the thick-
ness of a given interval measured at a right angle to the bedding surface in a vertical cross sec-
tion. It can be calculated by multiplying the TVT by the cosine of bed dip.
These various thicknesses are graphically illustrated in Fig. 4-24. In Fig. 4-24a, a well devi-
ated up-dip at an angle of 50 deg penetrates strata dipping 35 deg due west. The MLT of the stra-
ta is 127 ft. To correct the MLT to TVDT, the MLT (127 ft) is multiplied by the cosine of the well-
bore deviation angle (50 deg). The resultant TVDT is 82 ft.
Electric Log Correlation – Horizontal Wells 101
Figure 4-24b shows the same strata penetrated by a well drilled down-dip at an angle of 40
deg. The MLT of the strata in the well is 476 ft. Corrected to TVDT, it is 357 ft. The correction
factor equations developed in the previous section are used to calculate a TVT of 150 ft for the
penetrated strata in Fig. 4-24b. The TST of the strata calculated by multiplying the TVT (150 ft)
by the cosine of the bed dip (35 deg) is 123 ft.
Notice that for the well drilled down-dip, the MLT is 3.17 times greater than the TVT, and
the TVDT is 2.38 times greater than the TVT. For the well drilled up-dip, the MLT is less than
the TVT, and the TVDT is about one-half the TVT.
The understanding of these various measurements is very important in log correlation and
the determination of the value for the missing section. Very often, when a well is directionally
drilled, a TVD log is automatically prepared as part of the logging program. The TVD log is then
used to aid in correlation, determine missing or repeated section resulting from a fault, and to
count net sand and net pay. Figure 4-24 illustrates that in areas of significant dip, a TVD log may
provide little, if any, advantage over the deviated well log for correlation and can actually com-
plicate the estimation of missing section. Observe in Fig. 4-24b that the TVDT, which is the
thickness seen in a TVD log, is still 2.38 times greater than the TVT.
Consider a fault in this section (Fig. 4-24b) that exactly faults the entire unit out of the well.
The amount of missing section in that well would be 150 ft. By correlation with the well deviat-
ed down-dip, the missing section would be estimated to be 476 ft; with a TVD log, the estimat-
ed missing section for the fault would be 357 ft. We can conclude that the TVD log could pro-
vide a considerable error in the estimation of the amount of missing section. If the data are avail-
able, we recommend that a TVT log be prepared. This log can be used to aid in correlation and
to estimate the amount of missing section due to a fault, since it represents the TVT of the inter-
val logged.
The preparation of a TVD log for the well deviated up-dip in Fig. 4-24a could actually result
in additional correlation problems. Notice that the measured thickness of the strata in the deviat-
ed well is 127 ft. Converting this MLT to a TVD log thickness actually reduces the thickness of
the interval to 82 ft. This reduced thickness could be mistakenly interpreted as stratigraphic thin-
ning. If the TVD log were used to estimate the missing section for the fault, it would result in an
underestimate. For example, if we again use a fault in this section with a missing section of 150
ft, then by correlation with a TVD log for the well drilled up-dip, the missing section for the fault
would be estimated at 82 ft, or nearly one-half the actual size.
In areas of horizontal or nearly horizontal beds, TVDT is equal to or nearly equal to TVT
(Fig. 4-24c) and can be of significant help in correlation and estimating the value for missing sec-
tion. In areas of dipping beds, however, a TVD log may provide little help and can actually cause
additional correlation problems that may result in mapping errors.
parallel to the borehole, so it is possible to have sand on one side of the hole and shale on the
other. The tool would then be reading some averaged value for the two lithologies. Next, we
briefly discuss some of the basic methods of recognizing bed boundaries and correlating hori-
zontal wells.
log as it approaches horizontal, you may be able to recognize the target zone bed boundary.
Typically, the gamma ray and resistivity logs are used for TVD log correlation. However, if mul-
tiple bed boundaries are crossed or the borehole goes in and out of the target zone, a satisfactory
correlation of a TVD log is unlikely. A TVT log would more accurately represent the thickness
seen in the vertical well with which you are correlating, but the correction factor or reduction in
length can be extreme. The copy machine method of reducing part of a directionally drilled well
log to TVT for correlation would result in a very short, possibly unusable log section. For exam-
ple, if you used a log with a scale of 1 in. = 100 ft for a horizontal wellbore with 1 deg difference
in borehole angle and bed dip, a 1000-ft (10 in.) section would reduce to a TVT section of only
0.2 inches.
The TVD cross section can be constructed with standard cross-section methods and a few
additional considerations.
1. Plot the vertical section of the well on cross-section paper with a suitable vertical
exaggeration (e.g., 10:1).
2. Plot the bed boundaries for the part of the well as it approaches the horizontal segment of
the well and estimate the vertical thickness to the top of the objective zone using a nearby
vertical well. To determine the top of the objective zone, a pilot hole is commonly drilled
vertically to the objective stratigraphic section before drilling the horizontal part of the
hole.
3. Estimate the apparent bed dip in the plane of the cross section from maps, seismic data, or
correlations among nearby vertical wells. Draw the bed boundaries on the cross section and
try to pick the top of the objective zone where the projected top intersects the plotted
borehole.
GR Horizontal
LOG Wellbore
Vertical
Wellbore
GR
LOG
This log is upside
down and longer than
the vertical well log.
It cannot be simply
rescaled for correlation.
Figure 4-25 A gamma ray log in a wellbore that goes up stratigraphic section
is upside down compared to a log from a vertical well. (From Tearpock 1999.
Published by permission of Anadarko Petroleum Corporation.)
104 Chap. 4 / Log Correlation Techniques
4. Adjust the dip of the bed boundary if your correlation indicates a different penetration
point.
5. Continue projecting the beds and thicknesses along the cross section and look for
correlations where the wellbore would be intersecting the plotted bed boundary. Note that
if the wellbore goes up through a stratigraphic section, the log curve will be inverted
(Fig. 4-25).
6. If the wellbore cuts a fault, use the TVT in the vertical well to estimate the
missing section for the fault. That is, if your correlation shows the wellbore is 10 ft below
the top of the target zone before crossing the fault and 20 ft above the top after crossing
the fault, then the fault is a 30-ft fault. This is basically the estimated restored tops method
described later in this chapter.
In general, the TVD cross section requires that bed boundaries are recognizable on a meas-
ured depth log recorded in the horizontal borehole or at least on the computed TVD log.
Modeling of log response is often used to recognize the bed boundaries and a cross section is con-
structed to illustrate the correlations. Singer (1992) gives an example of using modeling and illus-
trating the correlations with a cross section.
TVT
G
Vertical
Thickness
between
TVD is the vertical depth
two points
TVD to each referenced point
in a well log
Different TVDs
TVD for points with
Ho the same TSD
r
W izon
ell
bo tal
re
Ta
r
Vertical Be get
d
Wellbore
TSD
Two points on the
same bedding plane
have the same TSD
Figure 4-26 A schematic vertical section along a horizontal wellbore with vertically exaggerated bed dip
illustrating TVD, TVT, and TSD. Points in a wellbore that are in the same stratigraphic position have the
same TSD. (From Tearpock 1999. Published by permission of Anadarko Petroleum Corporation.)
Electric Log Correlation – Horizontal Wells 105
TSD 1
TSD 1
TST
∆TSD φ = BED DIP ∆ TVD
DUE TO BED DIP
TOTAL TVD
Top o
f Targe
t Strat
∆TSD igrap
hic Z
one
VERTICAL
SECTION
TSD 2 BO
RE
HO
LE
VERTICAL WELL
Base
of Ta
rget S
tratig
raphic
Zone
∆ TVD
∆TSD = ( )-(
TOTAL
TVD DUE TO BED DIP )
∆TSD is the rescaled measured log thickness between
TSD points 1 and 2. This rescaled log can be correlated
back to a vertical reference well.
Figure 4-27 A schematic example of determining ∆TSD from TVD values to rescale a horizontal well log for
correlation with a log from a vertical well. This example shows a wellbore or borehole going down section.
(From Tearpock 1999. Published by permission of Anadarko Petroleum Corporation.)
106 Chap. 4 / Log Correlation Techniques
point 1 has already been correlated with the vertical well. TSD point 2 is a point in the wellbore
selected by the interpreter. From TSD point 1, the bed dip is plotted and a ∆TVD for TSD point
1 is calculated at a point vertically above TSD point 2. Subtracting this ∆TVD value from the
total TVD difference between the two TSD points leaves ∆TSD, which will be the rescaled log
thickness of the horizontal well to use for correlation (Fig. 4-27). In the computer program a slid-
er bar is used interactively to change the estimated dip and change the displayed log thickness.
The log is then stretched or squeezed until the curve is matched to the vertical well to make the
correlation. The program also generates a TVD cross section as correlations are made.
Figure 4-28 illustrates the method for determining rescaled log thickness for a wellbore
directed up section, which might occur if the wellbore extends too shallow relative to the target
zone. A deeper horizon, such as the base of the target zone, is correlatable with the vertical well,
and TSD point 2 is established and the apparent dip is drawn (Fig. 4-28). For the part of the well-
bore going up section to the selected TSD point 3, the total TVD difference between TSD points
2 and 3 and the ∆TVD for TSD point 2 add to give ∆TSD. This vertical thickness will be the
rescaled log thickness to use for correlations of the log section between TSD points 2 and 3. Of
course, the log curve will also be inverted compared to the vertical well if the wellbore is deviat-
ed up stratigraphic section (Fig. 4-25). The GRNAV program can correct an inverted log curve.
TSD 1
TSD 1
TSD 3
∆ TSD Between TSD
Top
Points 2 and 3
of T
arg
et S
tra
tigr
aph
ic Z
TSD one
2
TSD 3
le
Boreho Total
TVD
TSD 2 φ
Ba
se
of T ∆TVD ∆TSD
arg
et S Due to Bed Dip
∆TSD = ( TOTAL
TVD )+ ( ∆ TVD
)
DUE TO BED DIP
tra
tigr
aph
ic Z
one
∆ TSD is the rescaled measured log thickness
between TSD points 2 and 3. This rescaled log
can be correlated back to a vertical reference well.
Figure 4-28 A schematic example of determining ∆TSD from TVD values to rescale a horizontal well
log for correlation with a log from a vertical well. This example shows a wellbore or borehole going up
section. (From Tearpock 1999. Published by permission of Anadarko Petroleum Corporation.)
Electric Log Correlation – Horizontal Wells 107
Figure 4-29 is a generic illustration of the correlation of a horizontal well log and a vertical
well log from Kyte et al. (1994). Figure 4-29a is a gamma ray log from a nearby vertical well or
offset log with A, B, and C identifying three correlation markers. The wellbore geometry is
shown in Fig. 4-29b in a TVD cross section along the vertical section of the wellbore. A, B, and
C are the bed boundaries identified by the correlation markers in the vertical well. The wellbore
goes down stratigraphic section and up section with numbers indicating selected points illustrat-
ed in Fig. 4-29e, the horizontal well log. Figures 4-29c and d are horizontal well log segments
rescaled with the TSD method. The horizontal well log has been rescaled in Fig. 4-29c from point
1 to point 3 in the horizontal well log, Fig. 4-29e, to make the correlation to the vertical well.
From point 3 on the horizontal well log to point 4, the log has been rescaled and inverted in Fig.
4-29d, because the wellbore is going up section from points 3 to 4 (Fig. 4-29b).
The TSD method is an effective way to rescale a log from a horizontal well with a comput-
er program and correlate it to a nearby well. Good log character is required, and even with a
gamma ray log, the response may be more gradual than in a vertical well due to the low angle at
which the wellbore cuts across the bed boundary.
A A'
B 1
B'
VERTICAL WELL LOG 2
(a)
Wellbore
A
1 Line of Equal
B Stratigraphy 3
2
C
TVD C'
3
4
(b) Figure 4-29 A generic illustration of
the correlation of a horizontal well
Gamma Ray log with a vertical well log. (a)
Vertical well log. (b) A TVD cross
B' section drawn at true scale along
TSD
-A
Gamma Ray the vertical section of the wellbore.
Two gamma ray log curves have
TSD
-B 4
B been rescaled from the horizontal
well log for (c), a down-section part,
and (d), an up-section part. (e)
-C C Horizontal well log. (From Kyte et al.
DOWN-SECTION UP-SECTION 1994. Published by permission of
HORIZONTAL WELL LOG
Rescaled Log Rescaled Log Anadarko Petroleum Corporation.)
NORTH SOUTH
1 2
TYPE
LOG
MESA PET MESA OPERATING LIMITED PARTNERSH
OCSG-03587 A-5 OCS-G 03171 A-6
13 3380.3ft 13
200 Feet
Figure 4-30 Two full-width electric logs, displayed side-by-side, with a series of arbitrarily named
markers placed on the “type log.” Both logs are digital raster well log images with depth intelli-
gence added to the depth track. (Published by permission of W. C. Ross and A2D Technologies.)
MESA PET MESA PET MESA PETRO MESA OPERA MESA PETRO SKZ INC. SONAT EXPL
OCSG-03587A OCSG-03171A OCSG-03171B OCS-G 03171 OCSG-03171A OCSG-031717 OCSG-031719
13 13 3171 13 13 3171 13 13 3171 13
(a)
MESA PET MESA PET MESA PETRO MESA OPERA MESA PETRO SKZ INC. SONAT EXPL
OCSG-03587A OCSG-03171A OCSG-03171B OCS-G 03171 OCSG-03171A OCSG-031717 OCSG-031719
13 13 3171 13 13 3171 13 13 3171 13
(b)
Computer-Based Log Correlation 111
MESA PET MESA PET MESA PETRO MESA OPERA MESA PETRO SKZ INC. SONAT EXPL
OCSG-03587A OCSG-03171A OCSG-03171B OCS-G 03171 OCSG-03171A OCSG-031717 OCSG-031719
13 13 3171 13 13 3171 13 13 3171 13
(c)
Figure 4-31 (a) Cross-section window configured to simulate paper-based log correlation. Closely
spaced log windows, which here contain spontaneous potential and gamma ray curves centered with-
in them, simulate the log overlay techniques used in paper-based methods. The type log contains a
series of arbitrarily named markers to be used as starting points in a log correlation exercise. The wells
in the cross section have an average spacing of less than one-quarter mile and were constructed with
a combination of raster and digital (Log ASCII Standard) well log data. (b) Partially correlated panel of
well logs illustrating the progression of correlation work from the central type log. (c) Completed corre-
lation panel of well logs. (Published by permission of W. C. Ross and A2D Technologies.)
correlation. In our example, we place the SP and gamma ray curves in this window. By creating
narrow windows with single log curves (or single log tracks), a computer-screen display can sim-
ulate paper well-log overlay techniques and effectively increase the number of well logs that can
be displayed on the screen at one time.
Computer log correlation can now proceed by simply pointing and clicking on each well log
with the correlation tool at the various points of log correlation. Log slipping (or vertical scroll-
ing) can be performed to help line up each of the logs prior to correlation itself.
To simulate the log annotation techniques used in paper-based correlation, the following pro-
cedure can be used. First, nominate a type well for the starting point in your correlation exercise.
Next, place on the type log a series of markers (Fig. 4-31a). They are designated by an arbitrary
naming convention, such as GM1 (generic marker 1), GM2, GM3 and their purpose is analogous
to manual annotations on paper well logs. As with paper-based annotation, the markers should be
placed with as much frequency as is needed to capture all the log character deemed necessary for
detailed correlation.
Once the markers are placed on the type well (the starting point), the geologist can begin
transferring each marker from well log to well log across a panel of wells (Fig. 4-31b and c). The
112 Chap. 4 / Log Correlation Techniques
TYPE
WEST LOG EAST
1 2 3 4 5 6 7 8 9 10
ANADAR DAVIS O GULF OI TERRA R DAVIS O DAVIS O PETROL ARGO OI UPRR-N CHAMPL
UPRR-A DODSON SKYLINE UPRR 13- ARMOR- BARTA F GOVT-C GOVT-E UPRR 2-1 UPRR 3D
021.0N 10 021.0N 10 021.0N 10 021.0N 09 022.0N 09 022.0N 09 021.0N 09 021.0N 09 021.0N 09 021.0N 09
B2
B1
200 Feet
(a)
TYPE
WEST LOG EAST
1 2 3 4 5 6 7 8 9 10
ANADAR DAVIS O GULF OI TERRA R DAVIS O DAVIS O PETROL ARGO OI UPRR-N CHAMPL
UPRR-A DODSON SKYLINE UPRR 13- ARMOR- BARTA F GOVT-C GOVT-E UPRR 2-1 UPRR 3D
021.0N 10 021.0N 10 021.0N 10 021.0N 09 022.0N 09 022.0N 09 021.0N 09 021.0N 09 021.0N 09 021.0N 09
B2
B1
200 Feet
(b)
Computer-Based Log Correlation 113
TYPE
WEST LOG EAST
1 2 3 4 5 6 7 8 9 10
ANADAR DAVIS O GULF OI TERRA R DAVIS O DAVIS O PETROL ARGO OI UPRR-N CHAMPL
UPRR-A DODSON SKYLINE UPRR 13- ARMOR- BARTA F GOVT-C GOVT-E UPRR 2-1 UPRR 3D
021.0N 10 021.0N 10 021.0N 10 021.0N 09 022.0N 09 022.0N 09 021.0N 09 021.0N 09 021.0N 09 021.0N 09
B2
B1
200 Feet
(c)
Figure 4-32 (a) West-east stratigraphic cross section, hung on the B1 datum, constructed with conductivity
curves centered within narrow well log windows. The light gray shading depicts sands within the Almond
Formation. The darker gray pattern represents sands and shales of the non-marine Almond Formation and the
white region depicts shale facies within the Lewis Shale Formation. This cross section was constructed with well
log images. Style of drawing correlations is from log edge to log edge. The cross section is eighteen miles in
length with an average well spacing of less than two miles. The vertical exaggeration is 166:1. (b) Same cross
section with correlation draw style changed to center-to-center of logs, which accentuates basin-fill geometries.
Downlap on surface B1 is suggested. (c) Same cross section hung on the B2 stratigraphic datum. The higher
datum changes the geometry of the unconformable relationship and suggests onlap versus downlap. (Published
by permission of W. C. Ross and A2D Technologies.)
correlation work proceeds by correlating all the wells in the study area using this level of detail.
Whereas the level of detail may not be any higher than the paper-based log annotation techniques,
computer-based correlation of markers carries the advantage that each marker top is automati-
cally placed into a digital database of tops for each well. Consequently, as you build arbitrary
cross sections through your three-dimensional database of well control, all the marker correlation
work will carry forward onto these cross sections. The power of this technology for geometric
cross-sectional analysis, fault picking, and unconformity recognition is considerable, as we illus-
trate in the next section.
A subtle unconformity was discovered by careful well log correlation within the transgres-
sive Lewis Shale, just above the Almond Formation. To illustrate how this unconformity was
detected, we begin with a 10-well, west-to-east cross section constructed parallel to depositional
dip across the Almond and Lewis Formations (Fig. 4-32a). The shallow-marine Almond
Sandstone and the marine Lewis Shale represent the transgression of the Lewis seaway from east
to west (right to left) on the section.
The on-screen correlation involved the use of arbitrarily named markers to perform high-res-
olution time-stratigraphic correlations, following the sequence stratigraphic methodology of Van
Wagoner et al. (1990). Following this strategy, we established a type well at the Davis Oil Armor
State #1 (fifth well in cross section) and began correlating the initial markers across the panel of
well logs. In the early stages of log correlation it became apparent that there were distinct pack-
ages of rock within the Lewis Shale that could easily be distinguished and shown to be discor-
dant with respect to one another along an unconformity surface (Fig. 4-32a). The discordant sur-
face coincides with the top of a regional bentonite marker denoted as B1. The geometric discor-
dance is enhanced through extreme vertical exaggeration (V.E. = 166:1). The vertical exaggera-
tion is facilitated on a 21-inch computer screen by using narrow log windows, close well spac-
ing, and by changes in correlation draw style. Figure 4-32a and b illustrate the visual advantages
gained by changing the log correlation draw style from a conventional style where correlations
are drawn straight across the log (i.e., edge-to-edge) to a style where the correlations are con-
nected from the center point of each log (i.e., center-to-center). The use of center-to-center cor-
relation drawing style smooths the connections between correlation points and provides a clear-
er view of both discordant relationships and subtle basin-fill geometries. Upon initial inspection,
the discordant relationship (correlation markers terminating in an easterly or basinward direction
against a common surface) was interpreted as depositional downlap (Fig. 4-32b). Computer-
based cross-sectional analysis allows the interpreter to quickly choose an alternative cross-sec-
tion datum to evaluate alternative interpretations of unconformity type. For example, in Fig.
4-32c a higher marker (B2) is utilized as the stratigraphic datum, providing a view that would
suggest the unconformity is an onlap unconformity.
NORTH SOUTH
1 2 3 4 5
MESA PET MESA PET MESA PETROLEU SKZ INC. SONAT EXPL INC
OCSG-03587A-5 OCSG-03171A-3 OCSG-03171B-4 OCSG-031717 OCSG-031719
13 13 3171 13 13 3171 13
(a)
NORTH SOUTH
1 2 3 4 5 6
MESA PET MESA PET MESA PETROLEU MESA OPERATING SKZ INC. SONAT EXPL INC
OCSG-03587 A-5 OCSG-03171 A-3 OCSG-03171 B-4 OCS-G 03171 A-6 7
OCSG-03171 OCSG-03171 9
13 13 3171 13 13 13 3171 13
(b)
Figure 4-33 Cross section illustrating the technique to determine missing section. (a) Cross section with numer-
ous markers demonstrating a continuous, nonfaulted stratigraphic section. Average well spacing is less than one-
quarter mile. (b) Same cross section with an additional well (Well No. A-6) that has missing section. The detailed
correlation work demonstrates that the A, B and C correlations are cut out by a fault (i.e., where correlations con-
verge). (c) Same cross section with fault picked and log partially pulled apart (gapped) for Well No. A-6. (d) Same
cross section with missing section interval fully interpreted and missing tops (A, B, and C correlations) placed
within it. (Published by permission of W. C. Ross and A2D Technologies.)
116 Chap. 4 / Log Correlation Techniques
NORTH SOUTH
1 2 3 4 5 6
MESA PET MESA PET MESA PETROLEU MESA OPERATING SKZ INC. SONAT EXPL INC
OCSG-03587A-5 OCSG-03171 A-3 OCSG-03171 B-4 OCS-G 03171 A-6 OCSG-03171 7 OCSG-03171 9
13 13 3171 13 13 13 3171 13
GAP
(c)
NORTH SOUTH
1 2 3 4 5 6
MESA PET MESA PET MESA PETROLEU MESA OPERATING SKZ INC. SONAT EXPL. INC
OCSG-03587A-5 OCSG-03171A-3 OCSG-03171B-4 OCS-G 03171A-6 OCSG-031717 OCSG-031719
13 13 3171 13 13 13 3717 13
2059.0ft 1653.1ft 1419.2ft 1643.4ft 1211.4ft
200 Feet
(d)
Repeated Section 117
The log-splitting tool allows you to create correlation cross sections hung on stratigraphic
markers, with missing section (due to faulting or erosional truncation) displayed as missing sec-
tion gaps. Once the gaps are interpreted, you can pick tops within the missing section to allow
the cross section to be hung on missing top markers (Fig. 4-33d). The hand-generated equivalent
to this computer method is shown in Fig. 4-20b, for which a paper log was cut and pulled apart
by the amount of missing section.
The use of high-resolution marker correlations, narrow window correlation displays, and
interactive fault picking/gapping tools allow geologists to correlate hundreds of well logs within
a short period of time. Correlation exercises are significantly aided by this fault-gapping tech-
nology. By picking and gapping faults during a correlation exercise, significant amounts of dis-
tortion related to missing section can be eliminated (compare Fig. 4-33b and d).
There are many advantages of computer-based log correlation over paper-based techniques.
Perhaps the most significant are the capabilities for instant access to log data from a computer-
ized database, automatic rescaling of images and digital data, rapid redatuming capabilities, and
the ability to track and display hundreds of correlations/markers within a computerized database.
For many traditional geologists, the transition onto the computer will be a painful process.
However, many of the concerns can be overcome by adopting some of the techniques described
in this section.
Many geologists have expressed concern about being unable to see enough of the well log
on the computer screen at one time. This is a legitimate concern. However, technology has
improved dramatically in recent years, providing both high-resolution computer monitors and
image-enhancement technologies, which allow several thousands of feet of section to be viewed
on the computer screen for reconnaissance log review. Ultimately, geologists should agree that
the advantages of migrating onto the computer outweigh the disadvantages of learning a new set
of skills. As with geophysicists, who made the transition to geophysical workstations in the mid-
1980s, geologists are rapidly making the transition to geologic workstations. The enormous
increase in productivity for geologists will soon be on par with their geophysicist colleagues.
REPEATED SECTION
A repeated section in a well is defined as a part of the stratigraphic section that appears twice on
a log as the result of a fault, consequently lengthening the log section. A repeated section is com-
monly thought of as a compressional tectonic phenomenon, occurring as the result of a reverse
fault pushing the stratigraphic section in the hanging wall up and over the same section in the
footwall. Figure 4-34 illustrates the geometry required to result in a repeated section due to a
reverse fault. A repeated section can also occur with a normal fault. In this situation, a repeated
section requires a specific geometry between a normal fault and a directionally drilled well, or a
normal fault that cuts a structure dipping at a steeper angle than the fault.
So far in this book, we have looked at vertical and directionally drilled wells penetrating a
fault in what is called the normal sense; that is, from the hanging wall (downthrown) fault block
to the footwall (upthrown) fault block (Fig. 4-35). Figure 4-36 is a cross section illustrating the
geologic and deviated well parameters required to cause a repeated section in a log for a situa-
tion involving a normal fault. Geometries, as shown in Fig. 4-36, are not uncommon in areas of
directionally drilled wells. In offshore areas, deviated wells are drilled from a central platform
location and are often designed to parallel a known fault (Fig. 4-37). In this way, a well may be
drilled to penetrate a series of potential strata in the footwall (upthrown) block of an important
trapping fault. If care is not taken to precisely map the fault and drill the well, it is possible for
118 Chap. 4 / Log Correlation Techniques
the well to cross the fault and enter the hanging wall (downthrown) block of the fault. In Fig.
4-37, the well deviated toward the salt crosses a normal fault backwards, resulting in a repeated
section. Also notice that by crossing the fault, the well does not penetrate productive “Sand B,”
which is upthrown to the fault, but instead penetrates the nonproductive “Sand B” downthrown.
Figure 4-38 is an example of a repeated section resulting from an extensional (normal) fault.
The amount of wellbore deviation is shown on the right of the log section. Notice that the devi-
ation angle ranges from 40 deg 45 min to 42 deg 15 min. The well was drilled to parallel a fault
and to stay in the footwall fault block. Since wellbore deviation is measured from the vertical and
fault dip from the horizontal, it is necessary to subtract the hole angle from 90 deg to compare its
“dip,” or plunge, angle to that of the fault. By doing this, the well angle measured from the hor-
izontal, for the section shown, ranges from 47 deg 45 min to 49 deg 15 min. Based on the fault
VERTICAL DIRECTIONAL
WELLBORE WELLBORE
(a) (b)
Figure 4-35 (a) A vertical well intersecting a fault in the normal sense (from hanging wall
to footwall fault blocks). (b) A deviated well intersecting a fault in the normal sense (from
hanging wall to footwall fault blocks).
Repeated Section 119
REPEATED SECTION
FAULT SURFACE Figure 4-36 The fault and deviated well parameters required
for a repeated section resulting from a normal fault.
Figure 4-37 The well deviated toward the salt cuts the normal fault backwards (from the
footwall block to the hanging wall block), resulting in a repeated section.
surface map, the fault dip is approximately 52 deg. Therefore, the well inclines downward at an
angle less than the fault dip, establishing a well/fault geometry similar to that shown in Fig.
4-36. The result is the repeated section shown in Fig. 4-38.
Figure 4-39 is another excellent example of a repeated section resulting from a normal fault
cut backwards by a deviated well. Observe that the footwall fault block is productive of hydro-
carbons, whereas the hanging wall block is not.
120 Chap. 4 / Log Correlation Techniques
Figure 4-38 Repeated section in the Texaco LL&E Well No. 212 due to the well
cutting a normal fault backwards. (Published by permission of Texaco, USA.)
Repeated Section 121
Figure 4-39 Repeated section in a deviated well due to the well cutting a normal fault
backwards. Notice that the footwall (upthrown) fault block contains hydrocarbons, while the
hanging wall (downthrown) block is wet. (From Tearpock and Harris 1987. Published by per-
mission of Tenneco Oil Company.)
122 Chap. 4 / Log Correlation Techniques
The identification of a repeated section that has previously gone unrecognized can result in
excellent prospect potential. Look at the generic situation shown in Fig. 4-40. In this case, a well
was designed to stay in the footwall of an important trapping fault. Good control on the fault was
available at the time of drilling, but a fault surface map was not prepared. The well meandered
across the fault, resulting in an unrecognized repeated section on the well log. The stratigraphic
intervals penetrated in the footwall of the fault are productive of hydrocarbons, whereas the hang-
ing wall interval from A to B are wet. The repeated section went unrecognized and, consequent-
ly, the wet intervals penetrated in the hanging wall were thought to be in the footwall and were
condemned as nonproductive.
If an observant geoscientist recognizes the repeated section and understands that the non-
productive log section represents the well penetration in the hanging wall block, a new prospect
can be generated in an area previously condemned. In Fig. 4-40 the new prospect includes the
stratigraphic section c through e, present in the footwall block of the fault and not seen in the
deviated well.
FA
-7000'
UL
T
B
-8000'
A
b
c
-9000'
d
B
e
-10,000'
OIL f
UNTESTED
g
Figure 4-40 The identification of a previously unrecognized repeated section can result in a
potential prospect. In this example, Sands C through E are untested upthrown to Fault B.
Estimating Restored Tops 123
Vertical Wells
The procedure for determining the upthrown and downthrown restored tops (sometimes referred
to as equivalent points) in vertical wells is discussed here and illustrated in Fig. 4-42.
1. Figure 4-42 shows three vertical well logs positioned side by side with no horizontal scale.
By correlation, we determine that there is a 150-ft fault in Well No. 2 and that the
objective Unit C is faulted out. Well No. 1 is in the upthrown fault block and Well No. 3
is in the downthrown block. To determine the upthrown and downthrown restored tops,
these two nearby wells are used.
2. Downthrown Restored Top for Unit C. To estimate the downthrown restored top, Well No.
3 in the downthrown fault block is used to correlate with Well No. 2. Identify a good
marker, such as a field-wide sand or resistivity marker, in both logs above the faulted-out
section. In Fig. 4-42, the “B” marker is used as the marker in the wells. Line up the
marker on both logs and correlate down the logs until the top of “C” is reached in Well
No. 3. The equivalent measured depth reading for the top of “C” in Well No. 2 is 6855 ft.
This represents the estimated depth for the top of “C” if the well were in the downthrown
(hanging wall block).
3. Upthrown Restored Top for Unit C. To estimate the upthrown restored top, Well
No. 1, which is in the upthrown block, is used to correlate with Well No. 2. Identify a good
marker in both wells below the faulted-out section. In this case it is the “D”
marker. Line up the “D” marker on both logs and correlate up the logs until the top of “C”
is reached in Well No. 1. This equivalent depth reading for the top of “C” in Well No. 2 is
6705 ft. This represents the estimated depth for the top of “C” if the well were in the
upthrown block.
The accuracy of the restored top estimates can be checked by subtracting the upthrown
restored top from the downthrown restored top and then comparing the difference to the estimat-
ed value of the missing section for the fault. In this case, 6855 ft – 6705 ft = 150 ft. The differ-
ence in the restored tops and the amount of missing section in Well No. 2 are exactly the same.
Therefore, we conclude that the estimated depths for the restored tops are reasonable.
A similar method to restore a top is to calculate the vertical distance that the top should be
124 Chap. 4 / Log Correlation Techniques
above or below the fault in the well. In the example shown in Fig. 4-42, a point in upthrown Well
No. 1 can be picked that is correlative to the fault pick in Well No. 2; it would be just above “D.”
The true vertical thickness between that point and the top of Unit C can be determined. That
thickness of 105 ft is equal to the vertical distance between the fault and the upthrown restored
top of “C” in Well No. 2. In fact, that TVT can be used to calculate the restored top
105 ft above the fault, at 6705 ft, in Well No. 2. Similarly, the depth of the downthrown restored
top in Well No. 2 can be determined by using the 45-ft TVT of the interval between the fault-cor-
relative point and the top of Unit C in downthrown Well No. 3. In the next section, we use the
TVTs of such reference intervals to restore the top of a unit faulted out of a deviated well.
In the example illustrated in Fig. 4-42, the section in the three wells is approximately equal
in thickness. What if a reference well is of different thickness than the well with a fault? That
could result in an incorrect restoration of a top, so an adjustment to the thickness measured in the
reference well is necessary. Simply apply a thickness ratio. Determine an average thickness ratio
between the wells for stratigraphic intervals immediately above and below the fault in the fault-
ed well. Then use that ratio to modify the TVT of each reference interval as measured in the ref-
erence well. For example, if the stratigraphic section in the faulted well is 0.80 as thick as that in
the reference well, then multiply each reference interval measured in the reference well by 0.80
to provide a more accurate estimate for restoring each top.
WELL NO. 3
-5500
-6000
Upthrown Restored
Top -6230'
65 -6500 65
00 00
'S 'S
AN AN
D D
Downthrown Restored
Top -6910'
-7000
-7500
Figure 4-41 Basic concept of restoring tops. The upthrown restored top is the estimated
depth in the well for the mapped horizon if the well were in the upthrown fault block. The
downthrown restored top is the estimated depth in the well for the horizon if the well were
in the downthrown block.
Estimating Restored Tops 125
Figure 4-42 The method for estimating restored tops in vertical wells is shown by special
correlation of the three vertical wells.
The two restored tops should be honored in structure contouring of any mapped surface,
whether it is a marker, sand top, particular stratum, etc. If the map is referenced to sea level,
measured log depths must be converted to subsea depths for structure mapping. The specific tech-
nique to apply the restored tops in structure mapping is discussed in Chapter 8.
Deviated Wells
All the wells are vertical in the example shown in Fig. 4-38. Where deviated wells are present,
the estimation of restored tops requires additional steps because TVTs must be used instead of
the MLT in the deviated wells.
There are three separate cases for log correlation involving deviated wells: (1) a fault in a
deviated well correlated with a vertical well; (2) a fault in a vertical well correlated with a devi-
ated well; and (3) a fault in a deviated well correlated with another deviated well. Previously, we
showed that missing section is defined in terms of TVT, and we also used TVT in restoring tops
in vertical wells. When estimating restored tops in deviated wells, TVT must be used. Probably
the most accurate way to estimate restored tops in deviated wells is to convert the deviated logs
to TVT logs. In order to make a TVT log, bed dip and azimuth data are required in addition to
the well deviation angle and well azimuth. If time, costs, or other factors prohibit the preparation
of TVT logs, the restored tops can still be estimated with reasonable accuracy using the deviated
well logs and applying a correction factor to derive TVTs. The correction of MLT to TVT was
described in the section Estimating the Missing Section for Faults.
126 Chap. 4 / Log Correlation Techniques
For case No. 1, in which a fault in a deviated well is correlated with a vertical well, the same
basic procedure for estimating a restored top using two vertical wells can be applied, with one
significant variation. The depth of a restored top for the faulted-out unit cannot be transferred
(projected) directly to the deviated well log from the top of the unit in the vertical well log, as it
was in Fig. 4-42, because the deviated well log depths are not vertical depths and the thickness
of a given interval is not a TVT. A calculation for depth of each restored top must be made, and
it is based on the TVT of a reference interval in the vertical well.
The procedure for restoring a top faulted out of a deviated well is shown in Fig. 4-43. A fault
is identified in the deviated Well No. 3 at a depth of –6810 ft, with a missing section of 100 ft
(Fig. 4-43a). The B Sand is faulted out of the well. We wish to restore the upthrown and down-
thrown tops for the B Sand. Well No. 1 in the downthrown block and Well No. 2 in the upthrown
block are available to estimate the restored tops.
(a)
Figure 4-43 (a) The two vertical wells (Nos. 1 and 2) and the deviated well (No. 3) illustrate case No. 1, estimating
restored tops for a unit faulted out of a deviated well. (b) A structural cross section containing the three wells shown
in (a); it illustrates the precise positions of the restored tops with respect to Well No. 3: directly above (upthrown
restored top) and directly below (downthrown restored top) the actual fault pick in the well. The top of Sand B in each
fault block is projected to the appropriate restored top to establish the structural dip shown in the cross section.
Estimating Restored Tops 127
1. Upthrown Restored Top for the B Sand. To estimate the upthrown restored top, identify a
marker below the B Sand in Well No. 2 that is also in the faulted well. For this example,
the lower marker used and shown in Well No. 2 is exactly at the fault in Well No. 3. By
correlating up the logs, we determine that the top of the B Sand is 80 ft above the lower
marker in Well No. 3. Since the marker is exactly at the fault in Well No. 3, the upthrown
restored top for the B Sand in the well is 80 ft above the subsea depth of the fault. The
subsea depth of the fault is –6810 ft; therefore, the upthrown restored top is –6730 ft. It
is positioned vertically above the fault in the deviated well rather than within the inclined
wellbore (Fig. 4-43b).
2. Downthrown Restored Top for the B Sand. The same procedure is followed to estimate the
downthrown restored top. Using Well No. 1 in the downthrown fault block, identify a
marker above but as close to the fault cut as possible. In this case, the upper marker used
in Well No. 1 is exactly at the fault in Well No. 3. Correlate down the two logs from the
upper marker until the top of the B Sand is reached in Well No. 1. Using the TVT of that
interval, the restored top is calculated to be 20 ft vertically beneath the fault in Well No. 3,
at –6830 ft (Fig. 4-43b).
The depths for the restored tops were not projected directly from faulted Well No. 3, as was
done in the vertical well example (Fig. 4-42). Instead, for the upthrown restored top, the vertical
distance from the fault-correlative point to the top of the sand in unfaulted Well No. 2 was “sub-
tracted” from the subsea depth of the fault to estimate the upthrown restored top. Similarly, the
vertical distance from the fault-correlative point to the top of the sand in unfaulted Well No. 1
was “added” to the subsea depth of the fault to estimate the downthrown restored top.
Observe that the correlation marker used to determine each restored top was chosen as the
closest marker to the fault. It is recommended that, whenever possible, you choose a correlation
marker as close to the fault as possible in order to minimize any errors due to correlation or due
to stratigraphic variations between the wells being used.
In Fig. 4-43b, which is a structural cross section through the three wells, an X is placed 80
ft vertically above and 20 ft vertically below the fault in Well No. 3. These X’s show the precise
position of the upthrown and downthrown restored tops for the B Sand. It is important to under-
stand that the restored tops are located vertically above and below the location of the fault in the
well and not within the deviated well either in cross-section view or in map view. This under-
standing is critical when using these restored tops in structure mapping.
The 100-ft difference between the downthrown and upthrown restored tops agrees with the
100 ft estimated for the missing section in the well. Therefore, we conclude that the depths for
the estimated restored tops are reasonable.
For cases 2 and 3, in which a fault in a vertical well is correlated with a deviated well or a
fault in a deviated well is correlated with another deviated well, the procedure is basically the
same as for case No. 1 with one exception. When an unfaulted deviated well is used as a refer-
ence, the TVT from a marker to the mapped (faulted) horizon must be calculated to accurately
estimate the restored tops. The measured (deviated) log thickness cannot be used to properly esti-
mate restored tops. For example, if Well No. 2 in Fig. 4-43a were a deviated well, the interval
thickness of 80 ft measured from the upper marker to the top of Sand B would have to be cor-
rected to TVT before calculating the subsea depth of the upthrown restored top.
UNCONFORMITIES
Unconformities are present in all geologic settings, especially on steeply dipping growth struc-
tures. Excellent hydrocarbon traps can occur at unconformities where overlying impermeable
rocks form seals for reservoirs. Therefore, it is important to recognize, analyze, and often map
unconformities in the subsurface.
There are a host of interrelated versions of the broad term unconformity. Some are primari-
ly erosional, some are nondepositional, and others are combinations of both. The subject of
unconformities is extensive and beyond the scope of this book. We present important information
on the recognition of unconformities during electric log correlation.
Annotation and Documentation 129
An unconformity appears on an electric log as missing section. Since missing section can be
due to an unconformity or to a normal fault, care must be taken during correlation so that an
unconformity is not mistaken as a fault. In this section we discuss several general guidelines to
follow during correlation to recognize an unconformity.
1. Structural dip commonly is different above and below an unconformity. Dipmeter data can
be used to indicate this change in dip. The structural dip below an unconformity is
usually steeper (Fig. 4-44).
2. If missing section is recognized in two or more wells at the same or nearly the same
correlative horizon, such as “C” in Fig. 4-44a, an unconformity, rather than a fault, should
be suspected.
3 The amount of missing section resulting from an unconformity increases in the up-
structure direction. This is illustrated in Fig. 4-44. The missing section in Wells No. 1, 2,
and 3 increases in the up-structure direction; Well No. 3 has the least amount of section
missing and Well No. 1 has the greatest amount missing.
4. The stratigraphic section below an unconformity is truncated in a younger sequence in
down-structure wells than in up-structure wells. The truncated sequence becomes older in
the up-structure direction. In Fig. 4-44a, the sedimentary sequence just below the
unconformity in Well No. 3 (the G interval) is younger than the K interval in Well No. 1.
To recognize the sequence trend, you must correlate up the well logs.
5. If the depositional environment results in an onlap sedimentary sequence (Fig. 4-44b)
rather than the sequence shown in Fig. 4-44a, the missing section in logs correlated in the
up-structure position increases above and below the unconformity. The sedimentary
sequence just above the unconformity is younger in the up-structure direction.
6. Unconformities must be mapped. Since an excellent hydrocarbon trap can occur at an
unconformity, a map of the unconformity is vital. In order to identify the intersection of the
unconformity and the underlying sedimentary sequence, an unconformity map must be
integrated with structure maps of relevant surfaces within that sequence. The mapping
techniques required are discussed in Chapter 8.
Figure 4-45 illustrates the use of a dipmeter to aid in the identification of an angular uncon-
formity. The figure also shows a dipmeter response to beds affected by the proximity of salt with
an overhang. The dipmeter reaches a maximum dip of 62 deg at about –6500 ft; thereafter, the
dip slightly flattens to 40 deg at TD. This change in dip is in response to the proximity to the salt
face. Such dipmeter responses can be used in evaluating a salt feature for possible overhang, with
potential stratigraphic section beneath it.
In the mid-1990s, Dr. Richard Bischke developed a powerful new technique for evaluating
structural and stratigraphic information from well log as well as seismic data. This technique,
today known as the Multiple Bischke Plot Analysis (MBPA), has numerous applications (Bischke
1994b; Bischke et al. 1999). One application of the technique is to assist in the differentiation of
unconformities from faults in well logs where missing section is present. It is a rapid, robust, and
powerful method for application with unconformities, as well as validation of general log corre-
lations (see Chapter 13).
(a)
(b)
Figure 4-44 (a) Well log correlation can be used to recognize an angular unconformity. (b)
If an onlap sedimentary sequence is deposited above an angular unconformity, certain log
correlation guidelines can be used to recognized the unconformity.
Annotation and Documentation 131
Figure 4-45 Dipmeter data can be used to recognize an unconformity in the subsurface.
(Published by permission of Tenneco Oil Company.)
good quality maps depends upon a volume of accurate data. These data include faults, horizons,
tops and bases of stratigraphic units, net sand, net pay, and so on. In this section, we illustrate a
recommended method of annotating 1-in. and 5-in. electric logs and documenting the log data.
In Fig. 4-2 we illustrated the importance of marking logs with recognizable symbols and the
use of color. These markings are a form of annotation. They identify your correlations. The logs
in Fig. 4-2 are of intervals that do not contain any hydrocarbons. There are some additional data
that should be annotated on a 1-in. or 5-in. log with recognized pay. Figure 4-46 illustrates the
additional data that should be annotated in the pay section on a 1-in. log. The annotation includes
the name of each pay zone, perforation intervals, well status, cumulative production from each
interval produced and a note on why each interval went off production, and the measured and
subsea depth for the top of each important productive interval. Any intervals that appear produc-
tive on the log but have not been produced should be noted, such as Sands 2 and 6A in Fig. 4-46.
Finally, any recognized faults must be indicated on the log, such as Fault F, which faults out Sand
3 at a log depth of 8350 ft.
The annotation of a detailed 5-in. log is crucial, and an example is shown in Fig. 4-47. The
132 Chap. 4 / Log Correlation Techniques
information annotated includes name of productive zone measured and subsea depth of the cor-
related tops, perforation intervals and corresponding production, well status, net pay counts, limit
of pay (full to base of interval or water contact), and basic core data (at least the porosity and per-
meability data). Notice that the net pay on the 5-in. log in Fig. 4-47 is assigned per 10-ft inter-
vals on the left side of the log. This annotation is used to support the net pay count and later in
preparation of net pay isochore maps.
Finally, the documentation of well data means recording it in some format that can be easi-
ly used. There are various types of manual and computer data sheets available for documenting
mapping parameters. These data sheets should be used at all times to document the log correla-
tion data.
Figure 4-46 Electric log annotation in a hydrocarbon-bearing (pay) section. (From Tearpock
and Harris 1987. Published by permission of Tenneco Oil Company.)
Annotation and Documentation 133
Figure 4-47 Annotation of a detailed 5-in. log. (Modified from Tearpock and Harris 1987.
Published by permission of Tenneco Oil Company.)
CHAPTER 5
INTEGRATION
OF GEOPHYSICAL DATA
IN SUBSURFACE
MAPPING
134
Introduction and Philosophy 135
may not be familiar with seismic data and, indeed, may not understand how seismic data are
acquired and processed. We focus on practical approaches to using seismic data in the search for
hydrocarbon traps. The technical details of seismic acquisition and processing are beyond the
scope of this book as is the topic of theoretical geophysics. These are very important subjects that
a working geoscientist must understand. Many interpreters who have access to seismic data are
not geophysicists. However, it is our intent to illustrate techniques that will make the non-geo-
physicist comfortable with using these data in the construction of subsurface maps.
This chapter should make it obvious that valuable information is present in seismic data and
that an interpretation that properly integrates the subsurface geologic data with the seismic data
is always more accurate than an interpretation that ignores one of these data sets. It will soon
become apparent that the discussion has a strong regional bias in that most of the examples are
from the offshore Gulf of Mexico. There are several reasons for this. Perhaps the most obvious
reason is that this region has a greater abundance of high-quality seismic data than anywhere else
in the world. This fact means that (1) it is easier to get good examples from this region than from
most others, and (2) this region is highly prospective, which increases the likelihood that North
American geoscientists will work in this region at some point during their careers.
This regional orientation does not mean that the techniques outlined are limited to the Gulf
of Mexico. In fact, the techniques presented here can be used to establish the three-dimensional
geometric validity for subsurface maps in any tectonic environment, anywhere in the world.
Figure 5-1 Sampling differences between wells and surface seismic events. Well data are single
points, and correlations of points between wells must be inferred. Seismic events, however, expli-
citly demonstrate horizon continuity.
high bed dips, and extremely complex geology, all of which may invalidate many assumptions
necessary for the acquisition of good data. If these problems are present, they may present a chal-
lenge for even the best geophysicist. In such instances, the expertise of a geophysicist and struc-
tural geologist may be necessary to solve the complexities. In areas of complex faulting and
structure, a 3D data set may help resolve the structural complexities. However, even in these com-
plex areas, 2D seismic data may contain valuable information that can be used in creating a rea-
sonable subsurface interpretation. Examples of the usefulness of seismic data in these complex
areas are presented in the section that covers structural balancing (Chapter 10).
The techniques and parameters employed in field acquisition of seismic data can influence
the quality of data, so they should be based on knowledge of the geology of the area and the
acquisition procedures that are most suitable. Careful planning can result in achieving the best
possible data set, given the geologic constraints.
THE PROCESS
The procedure for making subsurface maps from seismic data is similar to the sequence of steps
used in constructing interpretations from subsurface well log information. The first step is one of
data validation; i.e., analysis of what the seismic data represent. Do the seismic data actually have
The Process 137
Figure 5-2 Effect of subsurface structure on actual subsurface reflection path of line. Subsurface
reflection points do not occur vertically beneath the surface location of the lines where reflected
from dipping subsurface horizons.
some relationship to the geology in the subsurface? This procedure is similar to the checking one
does when a log is first used. In the case of log data, decisions about the validity and meaning of
the log response must be made before the data can be used to form an interpretation.
The second step is the actual interpretation of the seismic section. This step is analogous to
the correlation of well logs when using subsurface well log information. Because the validity of
the remaining work rests on having an accurate and geologically correct interpretation of the seis-
mic data, validating the data is the most important part of the process.
Some aspects of seismic interpretation as they relate to the construction of subsurface maps
are covered in this chapter. However, we do not attempt to cover the subject exhaustively. Several
excellent books on seismic interpretation are listed in the references and are recommended to
those who may be unfamiliar with seismic interpretation techniques (Badley 1985). Just as a
basic knowledge of well logs is needed to use log data properly, a basic understanding of the
reflection seismic method is needed before seismic data can be interpreted correctly.
The third step involves extracting the information from the seismic data and transferring it
onto the map so that it can be used effectively. The 2D seismic workstation process collects data
in relation to a base map. Usually, transferring the data to a map is referred to as posting. This
procedure is practically identical to that used when recording subsurface well log data. As seis-
mic sections have a 2D aspect that well log data do not possess, there are some unique aspects to
138 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
this step when using seismic data. This step also represents the merger of the subsurface well log
and the seismic information. Both types of data should be posted and used to construct the final
interpretation. If you don’t understand the seismic data and require assistance, experienced inter-
preters are usually available. There is typically some usable information on even the worst seis-
mic data that can add to the confidence and validity of your final subsurface interpretation and
accompanying maps. A valid interpretation should agree with and satisfy all types of informa-
tion. If 3D data are available, all the data should be used.
We appeal to geophysicists as well as geologists to work in a synergistic manner. Very often,
there are two sets of interpretations and maps: the geological one and the geophysical one. A sub-
surface interpretation and maps should accurately represent the subsurface geology, incorporat-
ing both well log and seismic data into a seamless interpretation. There is only one configuration
of the subsurface, and it is the job of the interpreters (geologists and geophysicists) to create an
integrated and reconciled interpretation using all the data available.
The last step is the construction of the subsurface geologic maps. This step represents the
culmination of all previous work, and in many instances it will be the result by which your work
is measured. Any subsurface map is only as good as the information it contains, so do not rush
to begin this step before the previous steps are completely finished.
In practice, however, constructing a map is never the last step. Several iterations of valida-
DATA VALIDATION
DATA INTERPRETATION
DATA EXTRACTION
MAPPING
REVIEW
tion, interpretation, and mapping are typically necessary before a satisfactory subsurface map is
completed. Figure 5-3 is a conceptual flow diagram of this process. At some point, it will become
apparent that most of the major questions have been resolved and a satisfactory interpretation and
maps have been made. While pride and satisfaction in the result is deserved, always keep in mind
that additional data from either drilling or additional seismic acquisition will almost always
change some of your ideas. Furthermore, as seismic profiles are not geologic profiles, subsurface
maps can never perfectly represent the structural configuration within the earth. We lack perfect
and complete velocity functions over our data sets, and we lack the high-frequency wavelengths
required to resolve geologic subtleties. The more your ideas are actually tested, the more obvi-
ous it becomes that interpretation and mapping are both an art and a science. Ideally, you will
asymptotically approach the truth as more data and better interpretation techniques become avail-
able. The measure of an interpreter is his or her ability to approach the truth quickly with the lim-
ited data available.
Data Validation and Interpretation 139
The first incorrect assumption is that the reflections represent discrete layers of rock. A
reflection seen on a section may or may not represent a discrete sedimentary boundary. The ver-
tical complexity of the sedimentary sequence and the frequency content of the recorded and
processed seismic signal determine the appearance of the seismic wiggles. Figure 5-5 is a syn-
thetic seismogram illustrating the relative “size” of seismic wiggles in an average velocity
Tertiary section, such as that in the Gulf of Mexico, in comparison to a well log curve. It is obvi-
ous that the vertical resolution of a well log is vastly superior to that of a seismic trace. The seis-
mic wiggle trace is a composite of waveforms from reflections from many boundaries in the sub-
surface. Figure 5-6 illustrates how a series of interfaces can combine or convolve their reflections
to produce a simple seismic reflection.
At this point, you may be overwhelmed by the potential complexity of the seismic wave-
form. Let us say that in most cases, it is safe to assume that the individual reflections represent
mappable, isochronous, sedimentary unit boundaries or sequence boundaries (Mitchum and Vail,
in Payton 1977; Wilgus et al. 1988). This assumption usually will not sacrifice the integrity of the
final map in the least. In areas where there is no radical thinning or thickening of the sedimenta-
ry section, it is reasonable to assume that the reflections, at the very least, parallel the sedimen-
tary units.
Figure 5-5 Synthetic seismogram illustrating lack of vertical resolution in seismic data.
Data Validation and Interpretation 141
Figure 5-6 Composition of seismic trace from velocity and density contrasts. (After Anstey 1982.
Published by permission of the International Human Resources Development Corporation Press.)
The exception is noted in the situation where you may be forced to map a horizon that caus-
es no seismic event but is located in an interval between two diverging reflectors. In some cases,
the most likely position of the horizon is not parallel to either reflector. Mapping a nonevent is
often referred to as phantoming (Sheriff 1973).
Keep in mind that the vertical or horizontal resolution of seismic data will never be as good
as that of well log data, but experience has shown that the seismic reflections typically represent
isochronous geologic surfaces (Payton 1977). This fact makes it possible to map 2D seismic data
between wells.
A second incorrect assumption, which is shown in Fig. 5-4, is that a curved fault trace seen
on a seismic section represents a listric fault. Indeed, the fault shown in the figure is slightly
listric, but the reason for stating this is not because of the curved expression of the fault trace on
the seismic section; rather, it is the presence of the rollover seen on the seismic time section. A
perfectly linear feature in the subsurface may look curved when plotted on a seismic time section
– a situation often encountered when plotting direction well paths on seismic sections. You can-
not rely on the linearity or nonlinearity of a feature on a seismic time section to be a reliable indi-
cator of its actual geometry in the subsurface without first converting the feature from time to
depth and displaying the section with equivalent vertical and horizontal scales.
The most insidious assumption you can make regarding a seismic section is that the section
is really just a geologic cross section of the earth directly under the line. You must always keep
in mind that a seismic section is displayed in two very different dimensions: space and time, and
not in the geologic realm of space and depth.
A time section is simply a series of traces displayed next to one another on a piece of paper
or on a computer screen (ignoring variable density displays). The distance along the line is a
physical distance and represents a distance along the surface of the earth. Therefore, looking hor-
izontally along a section requires only that you understand the scale. If you are not working on a
workstation, a typical full-scale paper section might be 5 in. to the mile along the top of the sec-
tion. Figure 5-7 shows a typical time section with annotation illustrating the accepted working
terminology for the various parts of its display. On the workstation, the horizontal scale is typi-
cally posted along the top or base of the section, or it can be obtained in one of the windows.
142 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
It is the vertical dimension on seismic time sections that can lead you astray. It sometimes
seems reasonable to assume that the vertical dimension also translates directly into a scalable
physical distance. The vertical section is displayed in two-way time. It represents the amount of
time it takes for a seismic signal or wave front to travel from the surface, down through the earth
to the reflector, and back to the surface. It would be simple if the seismic velocity field remained
constant throughout the earth, but this is not the case.
Even in “geologically simple” areas, velocity changes with depth. In general, the deeper the
rock, the higher its velocity. Figure 5-8 is a time-depth table from checkshot data taken in a well
in Pliocene sediments. (A checkshot measures the actual time for a surface seismic source to trav-
el to a receiver lowered down a wellbore. This one-way time is converted to two-way time by
doubling the times for any given depth.) Underlined in the figure are the subsea depths at 1 sec,
2 sec, and 3 sec. The depth represented by 1 sec of two-way time is 3227 ft. The depth at 2 and
3 sec is 6996 and 11,642 ft, respectively. So in this example, depending on the depth, an incre-
mental 1 sec of two-way time may represent 3227, 3769, or 4646 ft. Therefore, before you make
conclusions concerning the listric shape of faults from a time section, you must convert the two-
way times to depth and display the depth section at true scale.
One expensive way to convert the two-way times is to have all the seismic section’s depth
converted. This may not be necessary, but this depth conversion process is routinely conducted
during 3D-depth migration (Chapter 9). An easier and less costly method is to convert all time
points to depth, using a valid checkshot, before constructing a geometric interpretation.
To demonstrate how different the perspective can become after converting everything to the
same dimension, look at Fig. 5-9, which shows a depth-converted fault trace plotted at the same
horizontal scale as the seismic profile in Fig. 5-4. Does the fault look as curved as the trace on
the time section? This example illustrates the effect time sections can have in distorting the true
geometry of a geologic feature.
Furthermore, when considering the listric nature of the fault, you cannot assume that this
seismic section orientation is perpendicular to the strike of the fault surface, since fault strike
cannot be determined on the basis of one line. Essentially, the only statement you can make is
that the trace of the fault on this section appears to be concave upward in time. Figure 5-10 shows
a hypothetical fault surface and seismic line showing its fault trace on the section. Observe that
the fault surface is curved, but the dip of the surface itself maintains a fairly constant angle. It is
enlightening to note that because of the orientation of the line with respect to the fault surface,
the trace of the fault, even on the depth section, appears to represent a listric fault, when in fact
this is not the case. The easiest mistake to make in seismic interpretation is to infer 3D geology
from observations based on a single seismic section, either from a 2D or 3D data set.
80
20
60
00
40
82
82
83
83
84
84
2,000'
4,000'
6,000'
8,000'
10,000'
12,000'
14,000'
16,000'
mon surface (i.e., geologic log cross sections or seismic sections) will show the intersection of
that surface at the same elevation on both profiles. This is illustrated in Fig. 5-11.
Even though this seems self-evident, the most common error in seismic interpretation is fail-
ure to ensure that all geologic surfaces that affect an interpretation have been tied around a loop
along the lines. This includes tying the faults from line to line. For example, our experience with
2D and 3D data sets demonstrates that the failure to loop-tie fault surfaces can result in mapping
two faults as one. The failure to loop-tie fault surfaces on 3D data can result in so-called trapping
faults that do not exist. This problem is most important where, in the strike direction, one fault
replaces another. This area is called the fault ramp or bridge (Chapter 11). The only cases where
tying faults is difficult are in areas where the fault surfaces are near vertical, or in areas of com-
plex deformation where the strike lines are poorly imaged. This may seem laborious (and often
is), but the ability to tie surfaces by following a laterally continuous seismic event is one of the
major advantages that seismic data has over well data. Well data forces the interpreter to infer a
continuous surface from point information, whereas seismic data shows explicit continuity for the
horizons and faults being mapped. By tying surfaces, you can eliminate some of the ambiguity
that may arise when just using point information from well data. In effect, the act of tying both
horizons and faults on a network of lines continually extends the surface and eliminates a num-
ber of possible surface configurations that may arise from the point data in wells.
Figure 5-12 is a set of diagrams illustrating the utility of tying surfaces in order to eliminate
this three-dimensional ambiguity. Figure 5-12a is a fault surface map that was constructed sole-
ly from fault depth data derived from well log correlation. Figure 5-12b shows a different fault
surface configuration that satisfies the same set of well log fault data. It is apparent that in some
cases the point data from the wells are insufficient to uniquely define a geologic surface. An inter-
pretation that satisfies the data may not be a unique solution. Figure 5-12c shows the same area
with a set of two seismic lines, and Fig. 5-12d is a representation of these two lines, tied at their
intersection and interpreted. To satisfy the requirement that all nonvertical surfaces should tie, it
is easy to see that all but one of the possible fault surfaces cannot be justified with the grid of
146 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
seismic data. Figure 5-12e shows a completed fault surface map, which is different than the other
two but more accurate because both well data and seismic data are integrated into the fault sur-
face interpretation.
Tying seismic data serves two important purposes. First, it establishes a relationship between
the traces of surfaces seen on seismic profiles. In other words, by tying the data, we can assure
ourselves that a given interpretation of a geologic surface on one line is indeed the same surface
as interpreted on an intersecting line. This principle applies to both 2D and 3D data sets. The sec-
ond benefit of tying seismic data is the ability to project the horizon being mapped into areas
where well control may not exist. This forms the basis for many wildcat prospects. As previous-
ly mentioned, few wildcat prospects have wells near them. Seismic data allow you to extend a
mapped horizon into areas with little subsurface control.
Figure 5-11 Intersection of seismic sections through a subsurface structure. Events must tie at
line intersections because they represent the same subsurface point.
Data Validation and Interpretation 147
(a) (b)
(c) (d)
(e)
Figure 5-12 (a) Possible interpretation of hypothetical fault data observed in a set of wells. (b) Another possible inter-
pretation of the same fault data. Multiple fault surface interpretations are possible from this set of data. (c) Hypothetical
seismic grid through faulted area. (d) Appearance of hypothetical seismic sections when interpreted and tied. Note that
fault traces on profiles meet (tie) at the intersection of the two lines. (e) Final fault surface interpretation, integrating both
seismic and well data.
148 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
Be aware of the tendency, however, to believe that a given interpretation must be correct
because it ties between lines. Any nonvertical surface being interpreted on a section can be drawn
to tie between any set of lines. Of course, when it is observed that seismic profiles show inter-
preted horizons crossing the actual seismic events, there is good cause not to believe in the inter-
pretation. An exception occurs in reverse-faulted and thrust-faulted terranes where horizons may
overlap because of faulting. More subtle problems can creep into the interpretation during the
process of tying faults. Fault traces on seismic sections are sometimes difficult to see, and tying
an array of closely spaced faults can present a near-impossible task, even on 3D data sets. An
insufficiently spaced grid of seismic data in an area of dense faulting can present problems of
aliasing, particularly if there is little data oriented along the strike of the faults. Figure 5-13 illus-
trates how aliasing can be a problem in highly faulted areas. Line Y is a strike line on which the
faults are poorly imaged.
The geometry of the faults in Fig. 5-13 may help with the tying problem. Notice that the
maximum upper limits of two of the faults on line Z is the same as the maximum upper limits of
the two faults on line X. It would be a reasonable first guess that the faults that die or are buried
at the same elevation are the same pair of faults. This supposition, however, should be consistent
with a believable tie and the construction of a reasonable set of fault surface maps.
Perhaps the most useful advice about tying data is to always think in terms of geologic sur-
faces and to regard what is seen on seismic sections as merely the trace of this surface intersect-
ing the surface of the seismic line. Remember, however, that the time sections are not depth sec-
tions. Making geologic interpretations based on one seismic profile defies the 3D nature of the
earth. Tying the data is a visualization tool that helps create a 3D representation of the subsur-
face.
Figure 5-13 Ambiguity when tying closely spaced faults. Line Y is a strike line. Multiple interpretations may
be possible.
Data Validation and Interpretation 149
Picking a Reflection to Interpret and Map. Seismic profiles (e.g., Fig. 5-4) contain numer-
ous reflections, and it is obvious that it would not be possible or practical to interpret and map
every event. The interpreter should look critically at the sections and decide which seismic hori-
zons are best to interpret and to map. Typically, the chosen events correspond to selected strati-
graphic horizons, although reflection strength or continuity can also influence the decision.
150 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
Today, interpreters typically map sequence boundaries because not only are these boundaries
commonly the most laterally continuous events on a section, but they are also directly related to
rock type and the geologic history of an area. Sequence boundaries are geologic unconformities,
or surfaces of erosion or nondeposition, and represent approximate isochronous surfaces (Payton
1977). They can generally be located on seismic sections by observing where reflections con-
verge or are truncated against a (usually) strong event.
Sequence stratigraphic analysis is an exciting use of geology in geophysical interpretation.
Covering this subject in any detail is a book in itself, and therefore we direct you to the references
in the bibliography (in particular, Payton 1977; Berg and Woolverton 1985; Wilgus et al. 1988).
You should be aware of the implications of this growing body of knowledge, as it has had and
will continue to have a major impact on petroleum exploration.
Annotating the Well Information. Now that you are confident that your geologic interpreta-
tion is underway, mark the position of any wells that intersect the seismic lines. Straight holes
require only that you have reasonably correct checkshot data close to the line. Directional wells
require that a directional survey be available. Directional surveys and the projection of deviated
wells into a seismic line are covered in Chapters 3 and 6. Remember that you must convert the
depth points to their equivalent two-way travel times in order to annotate them correctly on a seis-
mic section.
It is important to remember that any projection is a compromise and will commonly cause
some confusion. Figure 5-14a is an illustration of a projection of a directional well onto a seis-
mic section. Notice on Fig. 5-14a and b that an orthogonal projection of the well into the seismic
line suggests that the well penetrated the footwall of the interpreted fault. However, the fault sur-
face map (Fig. 5-14b) clearly shows that the well never crossed the fault from hanging wall to
footwall; it is entirely within the downthrown block. This illusion occurs because of the compro-
mises inherent in projecting a 3D entity onto a 2D profile. The routine orthogonal projection of
a well into a seismic line can also cause significant mis-ties of horizons and faults.
Tying Well Data to Seismic with Checkshot Information. Once the well position is anno-
tated, the information from the well data, in the form of geologic tops, must be located and
marked on the time sections or loaded into the computer and annotated on the profile. How do
you know where to find the event that corresponds to the geologic horizons? There are basically
two methods used to tie the geologic control into the seismic data: (1) using a time–depth func-
tion calculated from checkshot data, or (2) tying into the seismic data with a synthetic seismo-
gram.
The simplest but least accurate method of tying well data to seismic is to use the checkshot
data to convert the tops from the log data from depth to time, and post the equivalent horizons on
the seismic section at the proper times. The problem with this method is that you never know
what kind of assumptions may have been made in the processing of the seismic line to correct to
the proper datum. This is why data from different contractors may have static shifts between lines
when tied together. “Ground truth” is hard to ascertain in these circumstances.
There is also a temptation to place unwarranted faith in the checkshot data and to believe it
over all other information. We have seen cases where an interpreter tied a sand occurring in the
middle of a 1000-ft shale interval to a level within a no-data zone and above a distinct event, sim-
ply because a nearby checkshot indicated such a tie. Particularly when sonic and density infor-
mation indicate that a sand should generate a strong seismic response, it is likely that the sand
ties to the strong event in the middle of the mostly reflection-free shale zone. This is not to say
Data Validation and Interpretation 151
(a)
Figure 5-14a Seismic section with directional well projected onto section (straight-line perpendicular
projection used). (Seismic line published by permission of TGS/GECO.)
152 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
(b)
Figure 5-14b Fault surface map of fault observed in Fig. 5-14a. Projection onto line
makes well appear to have penetrated the upthrown block. Fault surface and the well’s
directional survey clearly indicate that the well remained in the downthrown block. Care
must be exercised in projecting wells onto seismic lines.
that shale intervals cannot generate strong seismic events. However, if a sand is present and fair-
ly close to tying an event on the line, then correlate the sand to the reflection and don’t be mis-
led by the checkshot listing.
Tying Well Data to Seismic with Synthetic Seismograms. Tying the well data into the
seismic data with a synthetic seismogram is the preferred method, as it will usually provide rea-
sonable results. Its usefulness, however, depends almost exclusively on the availability of good
quality sonic and bulk density log data from the wells. In some areas, these logs are run as a mat-
ter of course, whereas in other regions they are the exception rather than the rule. Particularly in
older basins, there may be a shortage of high quality sonic and density log data. It is imperative
that the geoscientist determines the quality of the data used to make the synthetic seismogram.
We have seen synthetics made from sonic logs obtained from washed-out wellbores that have
recorded mostly mud arrivals and cycle skipping. Needless to say, the synthetic seismograms
from such logs are useless.
Figure 5-15 shows a high-quality synthetic seismogram and its tie to a seismic profile. This
seismogram is shown adjacent to a seismic section through the actual well location. As you can
Data Validation and Interpretation 153
see, the match is good, though not perfect. The procedure for tying the proper event is to locate
the chosen horizon on the log plotted next to the synthetic, then draw a horizontal line over to the
synthetic trace. Lay the synthetic seismogram over the seismic profile at the appropriate location,
then shift the synthetic up or down to determine if it matches the seismic data. The horizon line
drawn earlier will show where the actual seismic event, corresponding to the log horizon, is locat-
ed on the seismic line.
Always be wary of forcing yourself to see correlations between the seismic data and the syn-
thetic seismogram. If there are problems with either the seismic data or the synthetic data, it may
be impossible to make a valid correlation. As a rule of thumb, a shift of the synthetic by more
than about one hundred milliseconds should be highly suspect. Also, if you can turn the synthet-
ic upside down and get equally good correlations, you should be suspicious of the validity of this
method for tying well horizons into seismic data.
If applicable, this method of tying in the well data is preferable over tying into a particular
seismic event with just checkshot data. If a correlation exists between the synthetic and the seis-
mic profiles, then the synthetic seismogram method will ensure, with a reasonable degree of cer-
tainty, that the event being mapped is the intended geologic horizon. This is particularly valuable
in areas of abrupt stratigraphic thinning and thickening. Figure 5-16 illustrates how an incorrect
pick for the mapping horizon can have a profound effect on the depth of the horizon away from
the well control. As shown, a small error in the thinner stratigraphic section will cause a much
larger error to occur where the stratigraphic section is thicker.
A vertical seismic profile (usually abbreviated VSP) derived from a synthetic trace is also an
excellent tool for tying into the seismic data. The methods for using the VSP to tie into the seis-
mic line are the same as those used for synthetic seismograms. Establish a correlation between
154 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
the traces on the VSP and the seismic, and use the plotted log as a guide for tying into the seis-
mic section. An added benefit to both the synthetic seismogram and the vertical seismic profile
is the ability to use these data to analyze the relationship between the lithology in the well and
the seismic character.
The subject of tying well information into seismic data is deceptively simple. There is a dan-
gerous tendency to believe that your first tie from a well is the correct one. Make sure that it is,
because being tied into the wrong horizon can cause you to miss the relationship between a given
log response and its correlative seismic response. The only other advice we can give is to avoid
a “railroad track” mentality (looking strictly at a narrow 80-millisecond strip of the seismic data),
or you may totally miss what is present above and below the horizon you are mapping.
Tying the Faults. We strongly recommend that fault surfaces be loop-tied prior to loop-tying
horizons. Geophysicists loop-tie surfaces for the same reasons that geologists loop-tie strati-
graphic markers. That is to make sure that you are on the same surface where you began, and not
on a different surface. Mapping faults first has distinct advantages. First, in areas of complex
structure the style of faulting helps determine the type of structures that exist in the area (Chapter
10). How can you determine how the structure formed if you do not understand how the faults
behave in three dimensions? Second, problems may arise when mapping horizons on portions of
a profile that contain a lower seismic frequency or where the data are not perfectly coherent. If
the position of recognized faults has not been identified in three dimensions, then geoscientists
Figure 5-16 Increase in magnitude of error in areas with rapid changes in interval thick-
nesses. Any uncertainty in tying formation top to seismic event can introduce a greater
uncertainty where stratigraphic section thickens away from well control.
Data Validation and Interpretation 155
may map horizons right through fault surfaces. This will result in a horizon mis-tie, as described
in the next section. Our examination of the horizon mis-tie problem on many data sets indicates
that the geoscientist may not even suspect that an existing fault produced the mis-tie. After all,
the assumption was made that the faulting is understood or that fault surface mapping is not
required. The result is that the geoscientist may create a nonexistent fault to solve the mis-tie
problem. The nonexistent fault constructed through semicoherent data or data that abruptly
changes reflection character creates additional mis-ties and more nonexistent faults and mis-
picked horizons. We have seen cases where prospects have been generated as a result of the
nonexistent faults. Needless to say, the wells were dry.
There is another common problem with the picking of nonexistent faults in semicoherent
data or data that rapidly changes reflection character. These nonexistent faults may actually be
axial surfaces or changes in bed dip that are common to compressional and salt-related folds
(Chapter 10). Where the geoscientist encounters a nonexistent fault, the horizons must be offset,
which results in a horizon mis-pick. A mis-picked horizon results in mis-ties and in other
nonexistent faults that are created to solve the mis-tie problem.
However, if fault surface maps are constructed for each of these nonexistent faults, the fault
may not map as a smooth curved surface, but will contain offsets and kinks. The presence of an
offset or kink typically implies more than one fault. Also, the overall geometry of a mapped fault
surface may be simply unreasonable. These types of fault surfaces are not viable geologic sur-
faces and should be rejected. Faults that map as smooth surfaces are considered more plausible.
Furthermore, with 2D data sets, questions typically arise as to which faults link to form a con-
tinuous fault surface. Fault surface maps can help resolve the fault correlation problem.
We find that mapping fault surfaces not only results in better interpretations and in higher
quality prospects, but the process saves time. The time taken to construct quality fault surface
maps is justified when you consider the time required to attempt to solve existent and nonexist-
ent mis-tie problems, the reworking of an interpretation that proves incorrect, and the costs of an
unnecessary dry hole.
Tying the Lines and Horizons. If you have followed the procedures so far, you now have a
tentative interpretation and well data annotated on each profile. To this point, we have not
described loop-tying the data. That is the next step. On a good day, portions of your preliminary
interpretation will probably be wrong when you finish loop-tying the data. Again, methodically
tying the loops will improve the chances of finishing the task correctly and in a timely manner.
The use of a workstation has advantages in that any errors encountered during the interpretation
can be readily corrected and alternative ideas can be easily tested.
We find it much easier to interpret the most obvious geologic features first and to tie them
together on all the lines. After the large features tie, begin another iteration of tying through the
data volume, concentrating on the smaller “second order” features. As the size of the features
being tied together decreases, the number of lines required to tie them also decreases, so the work
goes more quickly toward the end of the process.
It is important to pick a loop-tying scheme that will allow you to make the smallest number
of assumptions while carrying your surface around the map area. Tying a path that crosses the
fewest faults and that crosses faults at their location of smallest displacement will more likely be
correct.
The initial task for tying the loops is to post all the intersections of the seismic data on all
the lines. With paper sections, depending on the number of lines that are being tied, this process
can take anywhere from a morning to several weeks. Figure 5-17 shows a seismic base map with
156 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
two seismic lines. The corresponding profiles are shown in Figs. 5-18 and 5-19. Line A intersects
line B at a location just north of shotpoint 480 on line B.
Line intersections are seldom cooperative enough to fall on a downline (a downline, shown
in Fig. 5-7, merely refers to the dark vertical line printed on the time section at the shotpoint loca-
tions annotated on the maps). Depending on the precision required, you can either make a rough
estimate of the intersection or use a scale to determine that the intersection is exactly 150 ft, or
1.829 traces to the right of shotpoint 480 on line B. All the intersections that you intend to tie
together must be marked.
The next step, when using paper sections, is to fold one of the lines at the marked intersec-
tion. This fold must be vertical. Use a straightedge to ensure that the section is folded vertically.
Align the folded section with the unfolded section at the appropriate intersection. Figure 5-20
demonstrates a tie between the two paper sections. On the workstation, the horizon intersection
picks are posted automatically.
The first thing you will probably notice is that the lines may not tie perfectly when the inter-
secting lines are aligned at a common two-way time (Fig. 5-20). Sometimes they will match per-
fectly, but more likely, they will not match at a common time. At least one section will have to
be slightly shifted vertically to establish a good correlation between the lines. Figure 5-21 shows
how the two lines have to be shifted relative to one another to “tie” the events. Which line has
Figure 5-18 Seismic line A. Profile shows gently dipping reflectors cut by a small growth
fault. (Seismic data published by permission of TGS/GECO.)
158 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
Figure 5-19 Seismic line B. Profile shows gently dipping reflectors cut by a small growth
fault. (Seismic data published by permission of TGS/GECO.)
Data Validation and Interpretation 159
Figure 5-20 Seismic lines A and B intersected with each other. Notice line A appears
to be too “shallow” in relation to line B when the timing lines are aligned. (Seismic data
published by permission of TGS/GECO.)
160 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
been shifted? Has line A been shifted down 10 milliseconds (+ 10 ms), or has line B been shift-
ed up 10 milliseconds (– 10 ms)? You now decide which line shows the “real” two-way time to
the event.
In an interpretation and mapping project of any complexity, it is necessary to pick a refer-
ence seismic line so that all the other data may be posted with the appropriate time shifts relative
to the reference line. In effect, the procedure for choosing a proper reference line is similar to that
for picking a correlation type log (Chapter 4). When using wells, you are picking the log that best
demonstrates the geologic section of the mapping area. When choosing a reference line, pick a
line that has as many of the following characteristics as possible.
1. The reference line should cross as many of the other lines as possible. This is required to
reduce the number of indirect calculations of static shifts.
2. The reference line should be a dip line or as close to it as possible.
3. The reference line should be of high quality relative to the rest of your data. In short, it
must be one of the most believable lines in your collection.
Mis-ties
There are two kinds of mis-ties that must be corrected before posting data on a base map. There
are both static and migration mis-ties. The static mis-ties are reflection-time invariant corrections
made to the event times, and they are the easiest to recognize and correct. The migration mis-ties
are corrections that vary with the two-way time of the events being mapped, and they are more
difficult to correct. Real problems occur when both static and migration mis-ties are present in a
data set. The static component must be recognized and corrected first, and the migration correc-
tion is made after the static solution is determined.
Static Mis-ties. Static mis-ties can be recognized because they cause “bulk” shifts of the inter-
secting line, either up or down to achieve a good correlation (Sheriff 1973). They commonly
occur between data sets of varying vintages and contractors because of different datum correc-
tions and assumptions. Figure 5-20 shows a static mis-tie between two intersecting lines. The eas-
iest way to determine a static mis-tie problem is to search for any shift between flat-lying events
that are normally present in the shallow part of most basins. In Fig. 5-20, line A ties line B per-
fectly with a static shift. We do not know, however, if the times on line A are too large or if the
times on line B are too small. There is no absolute answer. This is the reason for picking a refer-
ence line; it establishes a reference or datum. The rest of the lines can then have the time picks
for a given event adjusted to the datum established by the reference line. If a line does not direct-
ly intersect the reference line, then its relative mis-tie with a line that does intersect the reference
line is added together with the adjustment value from the line intersecting the reference line. This
is harder to describe than to illustrate, so Fig. 5-21 shows an example for keeping track of static
mis-ties on lines that do not directly intersect the reference line.
Once a reference line is chosen, annotate the rest of the sections with their respective static
mis-tie values relative to the chosen datum. An important point to note about this process is that
it should be carried out only where you are tying events that are relatively low-dip (probably less
than 8-10 deg at most). High dip rates cause an effect called migration mis-tie, which is discussed
next.
Migration Mis-ties. A migration mis-tie is one of the more difficult aspects of interpretation of
2D seismic data, and it becomes problematic in areas of high bed dips. As noted previously, you
Data Validation and Interpretation
Figure 5-21 Keeping track of relative mis-tie through a grid of seismic data. All
161
absolute mis-ties are in reference to reference line.
162 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
may encounter varying amounts of mis-tie with different time ranges of events. In general, the
shallow, low-dip events will tie reasonably well with a static shift. But as the dip increases with
increasing depth (greater time on the section), you may notice that events on the dip lines appear
to be too deep (too large a time) relative to the intersecting strike line. Figure 5-22 is a sketch of
this phenomenon. This mis-tie problem is present because of the limits of 2D seismic data in
imaging a 3D surface.
To understand this problem, it is first important that you understand what the migration
process does to seismic data. Figure 5-23 is a simplified illustration of a 2D seismic line shot in
the dip direction. Two simple, normal incidence raypaths are drawn on the section from surface
positions A and B. By definition, a normal incidence ray will intersect the reflector Z at a right
angle. Raypaths drawn to satisfy this condition are A-A′ and B-B′. Assume the two-way travel
time for a reflection from point A′ is 2.0 sec and the two-way travel time for a reflection from
point B′ is 2.1 sec. Both these reflection points are being recorded at the surface at positions A
and B, respectively. So, on an unmigrated seismic section, the events appear to have the positions
shown by dashed line Z′. The seismic lines are recording data at surface locations A and B from
subsurface reflection points A′ and B′, which are located up-dip of surface locations A and B.
Figure 5-22 Appearance of migrated strike and dip lines at their intersection.
Notice events on strike line are too “high” (too small a time).
Data Validation and Interpretation 163
Figure 5-24 Strike line imaging. Reflection path for strike line is not underneath the
line. Data is being recorded from up-dip.
164 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
Migration is a fairly complex process that corrects the data by moving it back to its proper
time position relative to the surface location. In other words, after migration, a reflection point
should be positioned correctly with respect to the surface recording points. Migrating the exam-
ple would move the event Z′ to a position coincident with the actual event Z. Migration will
always steepen events or reflectors and cause a given event to appear “deeper” (occur at a larg-
er time) for any given shotpoint when compared to the unmigrated data. Migration of the seismic
time data is critically important in obtaining a reasonably accurate interpretation of the subsur-
face.
The problem with migration is that it can fully correct only a true dip line. Because the 2D
migration algorithm cannot move data from out of the plane of the line, a line that has an appar-
ent dip and not true dip will not be fully corrected. Figure 5-24 shows the extreme case: a true
strike line that is oriented perpendicular to dip direction and thus contains no component of dip
whatsoever. The data are still being recorded from up-dip, but the migration algorithm cannot
move data out of the 2D plane of the line. The apparent dip is zero, so the migration algorithm
really has no effect on the line. Event Z is positioned on the line at the two-way time of A-A′ and
B-B′, much shallower (smaller two-way time) than is really the case beneath the surface line A-
B. Figure 5-22 shows what occurs at the intersection of a true dip and true strike line. All the
events on the strike line appear to come in too shallow (too small a time). This is important to
remember when tying data. Only in very peculiar circumstances will a strike line event intersect
a dip line deeper (greater time) than the corresponding event on the dip line. (This assumes that
any static mis-tie has already been taken into account.)
Figure 5-25 illustrates what effect a migration mis-tie can have on a line intersection in areas
where there is increasing dip with depth: The mis-tie becomes larger and larger in the deeper
(greater time) part of the section. In an interesting twist to this problem, notice that the strong
event at about 2.6 sec on the strike line is actually “deeper” than its correlative event on the dip
line. How can this be possible if the data are coming from up-dip? Notice that all the events are
too shallow on the strike line until about 2.4 sec, when they begin to appear deeper. By tracing
horizontally (isotime) from the strike line to the dip line, we can see that the events are deeper
because they are being recorded from up-dip, which happens to be downthrown to a buried
growth fault, whereas the section ties the dip line in the upthrown block. Notice also that the fault
intersection point on the strike line is actually 300 ms deeper than a “mechanical” tie would indi-
cate. In growth fault areas, this problem is not uncommon. Many growth faults have larger dips
in the nongrowth sediments located beneath the fault at depth.
The problem can be handled in two ways. The first and easiest method is to use the strike
lines only to tie the events among the dip lines, but ignore them as valid sources of data points
for drawing maps. In areas where there is already abundant well control and adequate density of
dip-line coverage, this may be a viable option. The strike lines can still be used to tie events
among the lines, but their actual time values are ignored. The disadvantage of this method is that
the strike lines contain information on cross structures that is often critical to development of a
potential hydrocarbon trap (see Chapter 11).
The best option (and the most time consuming) is to explicitly correct the strike line data by
moving the data to their proper position relative to the surface locations. There is an easy graph-
ical way to accomplish this task that the 2D seismic workstation does not automatically provide.
Figure 5-26a shows hypothetical dip and strike lines posted on a structure map. Figure 5-26b
shows an intersection of migrated lines A and B, as they would appear on migrated time sections.
If you assume that the dip line A is properly migrated, then the data on strike line B at the inter-
section is actually coming from a position that is up-dip of the intersection with the dip line.
Data Validation and Interpretation 165
Figure 5-25 An unusual case of migration mis-tie. Strike section events occur earlier
(“shallower”) than same events on dip section until approximately 2.4 sec. At the bold
event at 2.6 sec, strike line continues to record up-dip; however, this is also downthrown
to an expanding growth fault. (Seismic lines published by permission of TGS/GECO.)
166 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
This concept is easier to illustrate than to explain, so Fig. 5-26b shows an imaginary hori-
zontal line drawn from the event intersection on the strike line to its real reflection point on the
dip line. The process for correcting a series of lines is to find the actual reflection points for all
the intersections of the strike lines with dip lines, and to mark these points on the map. These
points locate positions near where the reflected event actually originated (Point A on Fig. 5-26b).
If this procedure is carried out over a series of intersections, then a “corrected” base map can be
made that shows the approximate reflection path of the strike line. The dashed shotpoint display
in Fig. 5-26a illustrates how to “relocate” the strike line up-dip before posting the data points to
use in making a map. Once this corrected base map is made, interpolate the individual shotpoints
and post the times from the strike lines at their approximate subsurface locations. One important
point about this technique: Any correction is only valid for the particular event being interpreted
and loop-tied. For example, another event that is deeper may require another corrected base map
to be constructed for its structure map. In other words, the correction is not a fixed value, but
rather, varies depending on the depth of the event being mapped.
This discussion has been directed toward tying actual seismic events among lines. Faults
must also be tied, and in areas with high bed dips, the migration mis-tie problem can make map-
ping fault surfaces extremely difficult. It is vital to remember that if the fault surface reflections
are coming from somewhere other than beneath the line, then the fault surface reflection is sub-
ject to migration effects. What makes mapping faults especially difficult is the spatial aspect of
the migration mis-tie problem. It is entirely possible to have data points for a given fault surface
coming from further and further away from the line, simply because the profile crosses the fault
at an oblique angle. This causes the fault to change depth relative to the profile. Figure 5-27a
shows two seismic lines that intersect each other. Line A is a dip line and shows increasing dip
of the seismic events with increasing time (depth). Line B is a strike line with an obvious fault
cutting the events on the line. It is obvious from the tie that the events on line B are actually com-
ing from up-dip of the surface location of the line. The discontinuity in the events is caused by
fault X, and thus the image of fault X is also coming from up-dip of the surface location of the
line. Point “X” on Line A in Fig. 5-27a is the actual tie point for fault X. Figure 5-27b is an illus-
tration of the method used to construct a corrected fault surface map by moving the times for the
fault trace further and further up-dip with increasing depth of the fault. The complex nature of the
migration process illustrates the advantages of 3D migration.
DATA EXTRACTION
Picking and Posting
After the seismic lines have been interpreted, transfer all the information to a base map and begin
the process of making a subsurface map. As pointed out earlier, the seismic data should be post-
ed along with all the subsurface information from electric well logs. The mapping process is cov-
ered extensively in other chapters, and having seismic data on the map along with the subsurface
well data should not affect the techniques used for the actual mapping. The following discussion
pertains to posting data from interpreted sections. Most of these data are automatically recorded
during interpretation on a workstation.
Types of Data from Seismic. When interpreting by hand, the most obvious type of data to
post from seismic sections are the actual two-way travel times for the events that correspond to
the geologic horizons being mapped. This is analogous to posting formation tops on the map
when using well data. The same two-way travel times can be posted for any fault surfaces being
Data Extraction 167
(a)
(b)
Figure 5-26 (a) Approximate method for dealing with migration mis-tie involves relocating
strike line data points up-dip as determined from a true dip line. (b) Method for calculating
the distance that strike data must be moved to account for migration mis-tie. Find intersec-
tion of time observed on strike line (at line intersection) with the same event on dip line.
168 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
(a)
(b)
Figure 5-27 (a) Hypothetical dip and strike lines with fault observed on strike line. Notice the effect that
an increasing amount of dip, with increasing depth on the dip line, has on the correlative events on the
strike line. (b) Relocation of fault surface points to proper spatial location, using method outlined in text.
Notice that the change in the distance the points must be shifted actually changes as the record time
(depth) increases.
Data Extraction 169
mapped. There are several other types of mapping information that can be extracted from the seis-
mic data. One type of information that is extremely useful is the upthrown and downthrown inter-
sections (cutout points) of the horizon being mapped with the surface of a fault. These intersec-
tion points have both a vertical datum and a location associated with them. The significance of
these points, assuming the data is reasonably high quality, is that they can be used to help posi-
tion the upthrown and downthrown traces of the fault and the approximate width of the fault gap.
In practice, seismic methods overstate gap width by about a factor of 2, and the correct width of
the gap is finally derived from structural horizon and fault surface map integration (Chapter 8).
Many interpreters post a solid bar on the map in an easily identifiable color to indicate the
fault trace identification. In complex areas, you can assign a unique color code to each fault being
tied on the seismic sections and post the trace on a base map in the same color.
Depending on the area, there may be some other useful information that can be posted on
the map. If the seismic event being mapped has an amplitude anomaly associated with a hydro-
carbon-bearing sand, then the areal extent of the amplitude anomaly can be posted. In areas adja-
cent to salt domes, you may be fortunate to have data sufficient to identify a salt/sediment inter-
face. This contact can be posted and mapped. In areas with stratigraphic discontinuity in the
objective sands, you may be able to detect a unique seismic response indicating where the sand
is present and where it is not. The extent of the potential reservoir body can thus be mapped.
Extracting the Data. When tying the data around a loop by hand, use a colored pencil to color
in the troughs of the seismic data. At each downline or shotpoint marked on the line, mark with
a pencil a consistent part of the waveform, using either the maximum trough (which is easy to
see), the maximum peak, or the crossover.
Posting the Information. Now that the data have been interpreted, they have to be transferred
from the seismic lines to the base map. With an engineer’s scale, measure the two-way time in
milliseconds to the event being mapped. Next, add or subtract the constant value in milliseconds
that represents the amount of static mis-tie between this seismic line and the reference seismic
line. Post the two-way time on the map at the actual reflection point for the horizon being
mapped. If you constructed “pseudo” base maps with all the strike lines repositioned up-dip, post
two-way times at the “corrected” shotpoint locations. Remember that the adjusted strike line
location may vary, depending on the depth of the seismic event that is being mapped. Mark all
the intersections of the seismic events associated with fault surfaces, and post the upthrown and
downthrown intersection (cutout) points. Finally, record any other valuable information and post
it on the map.
Brute Conversion with a Time–Depth Table. The first method for converting time to depth
170 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
is extremely easy. The procedure for converting the values involves determining the depth values
corresponding to the posted time values in a time–depth table that has been generated from
checkshot data. Checkshot data is acquired during the evaluation phase of drilling a well. A
checkshot measures the amount of time it takes for the first arrival of a seismic wave to travel
from a surface source near the well to a receiver lowered down the wellbore. After the data are
acquired, they are usually interpreted to generate a set of one-way travel times for specific depths
in the well.
Conversion of these one-way times to two-way times involves multiplying the time values
by two. (The receiver only measures the time it takes for a wave to travel to a given depth, where-
as a seismic line measures the time required for a wave to travel to a reflecting horizon and back
to the surface.) The two-way time–depth pairs are then interpolated from the usually sparse set
of data points to generate a table with the time–depth pairs calculated at even increments of time.
Such a table was shown in Fig. 5-8. A plot of a particular checkshot can be shown graphically by
constructing a time versus depth graph. The actual times are shown as data points, and a line is
fitted to the data points using a cubic spline curve-fitting computer program.
Once the correct depth for a given two-way travel time has been found, you can post the
depth value beside the time value on the map. Using a contrasting color pencil to post depth infor-
mation is a good method for keeping the time values distinct from the depth values you will use
to contour the map.
Where should this method be used? Use this simple technique in areas where there does not
appear to be any large lateral variations in the relationship between seismic time and depth. The
method for determining whether this is appropriate is as follows. First, use a time–depth table
generated from the closest well to the mapping area (or a well in the middle of the area) to con-
vert all the times to depth. Second, examine your map, looking for obvious discrepancies between
the well tops posted on your map and the converted time values. If, for example, the depth from
a converted time value is 11,200 ft subsea, and a well top at the same location is 10,700 ft sub-
sea, then there is an obvious problem to be addressed.
Recognizing Velocity Problems. If the velocity information does not tie accurately to the
wells, there are two possible sources for the differences in the values posted. One possibility,
often overlooked, is that the wrong horizon was interpreted on the seismic line, and thus inap-
propriate time values are being converted to depth. This problem could arise from an incorrect
pick across a fault or from an incorrect tie to a well. The error can be caused by using too large
a shift to tie to a synthetic seismogram. Perhaps the synthetic seismogram contains problems
related to log quality or some other factor. There exists a whole multitude of possible causes sim-
ilar to these.
Another possibility is that there is a strong horizontal velocity gradient in the area that
causes the time–depth relationship to change laterally. Lateral variations are often easy to deter-
mine if you have logs and synthetic seismic events that are easy to correlate in two wells. In this
case you absolutely know that both the correlations and time picks on seismic data are correct
and the correlations in the wells are correct. However, the correlations between the wells don’t
agree with one another. In other cases, the problem may not be so easy to identify. If either the
well correlations or the seismic event ties are ambiguous, the presence of a gradient may not be
obvious because of the difficulty in deciding what interpretation to rely on initially.
Lateral velocity changes can be extremely difficult to manage. It is beyond the scope of this
book to attempt a complete discussion of the methods of handling velocity gradients. The key
question is when to recognize the need for expert help. A simple rule of thumb is this: If the gra-
Data Extraction 171
dient in the area being mapped is severe enough to cause the depth uncertainty for a given two-
way time value to exceed the average amount of closure on the features you are mapping, you
should be careful. In particular, if this depth uncertainty can be observed to occur between check
shots in wells that are as physically close as the average dimensions of a prospective closure, the
likelihood of correctly mapping a structural closure is small. If a problem exists, seek the assis-
tance of a senior geophysicist who understands all aspects of handling velocity gradients.
Accounting for Small Velocity Problems. If the magnitude of velocity changes over your
mapped area is not severe, you can often “eyeball correct” a map to account for a gradient. To
correct for gentle gradients, interpreters may make ad hoc adjustments to the map to account for
the horizontal velocity gradients. In practice, interpreters will attempt to find the geologic reason
for the velocity differences and use a different time–depth table on either side of the geological
“boundary” causing the anomaly. For example, in the Gulf Coast tertiary section, a large growth
fault will commonly have downthrown section that has a different velocity field than the
upthrown section. In some areas, the seismic velocities may be faster on the downthrown side
because of the increased amount of sand in the downthrown block. In other cases, the seismic
velocities may be slower in the downthrown block because the thickened section was deposited
rapidly and is undercompacted and slightly overpressured, and therefore the velocities are slow-
er. Whatever the situation, if you can determine the reason for the gradient, you can often adjust
the contours to honor the well control.
We have found that the easiest way to handle velocity gradients that do not appear to have
definite “boundaries” is to use the following technique, illustrated in Figs. 5-28a-d. A base map
with posted information obtained from both well logs and seismic sections is shown in Fig.
5-28a.
First, prepare a pure-time map. Map the isotime contours of the time values, as shown in Fig.
5-28b. Next, determine the average velocity in the depth range being mapped. Simply determine
the number of milliseconds of two-way time that the contour interval, in depth, represents at the
depth range you are mapping. A typical Tertiary value might be about 22 ms per 100 ft. At each
point of well control, use the time map as a guide and begin contouring in depth, using the dis-
tance between each time contour as a rough indicator of the magnitude of bed dip. Carry the con-
tours about halfway to the next well and then start at that well and contour away from that well
until you meet the previous contours (Fig. 5-28c). The discrepancy in the depth values can then
be adjusted by splitting the difference between the two sets of contours and gradually adjusting
the mis-tie in the spacing of the depth contours (Fig. 5-28d).
The points to remember about this technique are, (1) it is quick but imprecise, (2) it is appro-
priate when applied to minor velocity problems over large areas, and (3) it is useful when the map
must be finished quickly. A caveat: Never use this technique when the gradient is severe and is
present over a single structure. It is an appropriate and useful technique when you are mapping
on a large scale and need a method of “absorbing” the mis-ties in the synclines between the major
closures. Remember that seismic data do not have the vertical resolution of well data. A seismic
line sampled at 4 ms and picked to an accuracy of 10 ms will give you about 40 to 60 ft of error
in an average Gulf of Mexico Tertiary section. Using this technique to account for a 200-ft mis-
tie problem seems acceptable.
The most important point to remember about velocity problems is that it is very easy to
exceed your expertise with simplistic solutions. Look critically at your data and get assistance if
it is needed.
172 Chap. 5 / Integration of Geophysical Data in Subsurface Mapping
(a)
(b)
Figure 5-28 (a) Base map showing posted seismic two-way times and formation tops from
well data for a mapping horizon. (b) Isotime map of seismic two-way data. (c) Preliminary
depth map constructed from each point of control. Contouring begins at each point of well
control and moves outward. (d) Final depth map with horizontal velocity gradient “lost” in
the syncline between the two plunging noses. The difference in depth values has been
adjusted in the contouring of the syncline between the well control points.
Data Extraction 173
(c)
(d)
CROSS SECTIONS
INTRODUCTION
A contour map depicts the horizontal plan view configuration of a single attribute of a strati-
graphic unit, such as structure, thickness, and percent porosity. In contrast, a cross section depicts
the configuration of many units as typically viewed in a vertical plane. Since a map or cross sec-
tion alone cannot represent the complete subsurface geologic picture, both must be used to con-
duct a complete and detailed study.
Geologic cross sections constitute a very important geological exploration and exploitation
tool. They are useful in all phases of subsurface geology as well as in reservoir engineering. Cross
sections are used for solving structural and stratigraphic problems in addition to being employed
as finished illustrations for display or presentation. Used in conjunction with maps, they provide
another viewing dimension that is helpful in visualizing a geologic picture in three dimensions.
If a cross section is oriented perpendicular to the strike of the structure, it is termed a dip sec-
tion. If the section is oriented parallel to the strike of the structure, it is called a strike section.
Finally, if the orientation is oblique to the structural axis, it is termed an oblique section.
Various data can be used to construct a vertical cross section. It can be based on surface data
(dips), electric well log data (markers, unit tops and bases, dips, and faults), seismic data, or
entirely from completed subsurface maps. As mentioned in Chapter 1, all available data should
be used in the preparation of a subsurface interpretation, whether it is in the form of a structure
map or cross section.
175
176 Chap. 6 / Cross Sections
177
the New Orleans Geological Society.)
178
Chap. 6 / Cross Sections
Figure 6-2 Structural cross section prepared using well log sticks. (Modified
from Oil and Gas Fields of Southeast Louisiana, v. 3, 1983. Published by per-
mission of the New Orleans Geological Society.)
Structural Cross Sections 179
Figure 6-3 Typical cross section layout for a complex piercement salt structure.
such as that of a faulted diapiric salt structure (Fig. 6-3), the evaluation of the structure, salt, and
fault geometry may require a number of cross sections to be laid out in the direction of structur-
al dip and perpendicular to fault strike. Each structure and the problems to be solved must be
evaluated individually as to the planned direction and the number of sections required to ade-
quately study the geologic feature.
Oil and gas wells do not normally lie in a straight line, as shown in Fig. 6-4a, and so the
direction of a planned cross section may not be in a straight line. Instead, line segments between
adjacent wells may vary in length and direction, giving the line of section a zigzag appearance
(Fig. 6-4b). With a zigzag section, wherever the direction of the line of section changes, the
apparent dip of horizons and faults on the section also changes. An example of such a change can
be seen in Fig. 6-1 for Fault B. Notice that the dip of the fault changes from 46 deg between Wells
No. 3 and 4 to about 18 deg between Wells No. 4 and 5. This change indicates that the line of
section between Wells No. 3 and 4 is perpendicular or nearly perpendicular to the strike of Fault
B; but between Wells No. 4 and 5, the line of section is more parallel to the strike of the fault.
These apparent changes in fault dip are illusions of subsurface geometry caused by the
orientation of the line of section. Such illusions must be considered when laying out any line of
section in order to prevent confusion to an observer and also to make sure that the section shows
what it was intended to show.
180 Chap. 6 / Cross Sections
(a) (b)
Figure 6-4 (a) A cross section oriented through wells that form a straight line. (b) A zigzag line of section formed
by line segments between adjacent wells.
We recommend that structural cross sections be drawn with the same horizontal and verti-
cal scales whenever possible. With the same scales, the cross section is prepared true to scale
with no vertical or horizontal exaggeration. At times, however, exaggeration is required to permit
legible vertical detail. The effects of vertical exaggeration are discussed in detail later in this
chapter.
1. The logs that are chosen for a section must fit the scale of the cross section. The logs may
need to be enlarged or reduced to make the scale equal to that of the cross section.
2. If possible, use the same vertical and horizontal scale. This is usually convenient for field
studies; however, regional or semiregional sections may require exaggerated vertical
scales to get them down to a manageable size.
3. Data must be legible. Log heading data, correlations, depths, etc. may have to be posted
after the section is laid out in order for them to be legible.
How are electric logs mounted on a manually generated cross section? The procedure usu-
ally begins with the preparation of a film positive for each of the original logs to be used in the
Structural Cross Sections 181
Figure 6-5 Completed structural cross section from Bayou Jean La Croix Field, Terrebonne Parish, Louisiana.
(Reprinted from Oil and Gas Fields of Southeast Louisiana, v. 3, 1983. Published by permission of the New
Orleans Geological Society.)
section. If required, the logs are reduced to the appropriate vertical scale to accommodate the sec-
tion. The film positives are normally taped on coordinate grid cross-section paper with the cor-
rect proportional horizontal spacing between wells. It is good practice to plot your line of section
on a well base map or structure map (Fig. 6-13) so the spacing between well logs for the section
can be measured directly from the base map. For a structural section, the logs are normally hung
with sea level as the reference datum. Therefore, the measured log depths must be converted to
subsea depths for vertical position on the section. Finally, an ozalid or xerographic print of the
cross-section base is made, and you are now ready to begin the cross section interpretation.
182 Chap. 6 / Cross Sections
Cross sections are more easily prepared if you use a computer. In a few simple steps, a line
of section is laid out with proportionally spaced wells that can be hung on a given datum. The
datum is readily changed and the logs are easily manipulated in making interpretations. The pro-
cedures are described in the section Cross-Section Construction Using a Computer.
As the first step in an interpretation, we recommend that all recognized correlation markers,
as well as locations of faults in each well, be indicated on the logs used in the section. The faults
can be connected from well to well to reflect the proposed fault interpretation. Lastly, in devel-
oping your interpretation, lines can be drawn from well to well connecting the correlation mark-
ers, with the lines being offset across faults (Fig. 6-5). Straight-line sections are typically used to
initially evaluate a structure (Fig. 6-8a). They portray the dip in straight-line segments, so such
features as variations in dip between wells will not be apparent. Although such a section does not
represent the true attitude of the structure, during the initial phases of a study it provides signifi-
cant information on the general structure, fault geometry, and correlations.
Cross sections usually go through several stages of revisions, with each such revision
improving the accuracy and reasonableness of the interpretation. Remember that the final cross-
sectional interpretation must agree with the completed geologic maps, be geologically reason-
able, have three-dimensional geometric validity, and conform to the structural style of the area.
If possible, cross sections should be retrodeformable or structurally balanced (see Chapter 9).
Figure 6-5 is a completed structural cross section from the Bayou Jean La Croix Field in
Terrebonne Parish, Louisiana. It is a good example of what is called a Finished illustration
(Langstaff and Morrill 1981) structural cross section prepared from electric well logs (1 in. = 100
ft) and finished structure maps on the various horizons shown. The section is geologically rea-
sonable for the tectonic setting, illustrating the subsurface structural geology, including fault
geometry, correlations, and areas of hydrocarbon accumulation. The section was prepared true to
scale with the same horizontal and vertical scales shown graphically in the lower right-hand cor-
ner of the section.
Stick Sections
An alternative to electric logs is the use of log sticks in the preparation of a cross section. A stick
is defined as a vertical or deviated line that represents an electric log. A stick section has several
advantages over the electric log section, including simplicity, clarity, and ease of construction
(Lock 1989).
Since sticks do not show any correlation data (stratigraphic correlations or faults), it is nec-
essary to record the depth of all pertinent correlations and faults obtained from the actual electric
logs. Stick sections are often used to solve structural problems because of their simplicity and
lack of clutter. A typical stick structure cross section is shown in Fig. 6-2. This is the same cross
section shown in Fig. 6-1, which incorporates the actual electric well logs.
sen datum is horizontal. By using a horizontal datum, the distorting effects of structure (folds and
faults) are eliminated. This is equivalent to unfolding and unfaulting the strata. Figure 6-6 is an
example of two stratigraphic cross sections from southwest Kansas/northwest Oklahoma, each
hung on a specific stratigraphic datum. These sections were prepared as part of a detailed study
to evaluate the complexities of the structural and stratigraphic factors controlling the trapping of
hydrocarbons in the Morrowan Sandstones (Mannhard and Busch 1974).
In preparing structural cross sections, recall that you had the choice of using the actual elec-
tric well logs or stick representations of the logs. For stratigraphic cross sections, the actual well
logs must be used, since the work involves solving problems related to the lithology. Changes in
lithology from well to well can be evaluated only with real log data.
In the preparation of a stratigraphic section, the choice of an appropriate marker bed to use
as the datum is extremely important. The choice of a poor datum can pose problems, such as
incorrectly illustrating the original configuration of the stratigraphic units under study. For exam-
ple, a poorly chosen datum may incorrectly depict an actual channel sand as a bar.
Remember, one of the primary objectives in laying out a stratigraphic section is to recon-
struct the sand geometry at the time of deposition or shortly thereafter. When working in areas of
predominant sand/shale deposition, keep in mind that sands and shales compact to differing
degrees. The effects of differential compaction as the result of sediment burial are recorded on
the electric logs. If you have a good idea of the environment of deposition for the sands being
evaluated, the choice of the datum may be relatively easy. If the environments are in question,
however, you should ask yourself which marker is most likely to have been close to horizontal at
the time of deposition.
Figure 6-7 illustrates three cross sections of the Pennsylvanian Anvil Rock Sandstone, with
each section having a different reference datum. In the upper section, the No. 7 coal seam above
the sand is chosen as the datum. With this reconstruction, the cross section does not show the
effects of draping over the sand. In the middle section, the top of the Anvil Sandstone is used as
the datum. By using this datum, the sand may give the appearance of a channel fill sand regard-
less of the sand’s actual original configuration. The bottom section using the No. 5 coal seam
below the sand is the best choice in this particular situation because this last section depicts the
Anvil Sandstone as a channel-fill sand that shows the effects of differential compaction. The
channel sand interpretation is also supported by the lateral truncation of the 5A and 6 coal seams
against the sand. If the intent was to reflect the geometry at the time the channel was active and
the Anvil Rock Sandstone was deposited, the middle section hung on the sandstone itself would
come closest to showing this geometry.
We emphasize that care must always be taken when choosing the stratigraphic datum for a
stratigraphic cross section. There are no actual rules of thumb to apply when choosing the datum.
However, the answers to the initial questions that should be posed prior to constructing a section
often can serve as a guide in your choice of the best datum.
geologic interpretation. With this method, straight lines are drawn from well to well representing
the horizon and fault correlations. Obviously, the straight lines between wells may not illustrate
the true geologic picture, such as changes in bed dip between control points. However, in the ini-
tial stage of a geologic study, this type of section is very helpful in evaluating alternate correla-
tions and fault and structural interpretations. Remember, the interpretation must be geologically
reasonable and have three-dimensional geometric validity.
(a)
(b)
Figure 6-8 (a) Straight-line problem-solving cross section. (b) Finished illustration cross
section constructed from completed fault and structure maps (compare this with Fig. 6-8a).
Finished Illustration (Show) Cross Sections 187
cross section. The final illustration cross section constructed from all the fault and structure maps
is shown in Fig. 6-8b.
We can review Fig. 6-9a and b to illustrate the actual procedure for constructing this type
cross section. The upper part of Fig. 6-9a is a structure contour map on the 6850-ft Sand. The line
of section, drawn on the structure map, starts at the downthrown trace of Fault A, continues
through Wells No. 3, 9, and 10, and terminates at the –7100-ft contour line upthrown to Fault D.
It is always a good idea to review the planning and preparation procedures presented earlier
before beginning any cross section until you feel that you are thoroughly familiar with them. The
first task for a manually drawn cross section is to prepare the cross-section coordinate grid paper,
with the logs proportionately spaced, and then begin plotting all appropriate available data points.
Since all the data are acquired from completed maps and accepted log correlations, it is not nec-
essary to place the actual electric logs on the section; however, this is a matter of personal pref-
erence. The first points to plot are those for the top of the 6850-ft Sand, taken from the structure
map in Fig. 6-9a (upper). Starting with Well No. 3, the first data point to place on the section is
the top of the sand in this well, which is at a depth of –6703 ft. This point, shown at Well No. 3
on the structure map, is plotted at the appropriate depth on the log stick for Well No. 3 on the
cross section (position A). The second data point is the intersection of the section line and
the –6700-ft structure contour line on the structure map, 100 ft southeast of Well No. 3. This
–6700-ft subsea depth point is plotted on the cross section 100 ft from Well No. 3 (position B).
The third data point is the intersection of the section line with the –6600-ft contour line on the
structure map. This point is 725 ft southeast of Well No. 3 and is plotted on the cross section at
position C.
What is the fourth data point? It is the intersection of the section line and the upthrown trace
of Fault B, which is at an estimated subsea depth of –6585 ft and measures 900 ft from Well No.
3. This point is plotted on the cross section at position D.
This procedure is continued for each measurable data point along the line of section on the
structure map. Although there are no wells north of Well No. 3, the section extends to Fault A
using all contour lines as data points. It should be noted that measurable data points include such
items as gas/oil contacts, oil/water contacts, and upthrown and downthrown fault traces. You will
also notice that there is an obvious structural crest about 150 ft south of Well No. 9. The crestal
position and its subsea depth have been estimated and incorporated into the cross section.
Remember that this section is being constructed after the structure maps are in their final stages
of completion and all data shown on the maps should be used in the construction of the cross sec-
tion.
Each point for this cross section is identified on the structure map along the section line and
plotted on the cross section. When all the data points are plotted and connected with a correla-
tion line, a detailed interpretation of the 6850-ft Sand is illustrated in cross-sectional view (lower
part of Fig. 6-9a). This procedure should be completed for all the sands planned for the cross sec-
tion.
Turning to Fig. 6-9b upper, we see a contour map for Fault B with the line of section drawn
on the map. Following the same procedure outlined for the 6850-ft Sand, the location of each
fault data point is identified on the fault surface map and plotted on the cross section in the lower
part of the figure. The same procedure is followed for Faults A, C, and D.
We now have completed the plotting of all the available data points for the 6850-ft Sand and
all the faults for the finished illustration cross section in Fig. 6-8b. The interpretations of the other
188 Chap. 6 / Cross Sections
(a)
Figure 6-9 (a-upper) Structure map on the top of the 6850-ft Sand. Line of cross section shown on map. (a-
lower) Details for construction of a finished illustration section for the 6850-ft Sand. (b-upper) Fault surface
map for Fault B. Line of section shown on map. (b-lower) Detailed construction for Fault B for incorporation
into finished illustration cross section.
Finished Illustration (Show) Cross Sections 189
(b)
sands were constructed using the same procedure just outlined. The result is a detailed cross sec-
tion representing an accurate picture of the structural interpretation of the horizons and faults in
a vertical plane.
In summary, a finished illustration cross section is important for several reasons. It serves as
a visual aid for presenting a completed geologic interpretation and can also be helpful in identi-
fying mapping problems that might otherwise go undetected, particularly in fields with closely
spaced productive sands. An accurately detailed cross section might indicate any number of map-
ping “busts,” such as areas where thickness compatibility between horizons has not been hon-
ored. The cross section in Fig. 6-10, prepared to evaluate cross-fault drainage, is an example of a
detailed finished illustration cross section constructed almost exclusively from final structure
maps for each of the horizons shown on the section. The section shows that the interpretation for
each horizon, the faults, and the intersection of the faults with the individual horizons is geolog-
ically reasonable for the most part. However, a few areas on the cross section appear to indicate
some minor geologic busts, which are highlighted by asterisks. Although none of these problems
seem to be serious, the fact that they are visible shows the sensitivity of this detailed cross sec-
tion to interpretation error.
When mapping several very closely spaced horizons, it is not too difficult to cross contours
from one horizon to the next, if extreme care is not taken during mapping. It is good practice to
underlay the structure map currently being prepared with one already constructed on a horizon
immediately above or below. Such practice ensures compatibility in the structural interpretation
for a series of horizons being mapped and prevents major mapping busts. This topic is discussed
in detail in Chapter 8, but it is introduced here to show how detailed cross sections are used to
identify errors in structure contouring.
At times it is even possible to cross structure contours from two mapped horizons that are
actually hundreds of feet apart, resulting in a mapping bust. Figure 6-11 illustrates just such a sit-
uation. Figure 6-11a shows a structure maps on two different horizons. The prospective strati-
graphic section for an exploratory play lies between these mapped horizons (P-5 and P-7), which
are separated vertically by nearly 800 ft of stratigraphic section in the off-flank position labeled
“T” on the structure maps. A cursory review of the structure maps indicates that they appear geo-
logically reasonable.
Based on a study of the growth activity of a salt dome on the west, it was evident that the
salt mass was a positive feature during the time of deposition of the prospective section. So the
section between the P-5 and P-7 horizons should thin in the up-structure position to the west, and
it does. The stratigraphic thinning was incorporated into the structural picture, but because care
was not taken during the mapping process, the two structure maps depict the two mapped hori-
zons not only converging but actually crossing up-structure. The detailed cross section of the
P-5 and P-7 horizons in Fig. 6-11b shows the effect of this mapping bust very clearly. Notice that
the two mapped horizons intersect at the location marked C and then cross, creating a negative
structural volume, which is an impossible geologic situation. This mapping bust is due to care-
lessness in not using the technique of underlaying the completed P-5 structure map when con-
structing the P-7 map. Such a mistake can put a question in the minds of management as to the
reliability of the work and jeopardize what might otherwise be a great exploration prospect. The
preparation of a quick cross section would have caught this mapping bust.
A precise finished illustration cross section can identify many mapping problems and pro-
vide the time to correct the work before the final interpretation and maps are prepared. Also, these
cross sections are excellent aids for reviewing the possibility of juxtaposed reservoirs across a
fault (Fig. 6-10). Studies of such sand occurrences are important in evaluating the possibility for
Finished Illustration (Show) Cross Sections
Figure 6-10 A detailed finished illustration cross section constructed almost exclusively from com-
pleted maps, including fault maps, structure maps on the top and base of each of the seven strati-
graphic units shown, and a salt map. Notice that the detailed section indicates that the 6 Sand down-
thrown to Fault B is in juxtaposition with the 9B Sand upthrown to Fault B. Cross-fault drainage was
evident from production data and confirmed by this detailed section. Asterisks indicate locations of
possible mapping errors. The original scale of the cross section was 1 in. = 100 ft; therefore, detail
191
on the order of tens of feet could be incorporated into the section.
192 Chap. 6 / Cross Sections
(a)
(b)
Figure 6-11 (a) Completed structure maps on the P5 and P7 horizons. These two horizons
bracket a prospective section trapped by salt to the west and faults to the north and south. (b)
Structure cross section prepared from the completed structure maps on the P5 and P7 hori-
zons. The cross section clearly indicates that the completed structure maps were not con-
structed correctly.
Correlation Sections 193
cross-fault drainage from one fault block to another (Smith 1980). Another type of cross section
used for evaluation of juxtaposition across a fault is described in the section Fault Seal Analysis.
CORRELATION SECTIONS
In this segment of the chapter, we discuss a special type of cross section called a correlation sec-
tion. This section is a special type of stratigraphic cross section that is primarily used as a detailed
correlation aid. There are several important guidelines helpful in preparing this type of section.
1. Choose a stratigraphic datum that best serves the intended purpose of the correlation cross
section (see Fig. 6-7). A reliable shale marker is usually a good choice for a stratigraphic
datum (Sneider et al. 1977).
2. Limit the section to a short vertical log interval in order to show significant correlation
detail.
3. Position the logs as closely as possible with no horizontal scale to include as many logs as
needed in the section.
A correlation section can serve as an excellent correlation aid in defining the lateral and ver-
tical continuity of permeable units within a specific area and stratigraphic interval. These sections
can also be used as good prospecting tools to evaluate and illustrate the potential for hydrocar-
bons.
Figure 6-12 shows the layout for a typical correlation section. The actual electric log show-
ing all the curves can be used; however, to reduce clutter and provide sufficient space for the
inclusion of many logs, the SP or gamma ray and the amplified resistivity curves are often suffi-
cient. As mentioned in Chapter 4, the amplified resistivity curve is usually the most helpful for
correlation work. This discussion of layout and correlation procedure applies to both manually-
prepared and computer-based correlation sections (see the section Cross-Section Construction
Using a Computer).
In preparing a correlation section, follow the same correlation procedures outlined earlier in
Chapter 4. Since a correlation section is a type of stratigraphic section, hang it on a well-defined
datum such as the one shown in Fig. 6-12a. Next, correlate the shale sections using all correlat-
able shale markers indicated by the SP and resistivity curves. Figure 6-12a shows the shale cor-
relations for this section. Observe the parallel to semiparallel trend of the shale markers, which
indicates a fairly uniform stratigraphic thickness. A time-stratigraphic framework now exists
within which you can correlate permeable units. We use sands in this example.
We begin the sand correlations as shown in Fig. 6-12b. As can be seen, the sands are not as
vertically and laterally continuous and uniform as the shale sections. Lateral changes in deposi-
tional environment can cause sudden variations in sand thicknesses even within an area of limit-
ed lateral extent. Because the clays and silt that make up the shales are usually deposited in quiet
waters over large areas, abrupt changes in stratigraphic thicknesses in shales is not common. Such
extensive and consistent deposition of shale usually provides for good lateral correlation conti-
nuity from well to well.
The 8300-ft, 8500-ft, and 9200-ft Sands are separated vertically from the other sands, but
they appear laterally continuous from well to well. The stippled pattern for the 9000-ft Sand is
representative of the rest of the section. Within this overall gross sand interval there are several
distinct sand members in three of the wells. Based on this correlation section, it is speculative
whether there is vertical or lateral continuity of individual sand members from well to well. As
194 Chap. 6 / Cross Sections
(a)
Figure 6-12 (a) Correlation section is first laid out by hanging the logs on a reference
datum and correlating all recognizable shale markers. (b) Completed correlation section
shows the lateral and vertical continuity (or lack of continuity) of the individual sands
seen in each well.
for the other individual sands shown in each log, they appear to be laterally discontinuous from
well to well, indicating rapid changes in depositional environments and laterally limited sand
bodies. The rectangular areas between wells represent discontinuity of sands somewhere between
the wells.
This type of information regarding the continuity of sands is most important to the develop-
ment geologist and reservoir engineer. The layout of one such section or a number of sections can
Cross Section Design 195
1. Which well or wells must be perforated to maximize the drainage efficiency within a
reservoir?
2. Which sand interval or intervals should be perforated within a single wellbore to optimize
hydrocarbon recovery?
3. Is a specific reservoir competitive with an adjoining lease operator? In other words, can
another operator’s well on an adjacent lease drain your reserves because of possible
lateral continuity of sands across the lease? If so, what action must be taken to protect your
reserves?
4. Are any additional development wells required to optimize field production?
5. Can remaining reserves, identified in an abandoned well, be recovered with other existing
wellbores? By this we mean, is there continuity of the hydrocarbon sand from the location
of the abandoned well to a well capable of being completed?
We mentioned earlier that correlation sections can be used as an excellent exploitation tool.
Figure 6-13 shows part of a geologic and engineering prospect package designed to justify the
drilling of a development oil well into what was considered a depleted/nonproducible oil reser-
voir.
Figure 6-13a is a structure map on the top of a prospective sand called the “9300-Foot Sand.”
An oil reservoir is present upthrown to Fault A. Six wells penetrated the oil reservoir with Wells
No. 4 and 10 having produced minor amounts of oil and gas. Due to the minimum amount of oil
production from Well No. 10 and the minimum gas production and pressure decline in Well No.
4, the reservoir was considered depleted. However, further study, with the use of detailed maps,
log correlations, perforation data, production data, and the correlation section A-A' (Fig. 6-13c),
revealed that the reservoir consisted of three distinct sands: (1) a small, highly calcareous fringe
complex, such as that seen in Wells No. 7 and 10; (2) a major cut-and-fill channel sand seen in
Wells No. 3 and 4; and (3) an upper transgressive sand member separated from the fringe and
channel sands by a shale break. This transgressive sand member is seen in all wells.
The correlation section (Fig. 6-13c) shows that the oil production in Well No. 10 was from
the small fringe complex, and the gas production in Well No. 4 from the upper transgressive sand
member. Reserves were calculated volumetrically, from net sand and net hydrocarbon isochore
maps, and they were quite large compared to the actual oil and gas production. This suggested
that the fringe complex and transgressive sand are separate members not in communication with
the main channel sand (see delineation of major channel sand on the structure map Fig. 6-13a,
shown as a permeability barrier, and on the net sand map Fig. 6-13b, highlighted as the limit of
major channel sand). Therefore, significant reserves remain to be produced in this channel sand,
which can be recovered by a recompletion in Well No. 3, if possible, or through the drilling of a
new well. This prospect used the correlation section as an integral part of the prospecting process,
as well as a final illustration to present the idea to management.
(a) (b)
(c)
Figure 6-13 (a) Structure map on the top of the 9300-ft (hydrocarbon-bearing) Sand
trapped upthrown to Fault A. (b) Net sand isochore map of the 9300-ft Sand delineating
the limit of good quality major sand development. (c) A detailed correlation section
through the 9300-ft Sand Reservoir trapped upthrown to Fault A. The section clearly illus-
trates that the upper transgressive member, the fringe complex, and the channel sand are
separate and distinct members of the 9300-ft Sand package. Production data indicate
minor oil production from Well No.10 and that the transgressive member perforated in Well
No. 4 pressure depleted (it is a closed system separated from the main 9300-ft Sand by
a continuous shale break).
1. Cross sections can be run from well to well in a zigzag pattern or as a straight line with
data from each well not on the line projected into the line of section. Both types of sections
have inherent problems. Zigzag sections tend to distort the subsurface geology due to
changes in the strike direction of the section. If you treat a zigzag section as a series of two-
well straight sections, interpretation problems can be minimized.
Straight-line sections may require the projection of well data over long distances,
resulting in well projection problems. If you understand the various methods and
Cross Section Design 197
limitations of projecting well data, however, straight-line sections can be used very
effectively.
2. Deviated wells can be included in a cross section if the line of section runs along the plan
view path of the deviated well (Fig. 6-15).
3. If a line of section being laid out intersects two closely spaced wells, include the well that
penetrates the deepest section if the total depths are significantly important. Spacing may
not be the critical factor; instead, similar structural geometry may be critical. In such cases,
if two closely spaced wells reflect different geometry, choose the well that illustrates the
geometry expected in the cross section.
4. When preparing both strike and dip sections in the same field, it is good practice to tie the
sections together with a specific well placed on both sections (Fig. 6-14).
Use common sense in laying out all sections. Remember, a cross section is intended to help
you visualize the structure in three dimensions, give you another perspective view of the struc-
ture, and serve as an aid in solving a variety of problems related to general correlations, fault
geometry, or the structural interpretation.
Extensional Structures
Figure 6-14 shows the layout for two dip sections and one strike section for a typical extension-
al structure. The dip sections, which are perpendicular to the strike of the faults, provide the best
information for studying the faults. For single fault systems, these sections can be balanced to
help develop the best structural interpretation. Sections involving bifurcating or compensating
fault systems are difficult to balance due to out-of-the-plane motion, which is often difficult to
account for in balancing (see Chapter 11). Initially, the cross sections can be used as problem-
solving sections to help delineate the fault and structural geometry for the area under study. At
the initial stages of the geologic study, the straight-line method of construction for faults, hori-
zons, salt bodies, unconformities, and other features is recommended. During the later stages of
mapping, the sections can be upgraded or revised to help develop and illustrate the final inter-
pretation. Also, cross sections such as section A-A′ in Fig. 6-14 can be useful in resolving possi-
ble correlation problems in deviated wells.
Figure 6-14 Typical layout of cross sections in an extensional tectonic setting. (Map was
computer-drafted.)
Figure 6-15 Typical layout of cross sections for a complex piercement salt structure includ-
ing both straight and vertical wells. (Map was computer-drafted.)
Cross Section Design 199
includes a deviated well follows the path of that deviated well from the surface to total depth. The
portion of the line of section that follows a deviated well path is dashed on the figure for clarity.
Once the structural interpretation is complete, the sections can be upgraded to serve as dis-
plays to illustrate the final interpretation. The structure maps represent the horizontal view and
the cross sections show the vertical view of the three-dimensional geometry of the structural
interpretation.
Look at sections A-A′ and B-B′ in Fig. 6-15. With these sections, the general fault geome-
try in the southern portion of this field can be studied because the sections are laid out perpendi-
cular to the strike direction of the faults. They can also be used to study the juxtapositioning of
sands to evaluate the possibility of cross-fault drainage. The cross sections C-C′, D-D′, and
E-E′, which are dip sections, provide information on the correlations, structural dip, sediment/salt
interface, growth characteristics of the structure, and data on the down-dip extent of any hydro-
carbon accumulations.
Finally, notice that section E-E′ is laid out with no well control. This section may be impor-
tant in evaluating the structural interpretation developed for the eastern flank of the field. Such a
section is constructed using the detailed illustration cross-section techniques discussed earlier. It
is constructed solely from completed structure, fault, and salt maps.
Compressional Structures
The most common hydrocarbon trap in compressional areas is found in hanging wall anticlines,
which form as a direct result of thrust faulting and include such structures as fault propagation
folds, fault bend folds, and duplexes (Chapter 10). The anticlines commonly exhibit an asymme-
try with steep frontal limbs and elongated longitudinal axes (B-axes) perpendicular to transport
direction. In order to study the internal geometry of these plunging compressional folds, cross
sections are normally laid out perpendicular to the B-axis, i.e., parallel to transport direction.
Figure 6-16 is a structure map on top of the Upper Triassic Nugget Sandstone in the Painter
Reservoir and East Painter Reservoir Fields. Cross sections A-A′ and B-B′ shown on the map are
laid out parallel to transport direction and perpendicular to the strike direction of the thrust faults.
In compressional settings it is best to construct straight-line sections and plunge-project well data
into the line of section to preclude the possibility of physically moving structural geometry with-
in the fold. Such a section will also provide the best information for studying the thrust faults and
balancing the structural interpretation.
Figure 6-16 Cross-section layout for a typical compressional structure. Structure map is on the
Upper Nugget Sandstone in the Painter Reservoir and East Painter Reservoir Fields, Wyoming.
(Modified from Lamerson 1982. Published by permission of the Rocky Mountain Association of
Geologists.)
Vertical Exaggeration 201
VERTICAL EXAGGERATION
Earlier in this chapter, we mentioned that, whenever possible, cross sections should be con-
structed using the same horizontal and vertical scales (true-scale sections). Special considerations
may require that a section be prepared with different (exaggerated) scales, particularly when con-
structing large regional or semiregional sections. Typically, it is the vertical scale that is exag-
gerated. Vertical exaggeration can be incorporated into both structural and stratigraphic cross sec-
tions. Keep in mind that with the use of a vertically exaggerated scale comes various types of dis-
tortion, such as that of stratigraphic unit thickness and dip angles of horizons and faults.
Figure 6-17 Cross section layout for a strike-slip fault system with associated
faulted anticlines.
202 Chap. 6 / Cross Sections
HL
VE = (6-1)
VL
where
VE = Vertical exaggeration
VL = Length of unit distance on the vertical scale
HL = Length of unit distance on the horizontal scale
Say, for example, that you wish to prepare a cross section with a horizontal scale of 1 in. =
10,000 ft and a vertical scale of 1 in. = 2000 ft. By using Eq. (6-1), the vertical exaggeration for
this cross section is
10, 000 ft
VE =
2000 ft
VE = 5
If a map has a horizontal scale of 1 in. = 4000 ft and you wish to prepare a cross section with
a vertical exaggeration equal to 4, Eq. (6-1) can be rearranged to determine the vertical scale
required for this exaggeration.
HL
VL =
VE
Therefore,
4000 ft
VL =
4
VL = 1000 ft
We mentioned at the beginning of this section that there are certain situations in which a
cross section with a vertical exaggeration is required; in fact, it may have some advantages over
a cross section with equal scales. Several situations that may require a cross section with a verti-
cal exaggeration are (1) the preparation of a cross section in an area of low structural relief, (2)
the construction of a section in an area of gently dipping beds, (3) a section that would be unrea-
sonably long with equal scales, or (4) the need for extensive vertical detail. Consideration must
also be given to the cost and size limitations of available reproduction equipment. Figure 6-18
shows two cross sections across the Uinta Basin, Colorado, which are the same except for scales.
The upper section has a vertical exaggeration of 12; the lower section is true scale with the same
horizontal and vertical scales. Notice how much detail can be shown on the upper, vertically
Vertical Exaggeration
(a)
(b)
Figure 6-18 (a) A generalized cross section across the Uinta Basin, Colorado, showing the effect of verti-
cal exaggeration (12 times). (b) Same general cross section constructed to true scale. (Modified from Suter
203
1947; AAPG©1947, reprinted by permission of the AAPG whose permission is required for further use.)
204 Chap. 6 / Cross Sections
Tan [exaggerated dip (δE)] = (VE) Tan [true dip (δ)] (6-2)
Using Eq. (6-2), the exaggerated dip for any cross section can be calculated if the true dip
and vertical exaggeration are known. Likewise, if the exaggerated dip and vertical exaggeration
for a horizon or fault are known, the true dip can be calculated. Figure 6-19 is a graphic repre-
sentation of Eq. (6-2) from Langstaff and Morrill (1981). We can look at an example to illustrate
the use of the graph. Referring once again to Fig. 6-18, the Dakota Formation has a dip of 50 deg
at location A, and the cross section has a vertical exaggeration of 12. By entering the graph on
the X-axis at 50 deg, representing the exaggerated dip of 50 deg, and entering the Y-axis at 12,
representing the vertical exaggeration, the intersection of the two lines generated from these data
points indicates a true dip of 5.7 deg. You can see by this example that there can be significant
exaggeration of dip on a cross section with a large vertical exaggeration.
As pointed out earlier, thickness of stratigraphic units is also affected by vertical exaggera-
tion. If we consider that the exaggeration in the vertical direction on a cross section is a function
of Eq. (6-1), it becomes apparent that interval thickness in cross section varies as a function of
the exaggerated dip. Figure 6-20 shows two cross sections. Figure 6-20a is plotted true to scale
and shows a unit of constant thicknesses “T.” Figure 6-20b shows the effect on the unit’s thick-
ness with a vertical exaggeration of two. Notice that in the area of low dip the thickness of the
interval increased by a factor of two. In the area of steep dip, the interval thickness has been atten-
uated as a result of the exaggerated scale. It is interesting to note, however, that even though the
apparent true bed thickness (stratigraphic thickness) is attenuated in the area of steep dip, the ver-
tical thickness is still increased by a factor of two. This effect can also be seen in Fig. 6-18 across
the Uinta Basin. In the area of Rangely Dome, we see the effect of attenuation of the units on the
flanks of the dome and the apparent thickening of the units over the crest of the structure and in
the synclines on each side of the dome. In the preparation and review of cross sections with ver-
tical exaggeration, be sure to keep such thickness changes in mind.
The decision to use vertical exaggeration in the construction of a cross section must be based
upon the scale of the section and its intended use. Employed wisely, vertical exaggeration can be
a very useful tool. Finally, we emphasize that the use of both a vertical and horizontal graphic
scale is necessary in all constructed cross sections, whether they are true scale or incorporate
some type of exaggeration.
PROJECTION OF WELLS
It is best to construct straight cross sections directly through wells; however, for various reasons
this may not be possible, so it sometimes becomes necessary to project a well into a line of
Projection of Wells 205
12 5.7 deg.
VE
50 deg.
dE
Figure 6-19 Solutions to Eq. (6-2) for different values of VE (vertical exaggeration), and (true
dip). True dip occurs where VE = 1. (From Langstaff and Morrill 1981. Published by permission
of the International Human Resources Development Corporation Press.)
(a) (b)
Figure 6-20 (a) is constructed to true scale. (b) shows apparent attenuation due to vertical exaggera-
tion. Attenuation is greatest where true dip is steepest. Notice in the area of steep dip that, although the
stratigraphic thickness is attenuated, the vertical thickness is twice the vertical thickness of that shown
in (a), which is drawn with a true scale. (Modified from Langstaff and Morrill 1981. Published by per-
mission of the International Human Resources Development Corporation Press.)
206 Chap. 6 / Cross Sections
section. There are several ways in which a well can be projected into a line of section (Fig. 6-21).
These include
1. Plunge projection
2. Strike projection
3. Up-dip or down-dip projection
4. Normal-to-the-section-line projection (minimum distance method)
5. Parallel-to-fault-strike projection
All these methods can be used to project well data into a cross section. The best method to
use, however, depends upon various factors, including (1) the structural style of the area being
worked, (2) the orientation of the section to the axis of the structure, (3) the horizontal distance
of a well from the line of section, (4) whether there are faults on the section line, and (5) the gen-
eral dip of the structure. The most commonly used methods for projecting wells into a line of sec-
tion are the plunge and strike projections.
Plunge Projection
In structural settings such as compressional fold-and-thrust belts containing plunging structures,
the preferred method for projection of well data into a line of section is parallel to the B-axis of
the fold in the hanging wall, either up-plunge or down-plunge. Figure 6-22 is a map view of a
cylindrical fold. The longitudinal or B-axis of the fold is shown as B-B' , which is coincident with
the axial line (W. Brown 1984a). The dashed lines represent plunge lines (fold elements) which
are, by definition, parallel to the B-axis in a cylindrical fold. Notice that the dip rates are the same
along each individual fold element, including the axial line. This is true only where the plunge
rate is constant for the portion of the fold depicted.
One of the main objectives in plunge-projecting well data into a line of section is to preclude
the possibility of physically moving structural geometry within the fold. By definition, a cylin-
drical fold is one in which all fold elements are parallel to one another and to the B-axis, which
is parallel to the direction of plunge. A good way to visualize this is to equate a cylindrical fold
Figure 6-21 The five most common methods of projecting well data into a line of section include (1)
plunge, (2) strike, (3) up-dip or down-dip, (4) normal-to-the-section-line (minimum distance), and (5)
parallel to fault. (Modified from W. Brown 1984a; AAPG©1984, reprinted by permission of the AAPG
whose permission is required for further use.)
Projection of Wells 207
Figure 6-22 Map view of a plunging cylindrical fold. The longitudinal fold axis (B-axis)
is shown as B-B' . The dashed lines represent plunge lines that parallel the B-axis.
(Modified from W. Brown 1984a; AAPG©1984, reprinted by permission of the AAPG
whose permission is required for further use.)
to a cylinder of pipe. The edges of the pipe represent fold elements and are parallel to one anoth-
er for the entire length of the pipe. Thus, the shape of the pipe’s cross section is the same at each
end and everywhere in between. Similarly, a cylindrical fold contains fold elements which are
everywhere parallel, and the cross-section geometry is the same throughout the length of the
cylindrical portion of the fold. Therefore, projection of fold geometry into cross section A-A'
(Fig. 6-22) along a line that is not parallel to the plunge lines (e.g., sides of the pipe) would result
in the attempted physical movement of geometry to an improper position within the fold (e.g., pro-
jection of a 38-deg dip from plunge line No. 6 to plunge line No. 5, which has only 24 deg of
dip).
We now consider the application of plunge projection. Figure 6-23 is a structure map of a
plunging cylindrical fold cut by a reverse fault. We want to construct cross section A-A' as shown
on the structure map. Since only Wells No. 1 and 7 actually lie on the line of section and we wish
to include more data than that available from the two wells, we have no choice but to project the
additional wells into the line of section.
Well No. 1 is drilled off the flank of the structure and Well No. 7 is in the footwall in the
opposite fault block. Wells No. 2 through 6 can be projected into the line of section in several dif-
ferent ways: most commonly, either parallel to structural strike (dashed arrow lines) or parallel
to plunge (solid arrow lines). In the former case, Wells No. 2 through 5 are projected into A-A'
208 Chap. 6 / Cross Sections
along structural strike (parallel to the contours), and Well No. 6 cannot be projected at all. Figure
6-24, the cross section obtained by projection along structural strike, is obviously very confusing
and results in an unacceptable interpretation of the fault shape, as well as the relationship of
hanging wall and footwall rocks.
The direction of plunge of the B-axis is indicated on the structure map in Fig. 6-23 to show
the direction in which the wells must be plunge-projected. The B-axial direction can be obtained
by analysis of any form of dip data. The solid arrows show the planned projection path of the
wells into the section. Next to each projection is a “+” or “–” sign. These signs indicate whether
the structural elevation for a specific horizon or structural marker (fault, dip rate, etc.) in the wells
must be adjusted positively or negatively for the direction and rate of plunge over the distance
projected into the section line. For example, Well No. 2 has a subsea depth for the horizon of
–1340 ft. When we project this well along plunge, the top of the horizon intersects the cross sec-
tion at a subsea depth of –1410 ft. Therefore, the top must be adjusted downward by 70 ft to cor-
rect for the change in elevation. This correction in elevation can actually be calculated with
trigonometry by using the angle of plunge and the length of the projection from the well site to
the line of section. Each well must be corrected in this manner, with (+) values indicating
Projection of Wells 209
Figure 6-24 Structural interpretation along cross section A-A' (Fig. 6-23) based on
strike-projecting well data into the line of section. The strike-projected data results
in an unacceptable interpretation. (Modified from W. Brown 1984a; AAPG©1984,
reprinted by permission of the AAPG whose permission is required for further use.)
elevations raised because the well is projected up-plunge, and (–) values indicating elevations
lowered because the well is projected down-plunge.
Figure 6-25 is the completed cross section using all the data from the wells projected into
the line of section parallel to plunge, or the B-axis (in the hanging wall). In order to differentiate
between the wells that actually lie on the section line from those that have been projected, the
wellbore stick for each projected well is dashed. Notice also that the numerical sequence of wells
(1 through 7) is now in proper order.
In conclusion, for plunging folds, the plunge-projection method offers the most accurate
means of projection of structural geometry for the hanging wall block into a cross section. This
projection may be up-plunge or down-plunge, or both, and structural elevations likewise are
adjusted up or down. The projections are made in order to ascertain the true cross-sectional shape
of the structure.
Strike Projection
If you are mapping in a geologic setting involving diapiric structures or in a tectonic area with
nonplunging folds, projection along strike is often more beneficial than other types of projec-
tions. In the areas of extensional or diapiric tectonics, the dip rate on structures is typically
210 Chap. 6 / Cross Sections
Figure 6-25 Structural interpretation along cross section A-A′ (Fig. 6-23) based on
the well data projected into the line of section parallel to plunge of the B-axis. The
projection of data along plunge provides the most accurate method of projecting the
fold geometry into the cross section. (Modified from W. Brown 1984a; AAPG©1984,
reprinted by permission of the AAPG whose permission is required for further use.)
constant along structural strike (Fig. 6-26). Therefore, in order to preserve structural geometry as
well as stratigraphic relationships (syndepositional structures), wells should be projected parallel
to strike.
As you look at each example cross section, imagine each to be a seismic section into which
wells are projected. Consider whether the well data would tie closely with the seismic interpre-
tation.
We first look at an example of projecting along strike in an area without faults (Fig. 6-27).
In Fig. 6-27a, Well No. 2 is projected parallel to strike into the line of Section A-A′. In cross sec-
tion, the M Horizon point projects into the correct structural position on the section and the rate
of dip at that point would be proper for the cross section. Using this method, the projection does
not require any correction factors for elevation or dip rate. In Fig. 6-27b, Well No. 2 is projected
along strike and again projects correctly into the line of section B-B′.We can conclude that in this
type of structural situation, projection along strike is preferred.
Now we look at some examples of projecting a well along strike of a horizon cut by faults
(Fig. 6-28). Figure 6-28a is a portion of a structure map showing a west-dipping structure cut by
an east-dipping normal fault. East-west cross section A-A′ is plotted. Wells No. 1, 2, and 4 lie on
the section line. Since data from Well No. 3 is important to the interpretation, the well must be
projected into the line of section. Considering the specific geologic conditions, Well No. 3 is pro-
jected parallel to strike into the A-A′ cross section, as shown below the structure map in the
Projection of Wells 211
figure. The top of the horizon projects into the section in the correct structural position and main-
tains the correct structural dip, but the depth of the fault in the projected well is incorrect and must
be adjusted. This adjustment is required because as Well No. 3 is projected from its actual posi-
tion into the line of section, the distance from the well to the fault changes.
(a) (b)
Figure 6-27 (a) Projection of well data along strike. Upper figure is a portion of a structure contour map. The
lower figure is cross section A-A′ illustrating the projected data for Well No. 2. (b) Projection of well data along
strike. Upper figure is a portion of a structure contour map. The lower figure is cross section B-B′ illustrating the
projected data for Well No. 2. (Published by permission of Tenneco Oil Company.)
212 Chap. 6 / Cross Sections
(a) (b)
Figure 6-28 (a) Strike projection of well data on a horizon cut by a normal fault. The strike direction of the fault
does not parallel the strike of the mapped horizon. Upper figure is the structure contour map. In the lower figure,
cross section A-A′ illustrates that the well depth point for the mapped horizon fits the cross section; however, the
fault location in Well No. 3 is too deep and must be adjusted. (b) Parallel-to-fault-strike projection. The well data
are projected parallel to the strike of the fault surface. The fault projects correctly into the line of section, but the
horizon depth must be adjusted. (Published by permission of Tenneco Oil Company.)
You might already have asked yourself the question, if the well is projected parallel to the
fault, how will the fault and horizon project into the line? Figure 6-28b depicts this situation. Well
No. 3 is projected into the section line parallel to the strike of the fault surface. With this projec-
tion, the fault location in the well fits the section correctly, but the horizon point is projected too
shallow, requiring an elevation adjustment from –590 ft to –605 ft.
When projecting wells along strike of a horizon cut by faults, either the horizons or the faults
usually will require an elevation adjustment. Therefore, we caution that projected wells often
have limited use in a cross section and can, at times, cause more confusion than clarity. If a well
must be projected into a line of section, be sure to identify the projected electric log or wellbore
stick with a dashed line and clearly note any required corrections or adjustments.
Projection of Wells 213
Figure 6-29 Projection of well data into a line of section, using the normal-to-section and down-dip projection
methods. Neither method is recommended for projecting well data. (Published by permission of Tenneco Oil
Company.)
214 Chap. 6 / Cross Sections
Figure 6-30 Structure contour map with line of section A-A′. Deviated Well No. 6 is
projected into the line of section parallel to horizon strike.
4000-ft, 5000-ft, 6000-ft, and 7000-ft true vertical depths are not the same distance from the sur-
face location as the actual depths plotted along the directional path. This correction must be made
when the individual directional survey points are plotted in cross section. If each survey data
point is projected correctly and plotted correctly in cross section, the projected directional well-
bore stick will depict the true representation of the directionally drilled well in cross section.
When the actual electric log for a deviated well is used to project into a cross section, the log
must be cut at various locations to retain true vertical or subsea true vertical depth, whichever is
used. In Fig. 6-30, we see that the projected depth points plotted on the map are not the same dis-
tance from the surface location as the actual plotted depths. Therefore, adjustments must be made
to the log in order to include it in the cross section. These adjustments are made by cutting out
small sections of the log at various depths, as needed, to maintain true vertical or subsea depth
on the section after projection.
Figure 6-31 shows a deviated well log that is cut at various depths to retain its true depth in
cross section after projection. It is best to delete sections of the log that do not contain any impor-
tant correlations. If possible, shale sections should be used to make these adjustments.
Projection of Wells 215
Figure 6-31 The film of the log for Well No. 212 is cut at various depths to retain
true vertical depth of the entire log in cross section.
216
Chap. 6 / Cross Sections
Figure 6-32 North-south structural cross section in West Cameron Block 192, northern Gulf of Mexico.
The section line (see structure map insert) has a zigzag pattern to accommodate the directionally drilled
wells, thus minimizing the projection of well data into the line. (From Offshore Louisiana Oil and Gas Fields,
v. II, 1988. Published by permission of the New Orleans Geological Society.)
Cross Section Construction Across Faults 217
Figure 6-32 is a north-south structural cross section through West Cameron Block 192 Field,
northern Gulf of Mexico. The section is composed from six vertical and six directionally drilled
wells. The structure map on the U Sand is placed as an insert on the cross section. Notice that the
line of section has a zigzag pattern from north to south. Apparently, this zigzag pattern was used
to accommodate directionally drilled wells. In other words, the path of the section line changes
direction to parallel as closely as possible the path of the directional wells. By doing so, the pro-
jection of data from each deviated well into the line of section is minimized, thereby providing
the most accurate representation of the directionally drilled well data in the cross section.
1. Be certain that the velocity function is the most accurate one for the area. Incorrect
velocity information can cause some serious errors when projecting a well, particularly if
it involves a directionally drilled well. Also, remember that a seismic section is not a
geologic cross section and that the apparent angle of a well on a seismic section may not
resemble the angle of the well on the equivalent geologic cross section. It is also important
to know both the horizontal and vertical scales of the seismic section being used.
2. Consider dashing all projected wells posted on a seismic section. If the well path actually
crosses the plane of the line, mark the point clearly with a different color solid line or a
horizontal dash across the well path.
3. Remember that when a well is projected along structural strike, many of the faults may
appear to intersect the projected well path at different depths than the actual intersections
in the well. Be ready to field questions when a line like this is used for presentation
purposes.
4. Too often, the minimum distance (normal-to-the-line) projection method is used for
projecting well data into a seismic section. Whenever possible, do not use this method,
because the data frequently will be projected incorrectly, resulting in a conflict between the
well and seismic data (Fig. 5-14a and b).
5. It is not good practice to project well data over long distances into a seismic section.
Remember the geometry of any structure changes laterally even over short horizontal
distances. Projections of well data over long distances can and often do cause great
confusion in the interpreted seismic section.
missing or repeated section in a wellbore, resulting from a fault, as being equal to this fault com-
ponent. Therefore, when constructing a cross section with faults, the vertical separation must be
used to correctly displace a horizon or stratigraphic unit from one fault block to another. In this
section, the method for using the vertical separation in the preparation of a cross section con-
taining a fault is discussed in detail. The concept is basic and the correct technique of construc-
tion very simple, yet the technique is often misused. Before reading this section, it may be nec-
essary to become familiar with the fault component vertical separation and the difference
between this component and fault throw. These subjects are covered in the beginning of Chapter
7, particularly in the sections entitled Definition of Fault Displacement and Fault Data
Determined from Well Logs.
If the missing section in an electric log is mistakenly used as throw in the construction of a
cross section involving a normal fault, chances are the displacement of horizons across the fault
will be incorrect. The values for vertical separation and throw are the same only for a vertical
fault and for a fault that cuts through horizontal beds. With increasing bed dip, however, the val-
ues are different and can be significantly so with steeply dipping beds. Therefore, the bed dis-
placement error in a cross section, using the missing section as throw, will usually be greater with
increasing structural dip.
Figure 6-33 graphically depicts the correct method for using vertical separation (missing
section) to obtain the proper bed displacement across a normal fault. The incorrect method using
throw as the missing section is also shown for comparison. In this example, assume there is suf-
ficient control from well data or a structure map in the downthrown fault block to establish the
dip of the horizon (Bed A) and its intersection with Fault 1 in the downthrown block at –7100 ft,
as shown in Fig. 6-33a. The missing section for Fault 1 determined by log correlation is 300 ft.
With this information, the next step is to determine where Bed A intersects the fault in the
upthrown fault block. This is accomplished by correctly plotting the 300 ft of missing section as
vertical separation across the fault into the opposite fault block. Vertical separation is defined as
the measured vertical distance between a bed projected from one fault block across the fault to a
point where the projection is vertically over or under the same bed in the opposite block. The
missing or repeated section observed in a wellbore as the result of a fault is equal in value to ver-
tical separation and not throw. Applying this definition to the normal fault in the cross section in
Fig. 6-33b, the dip of Bed A is projected from the downthrown block through the fault, as if the
fault were not there, to the upthrown block until it is 300 ft vertically beneath the fault. The depth
at this point on the fault (–6900 ft) is the upthrown intersection of Bed A with the fault, and Bed
A is then constructed from that point.
Now look at the missing section plotted incorrectly as throw (Fig. 6-33c). Throw is defined
as the difference in vertical depth between the fault intersection with the bed in one fault block
and the fault intersection with the same bed in the opposing fault block, determined in a direc-
tion perpendicular to the strike of the fault. Applying this definition to the cross section in Fig.
6-33c, which is perpendicular to the strike of Fault 1, the depth of intersection of Bed A with the
fault in the upthrown fault block is –6800 ft.
There is a 100-ft vertical difference between the two methods in the estimated depth of the
upthrown intersection of Bed A with the fault. Incorrectly using missing section as throw places
the upthrown block 100 ft too high. This mistake could be costly, particularly if a well were
planned for the upthrown block based on the incorrect interpretation, which plotted the amount
of vertical separation as throw.
Because this technique is so important, we will review it with two other geologic examples,
one involving a normal fault and the other a reverse fault. Figure 6-34 shows cross section A-A'
Cross Section Construction Across Faults
(a)
(b)
219
(c)
220 Chap. 6 / Cross Sections
Figure 6-34 An illustration using the correct method for displacing a horizon
across a normal fault.
Cross Section Construction Across Faults 221
and an insert enlargement of a portion of the section. The missing section in Wells No. 2, 3, and
5 for the normal Fault 1 is 400 ft. The missing section is equal to vertical separation, so we use
this fault component value to prepare the cross section construction across Fault 1.
The structural attitude of the 6000-ft Horizon in the upthrown block of Fault 1 is established
by using the available well control in that block and a completed structure map. Well No. 5 is suf-
ficiently close to the intersection of the 6000-ft Horizon with the fault to establish the depth at
which this intersection occurs. Once this depth has been established, we must determine the posi-
tion (depth) at which the horizon intersects the fault in the downthrown block. The depth chosen
again depends upon your understanding of missing section. You must use the definition of verti-
cal separation to project the dip of the 6000-ft Horizon from the upthrown block, through the fault
as if the fault were not there, into the downthrown block until the projection is 400 ft vertically
above the fault. The point labeled I in the insert is the correct intersection of the horizon with the
fault in the downthrown block. Notice that the projection of the 6000-ft Horizon from the
upthrown to downthrown blocks is curved to follow the actual changing dip of the unit.
Remember, the missing section is not equal to throw in areas involving dipping beds.
Therefore, if the mistake is made of using 400 ft of missing section as throw, then the estimated
depth of the intersection of the horizon with the fault in the downthrown block would be incor-
rect. In this case, shown in the insert in Fig. 6-34, the intersection point would be too shallow by
approximately 75 ft.
We now look at one example involving a reverse fault. The vertical separation for a reverse
fault is equal to the repeated section in a well cut by the fault. Figure 6-35 shows cross section
A-A' and an enlarged insert of a portion of the cross section. The repeated section resulting from
Fault 1, as correlated with all surrounding well control, is 400 ft, which is the vertical separation
of the fault.
Assume, for this example, that due to more well control, the attitude of the 5000-ft Horizon
and its intersection depth with Fault 1 is first established in the hanging wall block. The task now
is to determine the displaced position of the horizon in the footwall block and the depth of its
intersection with the fault. With reverse fault geometry, the technique is actually easier than with
a normal fault and can be used to establish two control points for the 5000-ft Horizon in the
opposing fault block (in this case the footwall). Working with the enlarged portion of the figure,
follow along the horizon in the hanging wall until the bed is at a point that is 400 ft vertically
above the fault surface. This point, labeled I, is the footwall intersection of the horizon with the
fault. A second point of control for the 5000-ft Horizon can be established by measuring 400 ft
vertically down from the intersection of the 5000-ft Horizon with the fault in the hanging wall.
This point, labeled I' , corresponds to the depth to the top of the horizon in the footwall at the point
vertically beneath the intersection of the sand with the fault in the hanging wall.
As an academic exercise, estimate the intersection of the 5000-ft Horizon with the fault in
the footwall using the incorrect assumption that the 400 ft of repeated section corresponds to the
throw of the fault. As a second exercise, estimate the size of the actual fault throw considering
the correct geometry shown in Fig. 6-35.
It is important to note again that the preciseness of the projection of horizon dip, through a
fault, determines the accuracy of the displacement across the fault. If the bed dip changes in the
area where the projection is made, the projection must follow this changing bed dip (observe
again the curved projection of the 6000-ft Horizon in the enlargement insert in Fig. 6-34). In other
words, the projection of a horizon across a fault need not be a straight line. If the bed dip is not
constant across the fault, then the projection must follow the changing bed dip for you to accu-
rately construct the vertical separation.
222 Chap. 6 / Cross Sections
Figure 6-35 An illustration using the correct method for displacing a horizon
across a reverse fault.
Three-Dimensional Views 223
THREE-DIMENSIONAL VIEWS
Many types of computer software provide three-dimensional views of mapped surfaces and rock
volumes. These are more easily generated and preferable to hand-drawn two-dimensional repre-
sentations of three-dimensional geology. However, the latter are sometimes appropriate or nec-
essary, so we present here some methods that help in three-dimensional visualization of the geol-
ogy.
In petroleum exploration and exploitation, a map or cross section alone cannot always rep-
resent the complete subsurface geologic picture, because each is limited to two dimensions. Since
neither a map nor cross section presents data in three dimensions, a detailed study may require
the use of additional techniques that aid in visualizing the geology in three dimensions. Such
additional aids include log maps, stick or fence diagrams, isometric projections, three-dimen-
sional structural models, and block diagrams.
Log Maps
Log maps are the simplest means of combining plan view information with some vertical dimen-
sional data. Figures 6-36 and 6-37 illustrate two different types of log maps. In Fig. 6-36, elec-
tric log sections and data charts for the Miocene Middle Cruse Sand Member in Trinidad are
superimposed on a structure map of the same sand member. This type map allows you to quick-
ly review the vertical sand conditions and producing characteristics of the sand in each well with
respect to the well’s structural position within the reservoir. Notice that only the SP and the
amplified short normal resistivity curves were used in the log sections. By looking at this map,
you can very quickly see the complexity of the one or more sand members compromising this
reservoir and the relationship of sand quality to structural position and well location.
Figure 6-37 is another type of log map in which the SP curve for each electric log is placed
next to the appropriate well location in an attempt to portray the three-dimensional paleogeo-
graphic conditions during the time of sand deposition. In this field, the sands appear to be of two
types: (1) a major cut-and-fill channel sand, and (2) a very poor quality, highly calcareous fringe
complex. The calcareous nature of the sand was determined from cores taken in a number of
wells.
If the actual structure map for the 10,500-ft Sand were superimposed on this log map in a
manner similar to that shown in Fig. 6-36, this additional step would improve the evaluation of
the sand conditions with respect to the localization of the major channel sand and calcareous
fringe complex sand in comparison to the structural position of the individual wells, production
characteristics of the local reservoirs, and hydrocarbon-trapping conditions. Such information
can lead to the identification of overlooked workovers and recompletions, in addition to possible
development drilling locations.
Fence Diagrams
Fence diagrams, also called panel diagrams, consist of a three-dimensional network of cross sec-
tions drawn in two dimensions. They are designed to illustrate the areal relationship among sev-
eral wells that are located in close proximity to each other. Fence diagrams can either be struc-
tural or stratigraphic. Figure 6-38 is a typical example of a fence diagram in which actual elec-
tric log segments have been traced for use at each well location. Fence diagrams are often pre-
pared using only log sticks, although log sticks do not provide the correlation detail usually need-
ed for detailed work. Fence diagrams are also usually diagrammatic because they seldom have a
common horizontal and vertical scale.
224 Chap. 6 / Cross Sections
Figure 6-36 Log map of the Miocene Middle Cruse Sand Member (Trinidad). The SP and amplified short
normal log curves are superimposed on the structure map of the sand. The relationship of sand quality to
structural position within the reservoir can quickly be evaluated. (Modified after Bower 1947; AAPG©1947,
reprinted by permission of the AAPG whose permission is required for further use.)
Fence Diagram Construction. The construction of all fence or panel diagrams begins with a
well base map. The plane of the base map represents a chosen datum plane and each well loca-
tion is taken to be the point where the well intersects the datum plane. A vertical line is drawn at
each well control point and the well data are either hand-plotted along this line or a film of the
actual log curves, reduced to the proper scale, is attached to the line. For example, in Fig. 6-38,
the SP and resistivity curves were hand-traced from reductions of the actual log. Figure 6-39 is a
fence diagram constructed using the actual reduced log curves.
In the preparation of a structural fence diagram, the reference plane of the map is a chosen
datum, which can be sea level or any elevation above or below. For example, the chosen datum
Three-Dimensional Views
Figure 6-37 Another type of log map. The SP curve for each well is placed at the
surface location for the well in an attempt to visualize the three-dimensional charac-
225
teristics of the 10,500-ft Sand. (Published by permission of Texaco USA.)
226 Chap. 6 / Cross Sections
Figure 6-38 Layout of a typical stratigraphic fence diagram in which the actual electric
log segment for each well was traced at the well location. (Modified from Boeckelman,
unpublished. Published by permission of author.)
for the fence diagram in Fig. 6-40 is –1500 ft, which is a subsea datum equivalent to 1500 ft
below sea level. The log sections or hand-traced curves are hung on the vertical line at each well
location and are referenced to the chosen datum. The structural profile of the bed(s) on the sec-
tion is drawn by initially connecting the major correlation markers from each well. The number
of correlation markers connected depends upon the detail desired for the diagram.
It is important to try to minimize the interference of one panel with another to reduce the
effect of masking data. It is a good idea to begin to complete the panels starting with the east-
west sections first, and in particular, the panels nearest the lower or front end of the map (see pan-
els A, B, and C in Fig. 6-38). The northwest-southeast or northeast-southwest panels are com-
pleted next, up to a point where they intersect the east-west panels from the front or where they
disappear behind these panels from the rear, such as panels E and F in Fig. 6-38. Once all the cor-
relation markers have been connected on all the panels and the desired labeling added, the dia-
gram is complete. If dips are most prominent in the north-south direction, an isometric projection
may be more helpful, or the diagram may be drawn with north toward the right instead of toward
the top of the map, as is more common.
Three-Dimensional Views 227
For the construction of a stratigraphic fence diagram, the plane of the map is considered to
be either the bedding plane or an unconformity that is the chosen datum horizon. Just as with the
structure fence diagrams, a vertical line is drawn at each well location. Since most stratigraphic
fence diagrams are prepared to illustrate correlations, facies changes, permeability barriers, sed-
iment wedging, or some other stratigraphic feature, the actual log sections are normally required.
Connect all the correlation markers to complete the fence diagram (Figs. 6-38 or 6-39).
Figure 6-39 Layout of a stratigraphic fence diagram using the actual reduced log
curves. (From Langstaff and Morrill 1981. Published by permission of the International
Human Resources Development Corporation Press.)
Isometric Projections
Isometric projections are sometimes used to emphasize certain stratigraphic variations that might
not be apparent on the conventionally oriented fence diagram. In an isometric projection, the map
base is shown as if it were turned at an angle and tilted toward the front, which transforms the
projection from orthogonal to nonorthogonal axes (Langstaff and Morrill 1981). Figure 6-41
shows the difference between a standard fence diagram and its transfer into an isometric projec-
tion. Notice that the rectangular base map becomes a parallelogram in the isometric projection.
The base map used in Fig. 6-41 has a coordinate grid system with both north-south and east-
west scales provided. Any isometric projection will have distortion in any direction except in the
directions of the original grid system. In other words, all lines parallel to the original north-south
and east-west axes remain parallel in the isometric projection. Therefore, any data points to be
measured from the base map must be measured along one or both of these coordinated grids, and
by doing so they can be measured accurately. Measured distances in any direction except paral-
lel to the original grid coordinates will be distorted. They may be greater or less than the actual
distance. It is very important, therefore, to always include some type of coordinate grid system
on all map bases that are planned as isometric projections.
(a)
(b)
Figure 6-41 Comparison of (a) a standard stratigraphic fence diagram to (b) an isometric
projection. (From Langstaff and Morrill 1981. Published by permission of the International
Human Resources Development Corporation Press.)
230 Chap. 6 / Cross Sections
to reduce clutter on the model. The second choice is to hand-trace the reduced log curves on the
vertical line at each well location. The choice is more a matter of preference.
Our interest, in this case, is to hang the section on the top of the 9200-ft Sand in order to
study and review the entire 9200-ft Sand package. At Well No. 3, for example, the top of the
9200-ft Sand is at a depth of –9150 ft. Since the referenced datum is –9100 ft, the first depth point
is 50 ft below the datum. The log or hand-drawn curve(s) is positioned so that the top of the sand
is 50 ft below the reference grid at Well No. 3, as shown in Fig. 6-43d. For Well No. 1, the top of
the sand is at a depth of –9450 ft, which is 350 ft below the referenced grid datum of –9100 ft.
Therefore, the top of the sand represented on the actual log or hand-drawn curve(s) is placed 350
ft below the grid at the Well No. 1 location. The depth below the referenced datum is calculated
for each well, and the appropriate log curve(s) is posted at that location.
Now what about the 18 other grid intersections? Using the grid overlay on the structure map
of the 9200-ft Sand (Fig. 6-43b), determine the structural elevation for the top of the unit at each
grid intersection, calculate the difference between the unit top and the reference depth, and post
the structural top at the appropriate depth on the vertical line at each separate grid intersection.
After posting the elevation for the structural top of the 9200-ft Sand at each of the 24 grid inter-
sections, connect the points with straight lines as shown in Fig. 6-43d. These connected lines rep-
resent the top of the 9200-ft Sand on the three-dimensional model.
The control points connected with straight lines may not accurately represent the attitude of
the sand top, since the true top of the sand has curvature between many control points. The
process can be improved further by posting the depth values for all structure contour/grid inter-
sections and connecting these depths. Approximately 30 depth control points for contouring the
top of the sand are added using this technique. With these additional control points, the top of the
sand can be contoured to more closely represent the true attitude of the sand. Also, grid lines can
Three-Dimensional Views 231
(a)
(b)
Figure 6-43 (a) Completed structure map on the top of the 9200-ft Sand. (b) RAM grid posi-
tioned at 90 deg to dip. Grid lines must intersect each well to ensure the use of all geolog-
ic data. (c) Choose a datum and construct lines from the datum vertically straight down from
each well and grid location. (d) Post the elevation for the top of the 9200-ft Sand at each
grid intersection and connect the points with straight lines representing the top of the 9200-
ft Sand. (e) Overlay the grid system on the base of the sand map and post the base of the
sand at each grid intersection and each well location. (f) Begin detailed construction of the
sand and shale data for each panel starting at each well location, working outward in all
panel directions. (From Boeckelman, unpublished. Published by permission of author.)
232 Chap. 6 / Cross Sections
(c) (d)
(e) (f)
be added at strategic locations to refine the RAM as desired. The amount of detail depends upon
the individual interpreter and the intent of the reservoir model.
From the log sections for the six well locations, the base of the sand can be recognized. But
what about the base of sand at the other 18 grid intersections? We must now repeat the process
outlined above for the structural base of the sand. Overlay the grid over a structure map on the
base of the sand, determine the structural elevation of the base of the sand at each grid intersec-
tion, and mark this base on each vertical intersection line. Connect all the depth points for the
base of the sand, and the three-dimensional panels for the 9200-ft Sand are complete, as shown
in Fig. 6-43e. Notice that all the panels are at right angles to one another, so very little data on
any particular panel are hidden by other panels.
Finally, using the net sand and shale data from each of the electric logs, begin construction
of the model panel by panel. Start at each well location and work outward in all panel directions
to prepare a complete interpretation of the sand/shale distribution for the 9200-ft Sand. This
sand/shale work can be done in a general way or detailed based upon each foot of analyzed elec-
tric log (Fig. 6-43f).
This three-dimensional reservoir analysis model can be used to accurately evaluate the dis-
tribution of sand and shale, and to aid in developing a depositional model for the overall sand
package as well as the individual members. If hydrocarbons are present, water levels can be plot-
ted and an entire color scheme developed to represent sand, shale, hydrocarbons, and water.
There are advantages to this model not associated with other types of cross sections, fence dia-
grams (Fig. 6-44), or isometric projections.
Figure 6-44 Typical fence diagram layout for the 9200-ft Sand. (From
Boeckelman, unpublished. Published by permission of author.)
234 Chap. 6 / Cross Sections
Figure 6-45 Compare this three-dimensional RAM with the fence diagram in Fig. 6-38, which was
prepared from the same well data.
this model with the fence diagram in Fig. 6-38, which is constructed from the same well data.
Notice how clearly the structure is represented in the three-dimensional model (Fig. 6-45) com-
pared to the fence diagram (Fig. 6-38). Also observe how much more data are presented with
minimal obstruction in the three-dimensional model. Finally, true length and slope can be meas-
ured on the model, which cannot be done with the fence diagram.
Missing Section
Figure 6-46 Eight-well, closely spaced correlation cross section set up to show the SP/GR track of a series of
electric logs. The orientation of the section is constantly changing as all wells in the field study were included for
correlation purposes. Missing section due to faulting is restored using a fault-gapping tool. Wells numbered one
through eight at the top of the section are all Mesa Petroleum-operated wells. (Published by permission of W. C.
Ross and A2D Technologies.)
236 Chap. 6 / Cross Sections
(a)
(b)
Figure 6-47 Stratigraphic (a) and structural (b) cross sections of Almond Formation and Lewis Formation strati-
graphic relationships in the Green River Basin of Wyoming. Wells are spaced proportional to map distance. Facies
changes and sandstone pinchouts, interpreted in the inter-well space in stratigraphic mode (a), are automatically
repositioned properly in structural mode (b), and tops picked in the time-stratigraphic gap are used to determine
and draw the position of onlap and erosional truncation terminations above and below the sequence boundary
respectively. (Published by permission of W. C. Ross and A2D Technologies.)
Cross Section Construction Using A Computer 237
stratigraphic geometries, wells can be spaced proportional to map distance (or projected to a line
of section). For wells with deviated wellbores, the program used should be able to adjust the spac-
ing of the wells to be proportional to map distance using the latitude and longitude positions
along the chosen datum. Cross-section software should support the picking and termination of
correlations, facies boundaries, and unconformities in the space between wells (Figs. 6-47 and
6-48). As wells are re-ordered or the section re-datumed from stratigraphic to structural views,
the inter-well interpretation shifts appropriately (Fig. 6-47a and b).
Missing Section in Stratigraphic Cross Sections. Building stratigraphic cross sections can
be enhanced by restoring those portions of the cross section that are missing due to erosion,
nondeposition, or faulting. Unconformity/fault tools allow geologists to split or gap logs propor-
tionally to reflect the amount of section lost due to erosion or faulting. An example of restoring
missing section due to channel erosion is illustrated in Fig. 6-48a, where a siliciclastic package
of marine shales and shelf sandstones is truncated by an incised valley system. Figure 6-48b
shows the same system where the logs have been split, or gapped, at an unconformity (sequence)
boundary. The gap space represents missing time due to erosion (in the valley) and nondeposi-
tion (in the interfluves). Correlations from the marine shale interfluve are extended into the miss-
ing section gap and represent an estimate of stratigraphic geometry prior to erosion. The ability
to pick tops within these gaps allows geologists to hang cross sections on tops that have been cut
out by erosion. Tops picked within the channel feature are extended into the interfluve space.
When the gaps are closed, the software automatically determines the appropriate onlap and trun-
cation geometries in the inter-well space, above and below the unconformity surface respective-
ly (Fig. 6-48a).
(a)
(b)
Figure 6-48 Stratigraphic cross sections of shallow marine and incised valley-fill sediments. (a) Shallow marine
shelf facies are shown truncated by an incised-valley system. The base of the incised valley system is a
sequence boundary (SB). Erosional truncation (ET) of shelf strata below the sequence boundary and onlap (OL)
of channel-fill strata above the unconformity are automatically positioned using the timeline geometries shown in
(b). (b) The logs have been pulled apart, or “gapped,” at the sequence boundary to represent missing section due
to erosion and/or nondeposition. The intersection of the sequence boundary and tops picked in the time-
stratigraphic gap are used to determine and draw the position of onlap and erosional truncation terminations
above and below the sequence boundary respectively. (Published by permission of W. C. Ross and A2D
Technologies; and AAPG©1990, reprinted by permission of the AAPG whose permission is required for further
use.)
Cross Section Construction Using A Computer 239
NORTH SOUTH
1 2 3 4
Figure 6-49 Stratigraphic cross section of Pliocene sediments. The missing section fault gaps show the results
of fault identification through detailed correlation and vertical separation (VS) analysis. Estimates of missing sec-
tion (VS), due to faulting, have been performed at three locations and are shown within each missing section fault
gap. Fault names are generic. As a result, no fault correlation is required at this stage of fault interpretation.
Correlations A, B, and C are picked within the missing fault gap on the Mesa Petroleum OCS-G 03587 A-6 well.
(Published by permission of W. C. Ross and A2D Technologies.)
The ability to interactively interpret fault location (depth) in a well and the amount of miss-
ing section (vertical separation), and then to pick markers within these missing-section gaps, pro-
vides the geologist with a powerful set of tools for subsurface interpretations in regions domi-
nated by normal faults. Figure 6-52a and b show stratigraphic (with missing-section gaps) and
structural versions of a cross section from the Wilcox of South Texas (Edwards 2001). The large
faults have cut out hundreds of feet of section. Tops cut out by the faults have been picked with-
in the missing-section gaps across the entire cross section. Missing-section stratigraphic cross
240 Chap. 6 / Cross Sections
NORTH SOUTH
1 2 3 4
Figure 6-50 Stratigraphic cross section from Fig. 6-49 showing fault geometry after geologist assigns a specif-
ic fault name to each generically-named fault. (Published by permission of W. C. Ross and A2D Technologies.)
sections are useful for removing structure and providing an undistorted view of the correlation
framework (Fig. 6-52b).
NORTH SOUTH
1 2 3 4
Figure 6-51 Structural cross section with fault and bed-offset geometry resulting from application of vertical-sep-
aration fault function to each fault-correlation intersection. Note upthrown and downthrown bed offsets of miss-
ing correlations A, B, and C to the left and right of cross-section Well No. 3. (Published by permission of W. C.
Ross and A2D Technologies.)
and (4) possible vertical conductivity of the fault. Given the context of this book, we address only
techniques by which you can determine juxtaposition of rock units across a fault. This determi-
nation provides the structural basis for a complete fault-seal analysis. We describe in detail a
manual structural method and summarize a computer-based evaluation of juxtaposition.
Analysis of juxtaposition typically involves a depiction of the fault surface showing the
traces of stratigraphic units on the footwall and hanging wall. The interpreted fault surface itself
may be used, or a horizontal projection of the traces can be made from the interpreted fault sur-
face onto a vertical plane parallel to the fault. Either method portrays the juxtaposition of units
across the fault. Allan (1989) is generally credited with introducing the technique into the oil
242 Chap. 6 / Cross Sections
(a)
(b)
Figure 6-52 (a) Structural cross section of faulted Wilcox strata in Zapata County, South
Texas. Correlation offsets were calculated using interpolated vertical separation information
interpreted at fault positions in wells. (b) Stratigraphic cross section with missing section, due
to faulting, shown as gaps. Note increasing size of missing section fault gap in a down-dip
direction for lowermost fault. This pattern depicts increasing amounts of vertical separation
along the fault with depth. (Published by permission of W. C. Ross and A2D Technologies.)
Fault Seal Analysis 243
industry, but it was Rippon (1985) who first described the construction of fault plane sections for
use in the coal mining industry in the United Kingdom. Bouvier et al. (1989) apply a technique
known as fault slicing (seismic sections parallel to the trace of a fault) to examine the seal char-
acteristics of faults in offshore Nigeria. Fault surface sections are used extensively to evaluate
sealing characteristics of faults (e.g., Broussard and Locke 1995 and Yielding et al. 1997).
TVT TST
100' 94'
FA
SU U
R L
FA T
C
ALLAN
PLANE
E
Figure 6-53 A cross section of a faulted sand bed showing the correct projection of the sand
from the fault surface to the vertical Allan Plane. (After Broussard and Lock, 1995. Published
by permission of the Gulf Coast Association of Geological Societies.)
V.S. = 125'
(App. throw 90')
TVT TST
100' 94'
FALSE
JUXTAPOSITION
FA
SU U
R L
FA T
C
ALLAN
PLANE
Figure 6-54 A cross section of a faulted sand bed showing an incorrect projection of the sand
from the fault surface to the Allan Plane. False juxtaposition can be indicated by using bed thick-
ness for the base of sand or by comparing apparent throw to bed thickness. (After Broussard and
Lock, 1995. Published by permission of the Gulf Coast Association of Geological Societies.)
Fault Seal Analysis 245
9800' TO
P UT -1
0,
TO 10
P 0'
DT
T
TOP D 10,000' -10
,00
0'
-9900'
A
10,200'
B
3
2
4 -10,070'
1 10,400' -10,000'
-9980'
00' '
-10,0 00
1
'
0,
00
-1
0,2
10,600'
-1
200'
10,800'
0 1000'
Figure 6-55 A fault surface section is constructed by overlaying depth-scaled cross-section paper on a struc-
ture map of the top of sand, with the depth scale perpendicular to the fault trace. The depths of the upthrown
and downthrown traces for the top of sand are plotted on the cross section. Although not illustrated here, the
traces of the base of sand are then plotted from a base-of-sand map. Wells 1, 2, 3, and 4 and Platforms A and
B are used to register the cross section in the same place on both maps. (After Broussard and Lock, 1995.
Published by permission of the Gulf Coast Association of Geological Societies.)
Also, if apparent throw is compared to TVT of the sand and found to be less than the TVT, a false
juxtaposition could be inferred. To avoid these pitfalls, always use both the top and base of sand
to portray the sand face correctly on the fault surface section.
Figure 6-55 illustrates the process of constructing a fault surface section or Allan section
from structure maps. The steps are similar to building a finished illustration (show) cross section,
as described elsewhere in this chapter, except the line of section is the fault surface projected into
a vertical plane. To illustrate the method, we use one sand offset by a fault and evaluate the jux-
taposition of the sand.
1. For the fault surface section, select a convenient depth scale that covers the range of depths
along the fault for the top-of-sand and base-of-sand maps. Overlay the cross-section paper
on the top-of-sand structure map with the vertical depth scale more or less perpendicular
to the fault trace on the map. Mark some registration points, such as well locations, so you
can locate the cross section paper in exactly the same place from map to map. In the
example the vertical exaggeration is 5:1.
2. Determine the depth of the fault trace for the top of sand in the upthrown block and plot it
on the cross-section paper. For example, from the structure map find where the
–10,000-ft contour intersects with the fault trace upthrown and project parallel to the depth
246 Chap. 6 / Cross Sections
scale to the –10,000-ft depth on the cross section and mark a point. Continue along the
upthrown fault trace plotting the depths onto the cross section (interpolating the depth
where necessary). Then connect the points with a line to plot the top of the sand at the fault.
3. Plot the depth of the fault trace for the top of sand in the downthrown block in the same
manner. At intersections with other faults, estimate the depth in the corners of the
intersection, plot these points on the cross section and connect them with a straight line.
This is the top of sand as it would be on the subject fault surface at the intersecting fault,
according to the structure map.
4. Next, although we do not illustrate it, you would repeat the process for a base-of-sand map
by plotting the upthrown and downthrown traces of the base of sand. The same registration
points used for the top-of-sand map will allow you to locate the cross-section paper in
exactly the same place on the base of sand map.
5. Finally, as shown in Fig. 6-56, color or shade any overlap or juxtaposition of the sand, or
sands if you are plotting more than one. Minor variations in contouring between the top-
and base-of-sand maps usually should not affect the juxtaposition interpretation, but they
should be evaluated for their significance to the accuracy of the maps.
-9800'
TO
P (U
T)
-10,000'
TO
P
OF
SA
A BA ND
-10,200' S (D
E T)
B OF
SA
3 2 ND
4 (U
BA T)
-10,400' SE(
1 DT
)
-10,600'
200' -10,800'
0 1000'
Figure 6-56 A fault surface section of a thick sand with the area shaded gray where the sand
is juxtaposed to itself. (After Broussard and Lock, 1995. Published by permission of the Gulf
Coast Association of Geological Societies.)
Commonly, the fault displacement is sufficient to juxtapose different sands. For this situa-
tion, choose a depth interval suitable to the sand(s) of interest in one fault block and then deter-
mine which sand(s) in the other block fall within that depth range. Follow the steps above, using
the structure maps of the top and base of each sand within its respective fault block.
Fault Seal Analysis 247
Plotting the depth of the fault traces for the tops and bases of permeable intervals on cross-
section paper gives a graphical display of the faces of the intervals against the fault surface that
can easily be interpreted for juxtaposition of the intervals. Accurate fault surface maps and top-
and base-of-sand maps are necessary for a valid fault surface section.
Figure 6-57 Footwall (thick lines and dashes) and hanging wall (thin lines and dashes) pro-
jected onto a normal fault surface. Vertical lines with nodes represent the fault surface inter-
section with the vertical seismic lines. (Published by permission of Badley Earth Sciences.)
248 Chap. 6 / Cross Sections
Fault-Seal Potential. Fault seal can be due to juxtaposition of reservoir rock against non-reser-
voir rock or to fault-zone material of high entry-pressure (e.g., clay smear in the fault zone)
(Yielding et al. 1997). Both seal types are dependent on fault displacement and the nature and
distribution of lithologies in the wall rocks (Figs. 6-58 and 6-59). The three-dimensional repre-
sentation of the entire fault surface allows fault-seal attributes to be mapped onto the fault sur-
face. A detailed description of the various methods to predict fault-seal potential, with examples
using fault surface sections, can be found in Yielding et al. 1997.
Figure 6-58 Fault surface showing sands in the footwall of a normal fault. (Published by
permission of Badley Earth Sciences.)
Fault Seal Analysis 249
Figure 6-59 Fault surface showing sands in the footwall (hachure pattern) and the hanging wall (solid pattern),
and sand juxtaposition where the two patterns overlap. (Published by permission of Badley Earth Sciences.)
Figure 6-60 Fault surface showing the overlaps of juxtaposed reservoir units. (Published by
permission of Badley Earth Sciences.)
250 Chap. 6 / Cross Sections
Identification of juxtaposition of reservoir units across the fault surface (Fig. 6-59) uses
horizons mapped from seismic and well data and a detailed reservoir stratigraphy defined by iso-
chores. Probable seals due to juxtaposition (sands on one side of the fault in contact with imper-
meable lithologies, commonly shales) can be clearly identified along the fault surface (Fig. 6-59).
Similarly, overlaps of juxtaposed reservoir units (e.g., sand against sand) can be explicitly dis-
played (Fig. 6-60). Remember that the accuracy of horizon interpretation is paramount to the
precision of this juxtaposition display and analysis, and therefore to the ultimate depiction of the
reservoirs.
Once the overlap of permeable units is determined, as in Fig. 6-60, an analysis can be made
of whether a given contact is likely to support a pressure difference due to lithologic attributes
(e.g., gouge ratio, shale smear). Refer to Yielding et al. (1997) for a detailed discussion.
Conclusions
Fault surface sections, whether depicted as vertical sections parallel to the strike of the fault or as
3D representations of the fault surface, offer a unique view within the 3D data volume. The sec-
tions provide data about the fault, the quality of the interpretation, and especially the fault’s
potential for sealing hydrocarbons. The routine construction and analysis of fault surface sections
should be adopted as a crucial part of the workflow in exploration and production projects.
CHAPTER 7
FAULT MAPS
INTRODUCTION
Faulted structures play a very significant role in the trapping of hydrocarbons. Therefore, it is
imperative that anyone involved in the exploration for, or exploitation of, hydrocarbons should
have a significant understanding of faults within the area of study, including their origin and rela-
tionship to the formation of structures. Detailed interpretation and mapping of major faults is crit-
ical in the process of hydrocarbon exploration and development. This chapter presents the cor-
rect subsurface interpretation and mapping techniques required to prepare fault surface maps.
Chapter 8 presents the techniques for integrating faults into structural interpretations and maps.
Faults themselves are vital to structural development and to hydrocarbon migration and entrap-
ment. A reasonable structural interpretation, in faulted areas, begins with an accurate fault inter-
pretation resulting from the construction of fault surface maps and the proper integration of these
fault maps into the structural interpretation. We refer to constructed fault maps as fault surface
maps rather than the more commonly used term “fault plane maps,” since most fault surfaces are
not true planes.
The data required to construct a fault surface map are obtained from the correlation of well
logs, interpretation of seismic sections, and at times from outcrops. In Chapter 4, we present the
methods and procedures for recognizing a fault in a well log and determining its missing or
repeated section. In this chapter, we discuss the importance of mapping faults and present the
methods for constructing fault surface maps with fault data acquired from electric well logs and
seismic sections.
The preparation of accurate fault surface interpretations and maps requires a strong geolog-
ical background, three-dimensional thinking, and a good understanding of the structural style of
the area being worked. When we make reference to the understanding of structural style, we are
251
252 Chap. 7 / Fault Maps
Figure 7-1 Schematic diagrams of hydrocarbon traps associated with various structural styles. BC,
basement complex; T, displacement toward viewer; A, away from viewer. (From Harding and Lowell
1979; AAPG©1979, reprinted by permission of the AAPG whose permission is required for further
use.) (Salt-related closures modified from Salt Domes by Michel T. Halbouty. Copyright 1979 by Gulf
Publishing Company, Houston TX. Used with permission. All rights reserved.)
Faulted structures can be simple or complex. To provide the best structural interpretation
with the available data, the integrity of the structure must be shown to be sound and geological-
ly reasonable. To provide the most accurate and sound geologic interpretation in faulted areas,
the interpretation, mapping, and validation of the faults is the first step. The construction of fault
surface maps as a fundamental part of any geologic study is absolutely necessary. The integration
of fault surface maps with structure maps is also essential to support the structural interpretation,
to prepare accurate maps, to identify prospects, to design wells to be drilled, and to determine the
volume of potential hydrocarbons. The integration of fault and structure maps is discussed in
detail in Chapter 8.
Too often, geologic interpretations and the accompanying maps and cross sections are pre-
pared without giving much consideration to the three-dimensional geometric validity of the inter-
pretation. Testing the validity of geologic interpretations is discussed in some detail in Chapter
10 under structural balancing, but it needs to be stated here that the proper construction of fault
surface maps and their correct structural integration can go a long way toward providing three-
dimensional validity or consistency to any interpretation.
It is not sufficient to rely solely on what the well logs are indicating or what is seen on the
seismic sections. Cross sections and seismic sections can in themselves misrepresent true sub-
surface relationships by the simple nature of their orientation, as well as by other factors. A good
understanding of three-dimensional geometry is essential to any attempt at reconstructing a sub-
surface picture.
No hydrocarbons have ever been trapped by a fault trace. The trap is along the fault surface
itself. Therefore, the mapping of the surfaces of all-important faults is an integral part of any sub-
surface interpretation, particularly in areas involving multiple faults, where extremely complicat-
ed structural relationships can exist. Attempting to reconstruct a complicated structure by using
isolated fault data from electric well logs or seismic sections without the benefit of reasonable
fault surface maps and their integration with various structural horizons can provide erroneous
geologic interpretations. Shortcuts are often taken in our preparation of subsurface maps. Such
shortcuts include failure to construct fault surface maps, the preparation of a structure map of
only one horizon, or the use of a limited number of seismic sections to generate an interpretation.
In this chapter and in Chapter 8, we show how such shortcuts can often lead to structural inter-
pretations that are misleading, unreasonable, and therefore costly to any exploration or develop-
ment program.
The basic concepts and techniques discussed in this chapter apply to the use of data obtained
from both vertical and deviated wells, in addition to data from seismic sections. The discussions,
illustrations, and practice problems deal principally with extensional and compressional faulting
that reflect mainly dip-slip movement, but the methods are broadly applicable to all styles of
faulting.
FAULT TERMINOLOGY
“Probably no portion of geologic literature has a more confused terminology than that dealing
with faults.” This is a profound statement that is as applicable today as it was when made by
H. W. Straley in 1932. A literature search on the subject of fault nomenclature, or terminology,
shows that as far back as the turn of the century, there was great inconsistency in the use of fault
component terminology.
In 1908 the Council of the Geological Society of America appointed the Committee on the
Nomenclature of Faults. This committee was charged with establishing proper fault nomencla-
254 Chap. 7 / Fault Maps
ture. Since that time, there have been numerous papers on the subject of fault components and
their related terminology, usage, and nomenclature. However, despite these numerous publica-
tions, many geoscientists and engineers are still confused when it comes to the definitions and
usage of various fault component terms.
The correct understanding and usage of fault terminology with respect to certain fault com-
ponents is essential to the preparation of correct subsurface fault and structural interpretations
and maps. Therefore, in this section we discuss the fault components that are important to sub-
surface mapping in the petroleum industry. Many others are not discussed, not because they are
less important, but because they do not apply to the interpretation and mapping techniques that
are presented in this text. Figure 7-2 graphically defines several of the fault components of inter-
est to us.
Figure 7-2 Block diagram of Bed X displaced by a normal fault, illustrating four different
fault components. The front panel is perpendicular to the strike of Fault F-1. (Modified from
Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
In the literature, the definitions and uses of the terms vertical separation, throw and heave
are inconsistent and confusing. The correct usage of fault component terms by all geoscientists
and engineers may never be achieved. However, in order to correctly conduct interpretations and
construct subsurface maps, we must use the fault component terms vertical separation, throw and
heave in a consistent manner. We choose to use the terms as follows.
Vertical separation (AE) is the vertical component of bed displacement. It is measured as
the vertical distance between a horizon (such as the top of a stratigraphic unit) projected from one
fault block across a fault to a point where the projection is vertically over or under the same hori-
zon in the opposite fault block. It is that separation seen in vertical wellbores, vertical shafts, and
vertical cross sections (Dennis 1972).
Definition of Fault Displacement 255
Throw (AC) is the difference in vertical depth between the fault intersection with a horizon
in one fault block and the fault intersection with the same horizon in the opposing fault block,
determined in a direction perpendicular to the strike of the fault surface.
Heave (BC) is the horizontal distance between the fault intersection with a horizon in one
fault block and the fault intersection with the same horizon in the opposing fault block, deter-
mined in a direction perpendicular to the strike of the fault surface.
Missing section is the vertical thickness of the stratigraphic section faulted out of a well-
bore as a direct result of a normal fault cutting through the section. Missing section is sometimes
referred to as fault cut. Repeated section is the vertical thickness of the stratigraphic section
repeated in a wellbore as the direct result of a reverse fault cutting through the section. The miss-
ing or repeated section is determined by correlation of an electric log from one well with other
electric logs from nearby wells, as presented in Chapter 4. Technically, missing or repeated sec-
tion is equal in value to vertical separation. Vertical separation for a fault can also be determined
from seismic data, as shown later in this chapter.
We cannot overemphasize the importance for all geoscientists and engineers to understand
that the missing or repeated section recognized in well logs is NOT throw, nor is it equal to
throw, but rather that it is equal to the fault component vertical separation. A misunderstanding
of the technical point can create, and has caused, significant interpretation and mapping errors
resulting in millions of lost dollars from dry holes, failed recompletions, workovers, and more.
Of special interest is the integration of well log with seismic data. Major interpretation errors
occur regarding fault displacement when seismic data are misinterpreted by using throw, or
apparent throw, determined from seismic sections as if it were equivalent to the missing section
from nearby wells.
Notice in Fig. 7-2 that the value for the throw of the fault is not equal to the value for the
vertical separation. The fact that they are different components of a fault is of significance when
fault data obtained from well logs are used for integrating faults into a structural interpretation
(Tearpock and Bischke 1990). The terms throw and vertical separation are commonly misunder-
stood and, more importantly, often misused in the preparation of subsurface interpretations and
the accompanying fault and structure maps.
Because the understanding of these terms is so important to correct interpretation and con-
struction of subsurface fault and structure maps, we attempt to clarify the issue without causing
more confusion than currently exists. We discuss the terms in a general way and review them with
respect to subsurface mapping techniques. For a complete review of the subject of fault nomen-
clature, refer to the references at the end of the book.
sides of a fault. Separation is apparent movement on a fault with respect to a reference horizon
cut by the fault (Dennison 1968). Separations are measurable, whereas slip is usually calculated.
Numerous authors (including Reid et al. 1913; Hill 1947; Crowell 1959; Dennis 1972; Tearpock
and Harris 1987; Tearpock and Bischke 1990, and others) have emphasized the importance of dis-
tinguishing between fault components related to slip and those components related to separation.
We emphasize that components of the actual slip cannot routinely be measured in the sub-
surface due to a lack of conventional sources of data from which the measurements can be made.
Therefore, slip components are not routinely mappable. Some separation components, on the
other hand, are measurable fault components with conventional subsurface data and therefore are
mappable. They can be measured in a vertical shaft, an electric log from a wellbore, or on a seis-
mic section, regardless of orientation with respect to fault strike. Of these various separation
components, vertical separation is the most important parameter for constructing subsurface fault
and structure maps (Tearpock and Harris 1987).
Throw and heave cannot be measured by correlation of well logs, as we will demonstrate.
Using a seismic interpretation, measurements of throw and heave can be made only if the inter-
preted seismic profile is oriented perpendicular to the strike of the fault surface. The terms throw
and heave have limited practical application in subsurface petroleum interpretation and mapping.
In addition, they cause confusion and significant interpretation and mapping errors. We discuss
their application and relationship to other fault components later in this chapter and again in
Chapters 8 and 9.
We cannot leave this subject with the idea that fault slip is not important. Fault slip and re-
lated components are important, particularly in mining operations. Slip generally can be deter-
mined in mines where the actual fault surface is visible. In fact, the usage of the terms throw and
heave originally came from the coal fields of Great Britain where the strata are nearly horizontal
or the faults are strike faults (Reid et al. 1913). As mentioned earlier, these terms have found their
way into the petroleum industry and are probably here to stay. However, they have limited prac-
tical application in subsurface petroleum interpretation and mapping.
Throw = AC = AB sin θ
where
AB = dip separation
θ = fault dip
Thus, as the fault changes dip, the value of throw must also change. We emphasize at this
point that throw (which is related to fault dip and displacement) cannot be directly measured from
electric well logs. We do, however, present methods that enable you to calculate the amount of
throw, if desired, knowing the vertical separation and other properties, such as fault and bed dips.
However, throw does not normally enter into proper subsurface mapping techniques (Tearpock
and Bischke 1990).
Mathematical Relationship of Throw to Vertical Separation 257
The vertical separation (AE) in Figs. 7-2 and 7-3 is defined as the distance that a bed has
been vertically displaced during faulting (Hill 1947). This distance is of primary importance to
us because the vertical separation is recognized and determined from correlated electric well logs,
as described in Chapter 4. To illustrate this point, consider the following example. Assume that a
structure exists that contains beds that dip uniformly to the west (Fig 7-4). The SP from two wells
drilled into these beds is shown on the figure. The dashed line in Fig. 7-4 represents a future nor-
mal fault. The beds will be displaced in such a manner that the hanging wall portion of Well
No. 1 is placed in juxtaposition with the footwall portion of Well No. 2, as shown in Fig. 7-5.
The geometric configuration produces the following observations in Fig. 7-5. As the hang-
ing wall block is displaced, the top of Bed B in the hanging wall is brought into contact with the
(a)
(b)
Figure 7-3 (a) Block diagram showing surface displaced by normal fault. (b) Vertical cross sec-
tion perpendicular to strike of fault; in same plane as front of hanging wall block of block diagram.
DF is vertical slip; AD is horizontal dip slip; AC is the throw; BC is the heave; AE is the vertical
separation. DF, AC, and AE are all used by some geologists as throw. (Modified from Billings 1972.
Published by permission of Prentice-Hall, Inc.)
top of Bed C in the footwall. Therefore, the missing section in Well No. 1 of the hanging wall
includes the stratigraphic section from the top of Bed B to the top of Bed C. Inspection of the
electric well logs now reveals that the missing section as the result of the fault in Well No. 1 (in
the hanging wall) is represented by the coarsening upward sand sequence and the lower shale sec-
tion present in the hanging wall portion of Well No. 2 (shaded section in Fig. 7-5).
258 Chap. 7 / Fault Maps
Figure 7-4 Hypothetical example. Unfaulted structure with beds dipping uniformly
to the west. Dashed line shows location of future normal fault perpendicular to plane
of cross section. (Published by permission of D. Tearpock and R. Bischke.)
This example clearly demonstrates that the throw of the fault is not equal to the missing sec-
tion in the faulted well. However, the missing section is equal to the vertical separation as defined
in Figs. 7-2 and 7-3. We therefore have shown that electric well logs record vertical separation
and not throw, and that throw does not directly enter into subsurface mapping techniques
(Tearpock and Harris 1987; Tearpock and Bischke 1990). Vertical separation (as well as throw)
varies laterally and with depth on a fault. Expect these variations and be prepared to recognize
them in your interpretation of well and seismic data, and honor the variations in your maps of
faulted structures.
Quantitative Relationship
Vertical separation can be related to throw utilizing the relationships contained in Fig. 7-6
(Tearpock and Bischke 1990). Performing some trigonometry and using the Law of Sines, the
following relationship is developed.
AE sin ( π / 2 – φ a ) = AB sin (φ – θ a )
Substituting
AC
AB =
sin θ
Mathematical Relationship of Throw to Vertical Separation 259
yields
AE sin (φ a – θ)
=
AC [sin θ ⋅ sin ( π / 2 – φ a )]
Equation (7-1) has application in regard to the evaluation of subsurface structure maps
(Tearpock and Bischke 1990). This two-dimensional equation gives us the ability to check com-
pleted structure maps for accuracy of construction when the missing or repeated section is used
to prepare an integrated subsurface structure map (see the section Contouring Faulted Surfaces in
Chapter 8 for definition of integrated structure map). In Chapters 8 and 9 we discuss the appli-
cation of Eq. (7-1) to test the validity of structure maps constructed using well log or seismic data,
and we present a method for analyzing the magnitude of error if mapped incorrectly.
Figure 7-5 Beds are displaced such that the upper portion of Well No. 1 in the hang-
ing wall fault block is juxtaposed with the lower portion of Well No. 2 in the footwall
block. Missing section in Well No. 1 by correlation with Well No. 2 is highlighted on the
SP curve for Well No. 2. The missing section is equal to vertical separation and not to
throw. (Published by permission of D. Tearpock and R. Bischke.)
260 Chap. 7 / Fault Maps
Using data from the first edition of Applied Subsurface Geological Mapping (Tearpock and
Bischke 1991), J. R. Sonnad generated a three-dimensional equation for determining throw. The
geometric relationships here are similar to those presented in Chapter 4 regarding the Setchell
equation. Therefore, the data are used to establish the three-dimensional equation for calculation
of throw when the required fault and structural data are available.
AE (7-2)
AC =
tan φ cos α
1 –
tan θ
where
φ = true bed dip
α = ∆ azimuth between bed dip direction and fault dip direction
θ = true fault dip
and vertical separation, we were unable to find one figure that diagrammatically illustrates the
geometric relationship of missing or repeated section, as seen on an electric log from a vertical
wellbore, with respect to the fault components throw and vertical separation. Figure 7-3, from
Billings (1972), illustrates a vertical cross section perpendicular to the strike of a fault surface
showing such fault components as throw, heave, and vertical separation. Notice that the figure
shows at least three separate fault components defined by various geoscientists as throw, demon-
strating the confusion that surrounds the use of these fault component terms. Although this fig-
ure correctly illustrates the difference between throw and vertical separation, it does not relate
this geometry to what is seen in a well log.
Since fault data are in part derived from well log correlation, it is very important to under-
stand that the true vertical thickness of missing or repeated section, determined in a well by cor-
relation of electric logs, is actually a measurement of the fault component, vertical separation. The
cross section in Fig. 7-7 diagrammatically shows the geometric relationship between the missing
section in a wellbore and the vertical separation of the fault. The east-west structural cross sec-
tion, which is perpendicular to the strike of the fault surface, shows two beds (A and B) that have
been displaced by the normal Fault F-1. This normal fault cuts Well No. 1 at –5230 ft and is dip-
ping at an angle of 45 deg to the east. The beds are dipping at 30 deg to the west. By correlation
with Well No. 2, the missing section in Well No. 1 is determined to be 100 ft and the fault is
shown to have entirely faulted out Bed A in Well No. 1. As shown in the cross section, the throw
of Fault F-1 is represented by the vertical line AC, which is equal to 63 ft. The vertical separa-
tion, represented by the vertical line AE, is equal to 100 ft. Thus, by the correlation of Well No.
1 with Well No. 2 in Fig. 7-7, we have diagrammatically shown that the missing section obtained
for the fault in Well No. 1 is not throw nor equal to throw, but rather is a measurement of the fault
component vertical separation. With this particular set of conditions, the throw of the fault is only
63 percent of the vertical separation. As shown in Fig. 7-7, there can be a significant difference
between the throw and vertical separation of a fault. If mapped incorrectly, this difference can
result in significant error in an integrated structural interpretation and therefore the generated
structure maps.
For the most part, an understanding of fault terms and their application to subsurface map-
ping comes from academic studies and company-sponsored training programs. Textbook discus-
sions on faults commonly use very simplistic examples showing faults cutting horizontal beds.
These examples using horizontal beds lead to the misconception that throw and vertical separa-
tion are the same fault component or have the same value. Throw and vertical separation, how-
ever, have the same value in only three specific circumstances: (1) where the strata being faulted
are horizontal, (2) where the fault is vertical, or (3) in a cross section perpendicular to the strike
of the fault surface, where the fault strike is at a right angle to the strike of the strata. In the lat-
ter situation, the strata will have an apparent dip of zero deg. Figure 7-8a shows the situation
involving horizontal beds where the values for throw and vertical separation are the same. The
use of models where faults cut dipping beds (Fig. 7-8b and c) should eliminate the misconcep-
tion that missing section is always throw, an idea that leads to the preparation of incorrect maps.
The first set of conditions from our list of situations, where vertical separation equals throw,
is discussed in greater detail here because it is the most common situation presented in textbooks
and the one that has resulted in more misunderstanding of the fault component terminology than
any other. Because of exposure only to simplistic examples using horizontal beds, many geosci-
entists and engineers have failed to recognize that the values for these two different fault com-
ponents vary from each other depending upon the structural attitude of the formation. This major
misunderstanding can result in numerous mapping errors in a structural interpretation.
262 Chap. 7 / Fault Maps
Figure 7-7 Diagrammatic cross section illustrates the geometric relationship between the miss-
ing section in a wellbore and the vertical separation of a normal fault. The 63-ft value of throw was
calculated mathematically using Eq. (7-1). (Published by permission of D. Tearpock.)
Figure 7-8b and c illustrate the discrepancy in values of vertical separation and throw where
bed dip is considered. They further show the relationship where the horizons are dipping in the
same general direction as the fault and where the horizons are dipping in the opposite direction
to the fault. For example, as shown in Fig. 7-8b, the missing section in Well No. 1 is 1000 ft
although the throw of the fault is 1500 ft. In Fig. 7-8c, the missing section in Well No. 1 is 1400
ft whereas the throw is only 1000 ft. We point out again that with the well logs, only vertical sep-
aration (missing section) data can be determined. Throw cannot be determined from well log cor-
relation, but it is not important to know in most cases.
Vertical separation is directly measurable from correlation of well logs and is equivalent to
the missing section or repeated section caused by a fault, which is valid regardless of the appar-
ent attitude of any horizon considered. Throw and heave are dependent fault variables that change
with variations in the apparent attitude of the fault and horizon. For most petroleum-related inter-
pretation and mapping, the estimates for throw and heave have mainly academic value. They can
be measured only in a cross section or seismic section that is oriented perpendicular to the strike
of the fault surface, or on a structure map after the map has been completed using fault data cor-
rectly as vertical separation to construct a technically and structurally reasonable map. By using
Fault Surface Map Construction 263
(a)
(b) (c)
Figure 7-8 (a) The values for throw and vertical separation are the same where the displaced beds
are horizontal. (b) Where both the strata and the fault dip in the same general direction, throw is
greater than vertical separation. (c) Where the strata and the fault dip in generally opposite directions,
throw is less than vertical separation. (Published by permission of D. Tearpock.)
Eq. (7-2), however, the measured throw across a fault on a completed structure map can be used
to check the accuracy of the map. This is discussed in detail in Chapter 8.
In an area where a fault serves as the boundary limit of a hydrocarbon reservoir, the trap is
along the fault surface itself. Construction of the geologic picture involves the integration of all
fault surface maps with several key structural horizons. Therefore, construction of an accurate
fault surface map for each important fault, using all available data, is usually the first step in gen-
erating a structural interpretation where faults are present.
Earlier in this chapter we mentioned that the preparation of accurate fault surface maps
requires three-dimensional thinking and a good understanding of the regional tectonics being
studied. This is so because each tectonic setting has its own characteristic patterns of faulting. For
example, most of the faults in areas like offshore Nigeria or the northern Gulf of Mexico Basin
are normal faults typically downthrown to the basin and, although they may strike in any direc-
tion, the preferred strike direction is roughly parallel to the present or historical coastline.
Regional knowledge is very important in developing a geologic interpretation, comparing alter-
native geologic solutions, and generating final subsurface maps that are geologically reasonable.
Note that we use the term fault surface map, or just fault map, instead of the more common
usage of fault plane map. Fault surfaces tend to differ from true mathematical planes. They may
increase or decrease in dip with depth, as well as change strike direction reflecting a sinuous or
angular appearance, which may trend in a specific direction or represent an arcuate shape. In pro-
file they may be listric, antilistric, or kinked. Some fault surfaces are deformed. Fault surfaces are
therefore rarely perfect planes. However, on a very localized or field scale, some faults may
appear planar and can be mapped as such. Some of the basic fault examples in this textbook rep-
resent idealized data used to present and teach a specific technique. In these cases, the fault exam-
ples are often simplified as true planes.
The construction of fault surface maps has numerous benefits in the interpretation and under-
standing of the development of faulted structures. Fault surface maps
In the subsurface, faults can be recognized in one of three ways: (1) through the correlation
and interpretation of electric logs, (2) by the interpretation of seismic sections, and (3) by infer-
ence. In this section, we discuss the use of electric logs to obtain fault data required to construct
a fault surface map. The fault information begins with fault data points in electric logs. These data
points, which represent the intersection of a drilled well with a fault surface, establish the actual
Fault Surface Map Construction 265
(a)
(b)
(c)
Figure 7-9 (a) Normal fault resulting in a missing section. (b) Reverse fault resulting in a
repeated section. (c) Normal fault resulting in a repeated section. In example (c), the beds are
dipping at a steeper angle than the fault. (Modified from Bishop 1960. Published by permission
of author.)
presence of faulting (refer to Chapter 4). For normal faults, the fault is usually represented by a
loss of stratigraphic section, whereas a repeat of section is associated with reverse faults, as
shown in Fig. 7-9. There are exceptions to this generalization, however, such as the special case
of steeply dipping beds cut by a normal fault resulting in a repeated section shown in Fig. 7-9c.
For a normal fault, any key horizon encountered in a well above the fault is typically in the
hanging wall (downthrown) fault block, and any marker encountered below the fault is in the
footwall (upthrown) fault block (Fig. 7-9a and c). The only exception is in a deviated well cross-
ing a fault backwards (Fig. 4-36). For a reverse fault, any horizon encountered in a well above
the fault is in the hanging wall (upthrown) fault block; a marker encountered below the fault is in
the footwall (downthrown) block (Fig. 7-9b). This relationship between the fault pick and any
particular horizon thus indicates whether a well is in the footwall or hanging wall fault block for
any particular horizon being mapped. This relationship is further discussed in Chapter 8.
For each fault point in a well, two values are required for use in the interpretation and con-
struction of a fault map: (1) an estimate of the amount of missing or repeated section for the fault,
which we define as the vertical separation, and (2) an estimate of the subsea depth of the fault in
the well. In Fig. 7-10 the recognized fault data in the log of Well No. 2 are clearly marked to indi-
cate the amount of missing section as a result of the fault (150 ft), the depth of the fault (a meas-
ured depth of 7280 ft), and the well(s) used in the correlation (Well No. 1). If you are still not
sure how to identify faults in electric well logs, refer again to Chapter 4 on electric log correla-
tion.
266 Chap. 7 / Fault Maps
Fault data from at least three wells, not in a straight line, are required to accurately begin to
contour a fault surface in the vicinity of the well control. However, if you are familiar enough
with the area and if data from one or more seismic sections are available, then an accurate fault
map may be constructed with data from just one well. Obviously, the more fault data available,
the better the interpretation of the fault surface. Fault maps also can be prepared from seismic
data alone if the coverage is sufficient. This topic is presented later in this chapter.
Contouring Guidelines
In preparing fault surface maps, certain general guidelines should be followed. If sufficient fault
data are available, the fault surface can be contoured in the same way as the elevations of a key
horizon (Reiter 1947). Inasmuch as faults result from breaks rather than bends in the strata, they
pose some special problems in contouring. The rules differ from the general rules for contouring,
in that angular relationships may exist between two intersecting fault surfaces or between a fault
surface and a horizon. The general guidelines for contouring a fault surface are as follows:
1. Contours of a fault surface may be open-ended. They do not have to close upon
themselves. This is true because faults terminate laterally in the subsurface.
2. Changes in either fault strike or dip are assumed to be gradual unless evidence indicates
otherwise (cross-faulting). An exception to this guideline might occur in the case of
mapping a reverse-faulted ramp and flat surface (see Chapter 10); in these cases the
changes in dip can be abrupt.
3. Changes in fault strike for normal faults are usually represented by smooth curves rather
than by sharp angles. Exceptions to this guideline are deformed fault surfaces and the effect
of cross structures.
4. Changes in dip are generally mapped as smooth curves rather than plane segments
deflecting at sharp angles. Again, there are exceptions to this guideline, including those
listed in guideline 3 and when mapping some thrust faults.
Fault Surface Map Construction 267
5. Use the interpretive method of contouring outlined in Chapter 2 for preparing fault surface
maps.
6. Several fault surfaces may be contoured on a single base map. Contours of individual faults
may merge inasmuch as faults intersect one another in nature. Note: When constructing
compensating, bifurcating, or intersecting faults, denote the lines of termination,
bifurcation, or intersection on the fault maps.
7. Fault surface maps must be geologically reasonable for the area being mapped.
8. Fault surface maps are normally contoured with a 500-ft or 1000-ft vertical contour
interval, since fault surfaces are usually relatively steep. Thrust faults may dip at a low
angle, so a smaller contour interval may be appropriate.
The fault surface map will be integrated later with a structure horizon map to generate a
completed structure map. This procedure is described in Chapter 8. Three key concepts regard-
ing fault contours need to be remembered for the integration process.
1. The contours of a fault surface join those of a given mapping horizon at points of
intersection of fault contours and horizon contours of the same value.
2. Faults commonly dip at different angles than a key horizon being mapped and,
consequently, only some of the fault surface contours intersect with those of a given datum
(see Chapter 8).
3. The fault surface intersects horizons above and below a given horizon unless the fault is
extremely limited in vertical extent.
Fault surface map construction actually involves some subjective interpretation; thus, the
more data available for mapping, the less uncertainty in the interpretation. As each of us has a
different idea of the geologic picture of an area being worked, it is possible in areas with limited
well or seismic control to generate several fault surface interpretations using a single or similar
sets of fault data. Fault maps, like many other subsurface geological maps, tend to change with
time as new well and seismic data become available. Therefore, a fault surface interpretation is
never complete until the last well is drilled and all the seismic data to be shot have been shot and
interpreted.
(a) (b)
GEOLOGICALLY UNREASONABLE
Figure 7-11 (a) Fault surface base map showing the missing section and depth of the fault in each well. (b) Fault contours established in three areas of well
control. Each fault segment is part of the same fault. (c) Unrealistic fault surface interpretation results from connecting each fault segment with straight lines.
This is a mechanical approach to contouring. (d) Completed fault surface map using the interpretive form of contouring to reflect the expected geometry of
the fault surface. (e) Cross section A-A' passes directly through Wells No. 3, 6, and 2 and is laid out perpendicular to fault strike. Use of an interpretative
approach to contouring results in a gradual, rather than abrupt, change in fault dip with depth. (f) through (h) are computer-contoured maps based on the
same fault data as (d). They differ from each other because different gridding algorithms were used: (f) projected slope; (g) closest point; (h) point density.
Fault Surface Map Construction
-5000'
3
-6000'
-5740' -4890'
-6650' -5920'
-7000'
-6960'
-6995'
-8000' -7775' -6875'
-7370'
0'
-8630' -900 6 -8390'
-9240'
-9640' -9030'
'
-10,000
-9650'
-10,770'
-11,000' -10,505'
-11,060'
0 1000'
(e) (f)
-4000'
0' -5000'
3 00 0' 3
-6 -600 -5000' -6000'
-5740' -4890' -5740' -4890'
00
-7000'
'
-6960' -6960' '
-6995' -8000 -6995'
0'
-7775' -6875'
-6 -7
-7775' -6875'
-8000' -7370' -7370'
'
0'
-8630' -900 6 -8390' 6 -8390'
-8630'
-80
-9240' -9240'
-9640' -9030' -9460' -9030'
0'
-90
00 -9650' -9650'
-60000'
-10
0,
-7000'
-500 0'
-8000'
-9000'
-40
-10,000
00
-10,000'
-1
-11,000
-9000
2
,00
'
2
0
0'
'
-10,770' -8 -70 000 0' -10,770'
'
'
00 00 '
-10,505'
'
0' ' -10,505'
-11,0
-6 00 0'
00'
-5 00
-4
-11,060' -11,060'
0 1000'
0 1000'
(g) (h)
269
Figure 7-11 (continued)
270 Chap. 7 / Fault Maps
Figure 7-11b shows the contours established for the fault in the three areas of well control.
Based on the fault data, we assume that the three fault segments are parts of the same fault, so
the final step is to extend the contours into the area of no control and connect the fault segments.
There are two possible ways to do this. One method is to extend the contours from each segment
toward one another as straight lines until they intersect. This method, illustrated in Fig. 7-11c,
appears to be a more unreasonable or unlikely interpretation. The second and preferred method
is to use an interpretive form of contouring (Fig. 7-11d). With this method, some geologic license
is used in the interpretation to reflect the expected geometry of the fault surface in this tectonic
setting. We gradually change the strike direction of the fault connecting the adjacent segments
with a smooth curve.
Now that the fault map is complete, we can estimate the dip of the fault at any location. At
a depth of around 5000 ft subsea, the dip of the fault is 65 deg, decreasing to 55 deg at 8500 ft
subsea, and finally flattening to about 40 deg between –10,000 ft and –11,000 ft. This type fault
shape is common for growth faults (see the Growth Faults section in this chapter). The fault in
Fig. 7-11d is contoured as a listric (curvilinear concave-upward surface) growth fault; that is, a
fault whose dip decreases with depth, whereas the vertical separation or missing section increas-
es with depth.
Figure 7-11e shows a cross section (A-A' ) laid out in a northwest-southeast direction per-
pendicular to the strike of the fault. Three wells lie on the section with fault data for each well
posted. An interpretive method of contouring was used to contour the fault surface with depth.
This is the preferred method, which provides the most reasonable interpretation. Other methods
could have been used, including the mechanical contouring method, in which the dip rate is con-
stant between each pair of well control points but changes at each well. The equal-spaced con-
touring method also could have been used. Both methods provide a less reasonable interpretation
of the fault surface.
Three examples of the same fault data that were contoured by a computer-based mapping
program are shown in Fig. 7-11f through h. The Projected Slope gridding algorithm was used for
Fig. 7-11f. The result is a map similar to the hand-drawn map in Fig. 7-11d in that the fault sur-
face is listric but has a smoother curving lateral bend. A Least Squares gridding algorithm gen-
erated a comparable map except that the extrapolated 5000-ft contour did not conform well to the
trend of the other contours and the surface was not so smoothly listric. A Closest Point algorithm
was used to generate the fault surface map in Fig. 7-11g, resulting in a fault surface that is not
geologically reasonable considering the data, which indicate that the fault is a large growth fault.
The mapped surface is not listric and the contours are not credible at the limits of the data, both
shallow and deep. The Closest Point algorithm is best suited to a data set with more numerous
and more evenly distributed data points than in our example. Lastly, Fig. 7-11g is an extreme
example of a bogus fault surface map generated by an algorithm (Point Density) that is unsuit-
able to the data set. These three examples demonstrate how critical it is to select an appropriate
gridding algorithm (including suitable gridding parameters) for generation of a fault surface
interpretation or for integration with a structural horizon map.
Comparison to the hand-contoured map in Fig. 7-11d indicates that the Projected Slope algo-
rithm was the best choice among four algorithms to generate the most reasonable fault surface
interpretation. But that does not mean the Projected Slope algorithm is always the most suitable
for mapping a listric fault surface with a lateral bend. The best algorithm is dependent on the
number and distribution of data points, among other things. Beware of habitually choosing the
same gridding algorithm in computer-based mapping. The interpreter must be sufficiently famil-
iar with all the various algorithms and gridding parameters in order to choose the one most suit-
Types of Fault Patterns 271
able to the data set and the geologic surfaces in the area of study.
In developing a final fault surface map interpretation, keep in mind that a fault need not
remain constant in strike direction, dip, or vertical separation over its entire extent. Along the
strike, the vertical separation may increase, decrease, or remain constant, and the strike direction
may change. The vertical separation might increase with depth, decrease to zero up-section, or
even decrease with depth. A fault may die laterally, have its displacement transferred to other
faults or to folds, combine with other faults, or intersect with or terminate against another fault.
In areas of salt diapirs, a fault may terminate against salt, extend through it, or even be deformed
due to strata draping around the diapir or by salt movement. We again emphasize that a good
interpretation of a fault surface must have three-dimensional validity and comply with the tec-
tonic characteristics of the region being mapped, and you must use correct mapping techniques
in its construction.
Figures 7-12 and 7-13 are examples of completed fault surface maps. The fault map in Fig.
7-12 is that of the “S” Fault in the Indigo Bayou area, Iberville Parish, Louisiana. Although there
is some variation in the amount of missing section for this fault, for the most part it appears to be
a post-depositional (nongrowth) fault with a vertical separation ranging from 220 ft to 400 ft. The
fault exhibits little, if any, growth with depth. This fault surface map is contoured in accordance
with the guidelines and rules outlined in this section.
Contoured fault surfaces of selected normal faults in the Ivanhoe Field, U.K. North Sea, are
shown in Fig. 7-13 (Hooper et al. 1992). The map is an essential component of a computer-gen-
erated set of fault, structure, and isochore maps that were successfully combined in developing
more accurate structural and volumetric models of reservoir units than existed at the time. The
fault surface maps were used to improve accuracy in determining the intersections of mapped
horizons with the faults, and that in turn was the basis for more precise net pay maps. The use of
fault surface maps is essential in generating the most accurate maps possible. We present in
Chapters 8 and 9 our methodology for integrating fault surface maps with horizon structure maps
to generate accurate structure maps, and in Chapter 14 we describe their use in developing the
most precise net pay maps.
Figure 7-13 Computer-generated fault surface maps of selected normal faults in Ivanhoe Field, U.K.
North Sea. Dashed lines indicate that a fault surface extends beyond the contours shown on this
map. Wide lines are upthrown fault traces at one horizon. (Modified from Hooper et al. 1992;
AAPG©1992, reprinted by permission of the AAPG whose permission is required for further use.)
crestal grabens, as well as radial and peripheral faulting. Antithetic faults, also referred to as com-
pensating faults, are common in extensional areas.
In addition to single normal faults (structural and growth), there are three principal patterns
of normal fault intersections and terminations common in areas of extensional tectonics. These
patterns or systems, illustrated in Figs. 7-14, 7-17, and 7-20, are: (1) bifurcating, (2) compensat-
ing, and (3) intersecting. For each of the fault patterns discussed, a fault surface map, one or more
cross sections, and a block model are provided to explain and illustrate the pattern.
Bifurcating Fault Pattern. A bifurcating fault pattern or system results from two normal faults
that dip in the same general direction, as shown in Fig. 7-14. The strike direction of each fault is
274 Chap. 7 / Fault Maps
(a)
(b)
Figure 7-14
Types of Fault Patterns 275
such that the two faults merge laterally in the subsurface and continue on as one fault. The line
along which the two faults merge is called the line of bifurcation or intersection. The total ver-
tical separation of the fault across the line of bifurcation must be conserved. This means that the
vertical separation of the single fault, where only one fault exists, is equal to or nearly equal to
the sum of the vertical separations of the two faults, where two faults are present. The contoured
fault surface map in Fig. 7-14a shows two intersecting fault surfaces dipping in the same gener-
al direction. This interpretation was made using fault data from 11 wells, the contouring guide-
lines, and an understanding of the geologic setting.
Using Fig. 7-14, we review the fault system in detail and illustrate the specific characteris-
tics that classify this as a bifurcating fault pattern. Fault A is striking east-west and dipping 50
deg to the south. Fault A-1 is striking northeast-southwest with a dip of 50 deg to the southeast.
The two faults are dipping in the same general direction.
Fault A-1 merges with Fault A where the two faults intersect, as indicated by the dashed line
of bifurcation. It is the result of the intersection of contours of the same value on the two faults.
There are two faults present west of this line. Fault A has a vertical separation of 300 ft and the
vertical separation for Fault A-1 is 200 ft. East of the line of intersection only one fault (Fault A)
exists, with a vertical separation of 500 ft. These vertical separation values across the line of
bifurcation satisfy the conservation of vertical separation, also referred to as the additive proper-
ty of faults (see Chapter 8). Notice that the contours for Fault A are dashed west of the line of
(c)
Figure 7-14 (a) Bifurcating fault pattern resulting from two merging faults, dipping in the same gen-
eral direction. Line of bifurcation indicates where the two faults merge. (b) Cross section A-A′
bisects Faults A and A-1 in such a way that the two faults do not appear on the cross section as
merging faults, but instead appear as two parallel faults. (c) Cross section B-B′ is laid out almost
perpendicular to Fault A-1 and at an oblique angle to Fault A. In cross section, the faults appear to
merge with depth rather than to merge laterally.
276 Chap. 7 / Fault Maps
intersection, indicating that the contour values are deeper than those for Fault A-1. This is a good
contouring practice that helps reduce confusion on maps where more than one fault surface is
constructed on the same base.
Figure 7-14b and c are two cross sections with a different orientation to the two fault sur-
faces for each line of section. Remember, cross sections used in conjunction with maps provide
another viewing dimension that can be helpful in visualizing the geologic picture and solving
structural problems. The orientation of the section line, however, is very important. Chosen incor-
rectly, the line of section can be more confusing then informative.
In the two cross sections through the bifurcating fault pattern, the fault geometry appears dif-
ferent in each section. In cross section A-A′ (Fig. 7-14b), the fault pattern does not appear to be
bifurcating. Instead, the two faults appear as parallel faults. Is this real or an optical illusion as a
result of the line of section? In Fig. 7-14c showing the B-B′ cross section, Fault A-1 appears to
merge with Fault A with depth. Real or illusion? Although the two cross sections are geological-
ly and technically correct, they can pose problems for those unfamiliar with fundamental geo-
logic principles. When laying out a cross section, be sure to consider the purpose of the cross sec-
tion and your audience (Chapter 6).
The vertical separation values for the faults have been incorporated into each cross section
to represent correctly the offset of the 6000-ft and 7000-ft Sands by Faults A and A-1. The term
bed offset means that the horizon has been displaced by the fault and the displacement is defined
in terms of vertical separation. Earlier in this chapter, we showed that the missing section in a
wellbore as the direct result of a normal fault is the measurement of the displacement in terms of
vertical separation. Since this understanding is very important when preparing cross sections, we
detail the procedure for using vertical separation in the preparation of cross sections in Chap-
ter 6. Therefore, in the two cross sections in Fig. 7-14b and c, the offset for the beds is constructed
using the vertical separation from the wellbore fault data. We cannot overemphasize that well-
bore fault data are not throw; therefore, we cannot construct a cross section using the fault data
as throw.
(a)
(b)
Figure 7-16 Example of a bifurcating fault system. Note the conservation of vertical sep-
aration on either side of the line of bifurcation. (Published by permission of Texaco, USA.)
Figure 7-15 is a block diagram of a bifurcating fault pattern. A review of Fig. 7-15a and b
illustrates the geologic development of the fault pattern and individual fault blocks. Fault 1 devel-
ops first as the rocks fracture and Fault Block B moves downward, creating a fault with 300 ft of
vertical separation. Then Fault Block C moves and creates Fault 1A with a vertical separation of
200 ft. Because Fault 1A terminates at Fault 1, the surface of Fault 1 to the left of the intersec-
tion must accommodate the displacement of Fault Block C. Therefore, the vertical separation
increases to 500 ft on Fault 1 to the left of the intersection.
Figure 7-16 is an example of a bifurcating fault pattern. The fault system shown is that of
two faults dipping in the same general direction and merging laterally as indicated by the line of
bifurcation. The sum of the vertical separations in the area where two faults are present, east of
the line of bifurcation, is equal to or nearly equal to the vertical separation of the one fault west
of the intersection.
Fault L dipping to the north-northeast is contoured between –9500 ft and –14,000 ft. The
available well control east of the intersection of the two faults indicates that the missing section
for Fault L ranges from 65 ft to 105 ft. Fault K, which dips to the north, is contoured between
–10,500 ft and –14,000 ft. Based on the well control, the missing section for this fault appears to
be about 120 ft. The sum of the vertical separations for Faults K and L east of the line of bifur-
cation is ±185 ft to 225 ft; west of the bifurcation line, where both faults have laterally merged
into one, the vertical separation of the Fault L is 165 ft to 225 ft. The nearly equal values for the
vertical separation on both sides of the fault intersection show that the vertical separation across
the line of bifurcation has generally been conserved.
Compensating Fault Pattern. A compensating fault pattern or system consists of two normal
faults dipping in generally opposite directions toward one another (Fig. 7-17) with an acute angle
278 Chap. 7 / Fault Maps
(a)
Figure 7-17 (a) Compensating fault pattern resulting from two intersecting faults dipping in gen-
erally opposite directions. Line of termination of Fault B at Fault A indicates the intersection of the
two faults. (b) Cross section A-A′ illustrates the termination of Fault B at Fault A. Northwest of the
intersection, Fault A has a vertical separation of 300 ft, whereas southeast of the intersection
Fault A is 100 ft. Fault B has 200 ft of vertical separation. The vertical separation is conserved
across the line of termination.
between the strike directions of the two faults. At the line of intersection, one of the two faults
terminates against the other. Conservation of vertical separation is maintained on either side of
the line of termination, as demonstrated in the following discussion.
We can look in detail at the example fault surface map for the compensating fault pattern
shown in Fig. 7-17a. The fault data were obtained from 11 wells. Based on these fault data, the
general guidelines presented earlier, and an understanding of the expected fault surface geome-
try in this setting, two intersecting fault surfaces were contoured as shown.
The completed fault surface map illustrates the specific characteristics that classify this as a
compensating fault pattern. Notice that Fault A is striking east-west and dipping to the south with
a dip of 50 deg, and Fault B is striking northeast-southwest and dipping at 50 deg to the north-
west. The two faults are dipping in generally opposite directions and toward one another. Fault B
terminates against Fault A where the two fault surfaces intersect at equal subsea elevations, as
indicated by a dashed line referred to as the line of termination.
Southeast of this line of termination are two faults (Faults A and B). Fault A has a vertical
separation of 100 ft, and the vertical separation of Fault B is 200 ft. Northwest of the termination
line, only Fault A is present and it has a vertical separation of 300 ft. These displacement values
satisfy the conservation of vertical separation. Therefore, we say that Fault B is compensating
with respect to Fault A.
Types of Fault Patterns 279
(b)
Figure 7-17a is northwest-southeast stick cross section A-A' shown in plan view on the fault
contour map in Fig. 7-17a. The fault data from Wells No. 3, 2, 14, and 1, which lie directly on
the cross section, are posted on the section. Fault B terminates against Fault A at a depth of –7460
ft, which corresponds to the point on the fault map (Fig. 7-17a) where the termination line for
Fault B intersects the cross section. In the area where Fault A and Fault B are present, Fault A
has a vertical separation of 100 ft, and the vertical separation of Fault B is 200 ft. This is shown
in Well No. 1 by the fault cut point for Fault A of 100 ft at –8870 ft and 200 ft at –6090 ft for
Fault B. Northwest of the termination of Fault B, only Fault A is present with a vertical separa-
tion of 300 ft shown in the fault cuts in Wells No. 2, 3, and 14. In Well No. 14, for example, the
300-ft fault cut is at a depth of –7400 ft. The vertical separation values for the faults have been
incorporated into the cross section to correctly represent the offset of the 6000-ft and 7000-ft
Sands by Faults A and B.
Figure 7-18 is a block diagram of a compensating fault pattern. At times, you may hear the
following as an explanation for the fault displacements within a compensating system: “Think of
the system in this way. Northwest of the fault intersection, Fault 1 has a missing section of 300
ft. Since Fault 2 is 200 ft, it takes away 200 ft of displacement from Fault 1 southeast of the inter-
section of the two faults, leaving 100 ft of displacement for Fault 1.” This explanation may pro-
280 Chap. 7 / Fault Maps
vide you with a visual understanding of the missing section for each fault on both sides of the ter-
mination line, but technically it is incorrect and can lead to confusion. One fault cannot take dis-
placement away from another fault unless there is active inversion.
Looking at Fig. 7-19, think of the geologic development of the fault pattern and individual
fault blocks in the way they formed. Fault 1 develops as the rocks fracture and Fault Block B
moves downward, creating a fault with a vertical separation of 100 ft. Next, Fault Block C moves
and creates Fault 2 with a vertical separation of 200 ft. Fault 2 terminates at Fault 1, so Fault 1
to the left of the intersection must accommodate the displacement of Fault Block C. Therefore,
the vertical separation increases to 300 ft on Fault 1 to the left of the intersection. Some geolo-
gists refer to such movement as a “reactivation of the older fault surface by the younger fault.” If
we check for conservation of vertical separation, we see that the 200 ft for Fault 2 plus the initial
100 ft for Fault 1 to the left of the intersection equal the final 300 ft of vertical separation for Fault
1 to the left of the intersection. Comparing Figs. 7-15 and 7-19, can you see that the orientation
of the younger fault is the only fundamental difference between the bifurcating and compensat-
ing fault patterns? For simplicity, we use examples of fault systems in which one fault is implied
to be younger. It is also possible that the two faults are contemporaneous.
(a)
(b)
Intersecting Fault Pattern. So far, we have discussed two types of fault patterns in which one
fault merges or terminates against another fault at their intersection. Now we look at the inter-
secting fault pattern, which results from two faults (normal or reverse) dipping in such a manner
as to intersect in the subsurface; unlike the bifurcating system, in which the two faults merge, or
the compensating system, in which one fault terminates against the other, both faults continue
downward. The geometric relationship of intersecting faults is very difficult to visualize. Block
models can help illustrate this pattern in three dimensions, and 3D seismic data are at times a fan-
tastic data source from which to view these patterns.
Because of the complexity of this fault pattern, a correct interpretation is rarely achieved,
even in areas of adequate well and seismic control (Dickinson 1954). When considering the three
fault patterns discussed in this section, the intersecting fault pattern presents the most complexi-
ties and the solutions are not at all straightforward. With limited available data, a decision must
be made whether the intersecting faults formed contemporaneously (Horsfield 1980) or are of
two different ages. Without good seismic control, such as a three-dimensional survey, it is often
difficult, if not impossible, to determine if the faults formed contemporaneously or at different
times.
If the conclusion is that the faults are of two different ages, the next step is to determine
which fault formed first and which was second. Such conclusions affect the construction of the
fault surface maps, as well as completed structure maps. Even with today’s technology, rarely are
there sufficient data to make such decisions.
Because of the complexities and uncertainties surrounding this fault pattern, we recommend
that the intersecting faults be mapped as if there is no offset of one fault by the other. This
assumption does result in some error around the intersection of the faults and integrated structure
maps, but it does save time, and the actual interpretation may be impossible to determine from
data available. This compromise should usually introduce less error than an incorrect guess as to
the age of faulting. This subject is further discussed in Chapters 8 and 9.
For the intersecting fault example shown in Fig. 7-20a, we make one of two assumptions:
(1) the faults are contemporaneous, or (2) the relative age of the faults is unknown. With these
assumptions, we construct a fault map with both faults meeting at their intersection and continu-
ing downward, unaffected at and past their intersection. The fault maps were prepared using fault
data from the 11 wells shown on the map. Fault A is striking east-west and dipping to the south
at 50 deg, and Fault B is striking northeast-southwest with a dip of 50 deg to the northwest. The
intersection of the two faults is shown as a dashed line. Beneath their intersection, the two faults
continue downward with no change in vertical separation. These values are not affected by the
intersection of the faults as they are in the bifurcating and compensating fault systems.
Around the area of fault intersection, a chaotic zone may exist in which both the strata and
fault surfaces are disrupted. However, well control and seismic data rarely can identify such dis-
ruption and, therefore, the mapping around these intersections is inaccurate. This must be kept in
mind when mapping horizons affected by intersecting faults.
The cross section A-A′ shown in Fig. 7-20b illustrates the simplified (or compromised)
method of preparing the fault surface maps as if the two faults were contemporaneous. Although
neither fault surface is offset on the fault map, when the fault surface map is integrated with a
structure map, all four resultant fault traces may be offset. This is covered in detail in Chapter 8.
Figure 7-21 is a block diagram of an intersecting fault pattern resulting from two different
ages of faulting. Fault 1 developed first, followed by Fault 2. Notice in the figure that Fault 2 cuts
through Fault 1, displacing the older fault surface and causing a gap in this displaced surface.
282 Chap. 7 / Fault Maps
B
LT
U
FA
(a)
Figure 7-20 (a) Intersecting fault pattern resulting from two faults dipping in opposite directions. Unlike the
compensating fault pattern, both faults continue beneath the intersection. This fault map was prepared using
the simplified method of assuming neither fault surface is offset. (b) Cross section A-A' illustrates this inter-
secting fault pattern. Observe that the faults form a central “graben” block above the intersection and a central
“horst” block below the intersection.
Since both faults carry to depth, no change in vertical separation occurs for either fault below
their intersection.
An interpreted seismic line from a three-dimensional survey over an offshore Gulf of Mexico
field is shown in Fig. 7-22. It shows very clearly an intersecting fault pattern in which both faults
appear to pass through one another as if the faulting were contemporaneous, similar to the exam-
ple used in Fig. 7-20. Intersecting faults dipping in generally opposite directions, as shown in the
figure, are given the special name intersecting horst-graben faults. Looking at the figure, we
can see how this pattern gets its name. Above the fault intersection, the faults form a central
graben block; below the intersection is a horst block, thus the name horst-graben fault system.
In areas where there is significant seismic data (three-dimensional data), it is sometimes pos-
sible to determine the relative ages of the faults. If this is possible, fault surface maps can be con-
structed for both faults, showing displacement of the older fault by the younger one. Integration
of these fault maps with structure maps results in a more accurate representation of the fault inter-
section.
One final note on the different fault patterns: These patterns can be very complex, involving
numerous faults in a single area. Also, a fault need not remain as one pattern over its lateral
extent. In other words, a fault that is part of a compensating fault pattern in one area can be part
of an intersecting or bifurcating pattern in another (see Fig. 7-22).
Types of Fault Patterns 283
(b)
Combined Vertical Separation. The term combined vertical separation applies to the relation-
ship that results where two faults of different ages intersect. The zone of combined vertical sep-
aration applies to that segment of the intersecting (younger) fault which lies between the offset
surfaces of the displaced (older) fault. This zone is called the “zone of combined throw” by
Dickinson (1954); however, his use of the word throw is a substitution for vertical separation.
Figure 7-23a and b illustrate a sequence of faulting involving two normal faults that results
in a combined vertical separation. If a well penetrates the area of combined vertical separation,
only one of the two intersecting fault surfaces will be crossed (only one fault pick in the well),
but the interval shortening or missing section will be equal to the sum of the vertical separations
for both faults. The example in Fig. 7-23 shows two intersecting faults of different ages dipping
in generally opposite directions. The initial Fault 1 has 100 ft of vertical separation. The younger
Fault 2, which has 200 ft of vertical separation, has displaced Fault 1 in a manner similar to a
fault displacing a horizon. A review of the stratigraphic section in the area affected by both faults
shows a vertical shortening (or missing section in the well) of 300 ft, which is equal to the com-
bined vertical separation for Faults 1 and 2.
Figure 7-24 is a fault surface map for Fault J. This fault is the north-dipping component of
an intersecting fault system composed of two faults of different ages and dipping in opposite
directions. The north-dipping fault has a vertical separation of about 80 ft, and the vertical sep-
aration of the south-dipping Fault D is about 150 ft. Notice along the line of fault intersection
that the fault cuts in Wells No. 2 and 109 are unusually large (235 to 250 ft) compared to the
vertical separation of Faults D or J. These two larger fault cuts result from a combined vertical
separation.
When working in an extensional area of complex or intersecting faults where an unusually
large fault is present in one or more wells, keep the idea of combined vertical separation in mind.
An unusually large cut could be the result of a new, previously unrecognized fault, a bifurcating
or compensating fault pattern, a buried fault, or a combined vertical separation resulting from
intersecting faults of different ages.
The effects of intersecting normal and reverse faults are illustrated in Fig. 7-25a through d.
The upper part of each figure shows a structure contour map depicting the interruption of a dip-
ping horizon “O” by various combinations of intersecting normal and reverse faults. The cross
section in the lower part of each figure shows the structural effects of the intersecting faults on
interval “O-P,” which has a constant vertical thickness defined as “c.” Although the vertical sep-
aration for the fault in the zone of combined vertical separation is a function of many variables,
including horizon dip, fault dip, vertical separation for each fault, and the relative movements of
the individual faults, it is usually equal to the algebraic sum of the vertical separation of both
faults.
Where the intersecting (younger) fault is normal (Fig. 7-25a and b), the vertical separation
of the intersecting fault in the zone of combined vertical separation is equal to the algebraic sum
of individual vertical separations, that is, it is equal to (– b) plus (+ a). The figures illustrate that
where both faults are normal, the vertical separation across the intersected fault in the zone of
combined vertical separation will also be normal. Where the intersected fault is reverse, howev-
er, the vertical separation will be normal only if the intersecting fault has the greater vertical sep-
aration and will be reverse only if the intersected reverse fault has the greater vertical separation
(Dickinson 1954).
Figure 7-25c and d show the resulting fault geometry where the intersecting fault is reverse.
In these two cases, the vertical separation across the intersecting fault in the zone of combined
vertical separation is equal to the algebraic difference between the vertical separations of the
intersecting and intersected faults; that is, it equals (+ b) minus (+ a). The vertical separation
across the zone of combined vertical separation will be normal where the intersected fault is
reverse and has a vertical separation greater than that of the intersecting fault. Where the inter-
sected fault is reverse and has a vertical separation smaller than that of the intersecting fault, or
where it is normal, the vertical separation across the zone of combined vertical separation is
always reverse for a reverse intersecting fault. Unlike the geometry involving normal faults where
286 Chap. 7 / Fault Maps
100’
(a) 0 (b)
Figure 7-23 Schematic cross sections illustrating the zone of combined vertical separation (zone of combined fault
cut), which develops from two intersecting normal faults. (a) Fault 1 with vertical separation of 100 ft. (b) Younger Fault
2, with vertical separation of 200 ft, offsets Fault 1. Well No. 2 penetrates one fault and has 300 ft of missing section.
(Published by permission of D. Tearpock and J. Brewton.)
Figure 7-24 Wells No. 2 and 109 each have combined fault cuts as the result of the intersection of Faults
D and J, Golden Meadow Field, Lafourche Parish, Louisiana. (Published by permission of Texaco, USA.)
(a) (b)
287
faults. (b) Displacement across zone of combined vertical separation for a reverse fault intersected by a normal fault. (c) Displacement across zone of
combined vertical separation for intersecting reverse faults. (d) Displacement across zone of combined vertical separation for a normal fault intersected
by a reverse fault. (Dickinson 1954; AAPG©1954, reprinted by permission of the AAPG whose permission is required for further use.)
(c) (d)
288
Chap. 7 / Fault Maps
Figure 7-25 (continued)
Types of Fault Patterns 289
only one fault is seen in the zone of combined vertical separation, the geometry in Fig. 7-25c and
d results in three faults (i.e., a well in that zone will penetrate three faults).
Compressional Faulting
Reverse faulting is defined as motion along a dipping fault surface on which the hanging wall
block rises relative to the footwall block. End members of the reverse fault spectrum consist of
vertical and horizontal fault surfaces, as is the case with normal faults. An idealized reverse fault
is shown in Fig. 7-26. As in Fig. 7-2, for normal faults, this figure graphically defines the com-
mon fault components used in mapping. Depending upon relative amounts of fault and formation
dips, normal faults usually omit section, whereas reverse faults usually repeat section in electric
logs. The terms normal and reverse are in wide use to indicate these conditions. However, we
should recognize the origin of section omission and repetition and, in doing so, use the genetic
terms extensional and compressional faulting. The previous section of this chapter dealt with the
practical aspects of extensional faulting. In this section we discuss principles and practical
aspects of compressional faulting.
In compressional areas, the determination of fault data begins with well log correlation, just
as it does in extensional areas. For a reverse fault, the repeated section, recognized on an electric
log by correlation with surrounding well logs, is equal in value to the vertical separation
(Tearpock and Harris 1987). Figure 7-27 diagrammatically represents the determination of fault
data by the correlation of well logs. This east-west structural cross section perpendicular to the
strike of the fault surface shows two beds that have been cut by a reverse fault (F-1). The reverse
fault cuts Well No. 1 at –5300 ft and is dipping at an angle of 35 deg to the east, whereas the stra-
ta are dipping at 30 deg to the west. By correlation with Well No. 2, the repeated section in Well
No. 1 is determined to be equal to 100 ft and is shown to have completely repeated Bed A in Well
No. 1. As shown in the cross section, the throw of Fault F-1 is represented by the vertical line AC,
which is equal to 55 ft. Vertical separation, represented by the vertical line AE, is equal to the
repeated section, which is 100 ft. This correlation example demonstrates that the repeated section
obtained for the fault due to a reverse fault is equal to vertical separation.
Figure 7-28 is another reverse fault example used to illustrate the representation of a repeat-
ed section by log correlation. The figure shows a cross section with two wells, each of which has
intersected reverse Fault F-1. The fault and beds are dipping in the same direction. By log corre-
lation, we see that the repeated section in each well is equal to 100 ft, which is the vertical sepa-
ration (AE). Unlike the first example in Fig. 7-27 in which the throw was about one-half the size
of the vertical separation, here a measurement of throw reveals it to be over twice as large as the
vertical separation. We once again diagrammatically show that the throw of Fault F-1 cannot be
measured by log correlation and also that the throw is not equal to the repeated section.
Vertical separation is directly measurable as repeat section in well logs and on seismic sec-
tions. It has a fixed value at any given point on any particular fault, which is valid regardless of
the apparent attitude of the strata or fault. Depending upon the attitude of strata or a fault, throw
can be less than, equal to, or greater than the vertical separation. These geometric relationships
are vital to understanding the subsurface and its depiction in map view.
Horizontal shortening of section is one component of compressional faulting, whereas ver-
tical motion is another. We do not attempt to determine which is the main driving force in moun-
tain building. We can say that crustal shortening in the form of compressional faults and folds has
been documented by drilling in many regions of the world.
290 Chap. 7 / Fault Maps
Figure 7-26 Block diagram of Bed X displaced by a reverse fault, showing four different fault
components. The front panel is perpendicular to the strike of the fault. (Modified from
Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
WELL #1 WELL #2
100' / –5305' 100' / –5525'
N
IN
TIO
–5000'
EL
–5000'
EC
L
TS
#1
A
PEA
O 100'
RE
30 –5100'
–5100' DIP
FA E
U
DI LT A
P: F–
35 O 1 D
–5200' BE –5200'
A
–5500' –5500'
BASE OF
BED B REPEAT
–5600' B SECTION –5600'
D
BE 100'
100'
0 100'
AC = THROW = 55'
AE = VERTICAL SEPARATION = 100'
Figure 7-27 Diagrammatic cross section illustrating the geometric relationship between the
repeated section in a wellbore and the vertical separation of a reverse fault. (Published by
permission of D. Tearpock.)
Types of Fault Patterns 291
Thrust faults are dip-slip faults in which the hanging wall has moved up relative to the foot-
wall. Thrust faults generally dip less than 30 deg during active slip and commonly dip between
10 and 20 deg at their time of formation (Suppe 1985). Many thrust faults lie along bedding
planes over part of their length.
Thrust faults develop in two primary settings: (1) compressional plate boundaries and (2)
secondary faulting in response to folding. The foreland of fold-and-thrust belts have been the
most extensively studied areas of thrust faulting because of significant potential for both petrole-
um and coal reserves. There are two basic types of fold-and-thrust belts: (1) those dominated by
thrust faulting such as the southern Appalachians and the Cordilleran belt in western Canada and
the United States, and (2) fold-dominated belts such as the central Appalachians and Jura moun-
tains of Switzerland (Suppe 1985).
Individual thrust faults commonly cut up the stratigraphic section in a sequence of discrete
crosscutting segments called ramps and flats. The thrust fault separates the deformed hanging
wall structure from the substrate along a bedding plane zone called a decollement (from the
French word meaning “unglue”). Thrust and reverse faults, as with normal faults, can bifurcate
in the ramp areas with total displacement distributed among the many fault splays, and also
absorbed in folds. In many cases the hanging wall blocks of thrust faults are folded; this results
from bending of the fault blocks as they slip over nonplanar fault surfaces. These folds are called
fault bend folds (Suppe 1985). The best-known fault bend folds are those associated with the
stepping-up of thrust faults from lower to higher decollement surfaces.
Figure 7-28 Diagrammatic cross section illustrating the relationship between a repeat-
ed section in a wellbore and the vertical separation where the beds are dipping in the
same direction as the reverse fault. Compare the relationship in this figure to those
shown in Fig. 7-27.
292 Chap. 7 / Fault Maps
Three names are commonly applied to compressional faults: (1) high-angle thrust faults,
usually referred to as reverse faults (steeply dipping), (2) low-angle thrust faults (moderately dip-
ping, sometimes defined as less than 45 deg of dip), and (3) overthrust faults (very shallow dip,
usually with long horizontal transport).
How do we map the surface of a reverse fault? Most of the fault surface map and cross-sec-
tion techniques already described for extensional faulting apply to compressional fault surface
mapping. For the most part, the differences lie in the complexities of the compressional fault pat-
terns and the limit of good quality well and seismic data. Probably more than in any other tec-
tonic setting, we must have a thorough understanding of the principles of structural geology and
the structural styles at play, and we must be able to think and visualize in three dimensions.
Intersecting Compressional Faults. A compressional fault pattern of two reverse faults dip-
ping in the same general direction is shown in Fig. 7-30a. The data from 12 wells, the general
Figure 7-29 Fault surface map for a reverse fault with a vertical separation (repeated sec-
tion) of 300 ft.
Types of Fault Patterns 293
guidelines presented earlier, and an understanding of the compressional tectonic setting were
used to construct the fault surface map interpretation for the two reverse faults A and B. Fault A
is striking east-west and dipping to the south at 35 deg, and Fault B is striking northeast-south-
west and dipping 35 deg to the southeast. The two faults merge laterally as indicated by the
dashed line of intersection. We can see by reviewing the fault surface map that the vertical sepa-
ration is conserved across the line of intersection; this means that the combined vertical separa-
tions for Fault A (125 ft) and Fault B (150 ft) west of this line equal the vertical separation of
Fault A (275 ft) east of the intersection.
The change in displacement for these reverse faults across the intersection line is the result
of their intersection. Do not mistake the change in displacement in this example with changes in
fault displacement resulting from its transfer from one structure to another. In other words, this
example is not related to what Dahlstrom (1969) refers to as the compensatory mechanism for
reverse faults (thrusts) wherein fault displacement that is diminishing in one fault is replaced by
increasing displacement on an echelon fault. Those displacement changes occur in a transfer
(a)
Figure 7-30 (a) Fault surface map of two intersecting reverse faults. (b) Cross section A-A'
crosses the two faults at an orientation that makes the two intersecting reverse faults appear
to be parallel.
294 Chap. 7 / Fault Maps
(b)
zone and result from all the associated faults being rooted in a common sole fault. This type of
faulting and change in displacement is discussed in Chapter 10.
Figure 7-30b shows cross section A-A' . The displacement for each fault has been incorpo-
rated into the offset of each stratum shown on the cross section. Remember, just as in the exten-
sional faulting examples, the displacement of the strata is described in terms of vertical separa-
tion. Therefore, the footwall and hanging wall intersections for each stratum with Faults A and B
are constructed with the vertical separation technique used to displace a stratum correctly from
one fault block to another in cross section. This is the same technique used to construct the cross
sections for the extensional and compressional fault patterns, which was introduced in Chapter 6.
Ramp and Flat Thrust Faults. Ramp and flat fault surfaces develop such that the fault surface
moves subparallel to bedding in incompetent strata and ramps to higher structural levels in com-
petent rock. If the ramp intersects another incompetent rock layer, then the ramp may transform
again into another flat. This geometry is illustrated in Fig. 7-31. Figure 7-31a is a fault surface
map of an idealized ramp and flat thrust fault. As with all fault surface maps, the interpretive
method of contouring is recommended. This gives you the geologic license to construct the fault
surface map to reflect the expected geometry in the area of study.
Figure 7-31b is cross section A-A' , which shows the fault surface and the disrupted beds.
Notice how the strata in the hanging wall are thrust up and over the same strata in the footwall.
The geometry of these faults is detailed in Chapter 10.
Types of Fault Patterns 295
(a)
(b)
Figure 7-31 (a) Fault surface map for an idealized ramp and flat thrust fault.
(b) Diagrammatic cross section, parallel to the transport direction of the thrust fault
shown in Fig. 7-31a.
296 Chap. 7 / Fault Maps
Types of Fault Patterns
297
Figure 7-32 Fault surface map of the Absaroka thrust fault in the Fossil Basin of Utah and
Wyoming. Note the changes in fault dip from east to west. (From Lamerson 1982. Published by
permission of the Rocky Mountain Association of Geologists.)
298 Chap. 7 / Fault Maps
A fault surface map of the Absaroka thrust fault in the Fossil Basin of Utah and Wyoming is
shown in Fig. 7-32. Twenty oil and gas fields have been found in the hanging wall block of the
Absaroka thrust fault in the southern Fossil Basin. The fault surface map in general reflects a
steep ramp to the west from about –20,000 ft to –12,000 ft, a flattening or gentle rise in the cen-
tral portion of the fault map between –12,000 ft and –5000 ft, and then another steepening ramp
to the east (Lamerson 1982).
A portion of the Hogsback ramp and flat thrust fault is shown in cross-sectional view in Fig.
7-33. This cross section clearly illustrates the features of the ramp and flat portions of this type
of thrust fault. Observe that the flat parallels the Cambrian rocks, which serve as the zone of
weakness for the basal detachment zone. As the Hogsback thrust fault cuts upsection across
Paleozoic and Mesozoic rocks in the footwall, it creates a ramp with west dip of 20 to 30 deg in
the hanging wall rocks of the thrust.
Fault surface mapping is not commonly done in compressional areas, although there is no
reason why it should not be used and, indeed, it is often required to aid the understanding of com-
plex thrust fault geometry. Hydrocarbon traps are not common along thrust faults, but instead the
anticlines that form in the hanging wall blocks act as the primary trapping mechanism. These
structural anticlines are often sufficient for trap delineation with little, if any, fault interaction. We
strongly believe, however, that the mapping of faults is an integral part of any geologic study
regardless of the tectonic setting and can significantly add to the understanding of the geology.
Figure 7-34a is the fault surface map for the Aguaclara ramp and flat thrust fault from the
southeastern thrust front of the Eastern Condillera of Colombia, S.A. (Rowan and Linares 2000).
This fault surface interpretation and map were generated from both well log and seismic data.
The fault surface in this case is mapped in time (msec) rather than depth. Observe the classical
ramp, flat, ramp geometry of the fault surface. One of the seismic sections used in this interpre-
tation is shown in Fig. 7-34b. Observe how the hanging wall beds have been folded due to the
effects of thrust faulting.
1. The data being used are of reasonable quality and have been correctly processed up to and
including migration of the data.
2. The two-way time-to-depth conversion is known (i.e., a reliable velocity function exists)
and does not vary significantly across the area being mapped.
3. The lines have been interpreted correctly, with all faults tied at line intersections. In areas
of complex structures, this may not be possible. Where 2D data are used, steep dips pose
difficulties due to the migration mis-tie between strike and dip lines. In these cases you
may have to rely mostly on the dip-oriented lines. This sort of problem makes fault surface
mapping a very helpful tool in identifying faults on successive dip lines (Chapters 5
and 9).
Fault Data Determined from Seismic Information
Figure 7-33 Cross section illustrating the geometry of the Hogsback ramp and
flat thrust fault. (Lamerson 1982. Published by permission of the Rocky Mountain
299
Association of Geologists.)
300 Chap. 7 / Fault Maps
ramp
00
50
00
46
0
420
00
38
3400
flat
00
30
00
c)
00
26
m se
22
e(
ramp
ti m
in
s
ur
to
0
on
180
C
ce 100
0
u rf a
u l tS
Fa
Seismic
Line
ramp 00
30 N
00
26
10 km
00
22
(a)
Figure 7-34 (a) Fault surface map for the Aguaclara ramp and flat thrust fault, Colombia. Contour interval is 400 msec,
relative to an arbitrary datum near the surface. (b) Interpreted seismic profile; location shown in (a). Solid lines are axial
surfaces. (Modified after Rowan and Linares 2000; AAPG©2000, reprinted by permission of the AAPG whose permis-
sion is required for further use.)
Fault Data Determined from Seismic Information 301
(b)
Figure 7-34 (continued)
Figure 7-35 Illustration of the usefulness of fault-tying in propagating a fault surface over a
series of seismic lines. The lower diagram is a map of the fault pattern at the level of Bed A.
(Prepared by C. Harmon. From Tearpock and Harris 1987. Published by permission of
Tenneco Oil Company.)
If these assumptions are met, seismic data can provide three very useful types of fault infor-
mation that aid in the integration of fault data from wells to generate a fault surface interpreta-
tion and map. First, seismic data offers a method of establishing a fault identification (ID).
Second, it provides the location of a fault surface in space. Third, the amount of vertical separa-
tion for a fault can be measured.
302 Chap. 7 / Fault Maps
Figure 7-36 Seismic section shows the effect of velocity function uncertainty on the location
of a directional well. (Prepared by C. Harmon. Modified from Tearpock and Harris 1987.
Published by permission of Tenneco Oil Company.)
The tying process provides us with a method of establishing a fault ID. In short, by tying a
fault on a series of lines, we can be sure that the fault seen on various lines is formed by the same
fault surface. A fault interpretation that does not tie between lines cannot be correct. Merely tying
the faults, however, does not ensure that the fault interpretation is valid, since any nonvertical
surface can be tied between a series of lines. To establish validity, a fault surface map must be
constructed to confirm that both the fault dip and the changes in strike of fault contours are geo-
logically reasonable.
Seismic data offer a method of visualizing a fault interpretation through a series of lines. By
tying the fault traces from these seismic sections, we can be assured that a given fault seen on one
section is the same fault seen on other profiles. A schematic illustration of this is shown in Fig.
7-35.
Fault Data Determined from Seismic Information 303
Seismic data give us a location of the fault surface in space, and unlike well data, data from
seismic sections are over a series of points along the profile of the line. This enables us to map
fault surfaces over a greater area, and in areas where no well control exists.
It is important to remember that accuracy of points plotted from seismic data is dependent
upon the accuracy (appropriateness) of the velocity function used to convert seismic time to a
depth value. Figure 7-36 illustrates what influence a velocity function can have on dipping sur-
faces. The figure shows two positions for a proposed directional well plotted on a seismic line
based on two different velocity functions. Both functions are possibly valid for this prospect. As
the illustration shows, the use of the slow velocity function places the directional well in the
upthrown fault block; however, the fast function places the well in the downthrown fault block.
The upthrown block is the target block to test horizons 1 and 2. The use of the wrong velocity
would cause the exploratory well to be drilled in the wrong fault block. In this particular situa-
tion, since the correct velocity was not known, a straight hole design was chosen rather than risk
missing the prospective fault block with a directional well. The same problem can apply to fault
interpretations. In some cases, there may be some uncertainty as to the two-way seismic time cor-
responding to a particular depth. If the position of a fault is especially critical, error bars can be
placed around a posted point that will graphically illustrate spatial uncertainty for a particular
datum.
The third type of information seismic data can provide is a measurement of the vertical sep-
aration of a fault. There is often disagreement between geologists and geophysicists with regard
to the amount of missing or repeated section for a fault estimated from well log correlation as
compared to that estimated from a seismic section. In many cases, the difference between what
the geophysicist is calling displacement and what the geologist is calling displacement is the
result of a misunderstanding of fault component terminology. If we assume the vertical resolu-
tion limitation of seismic data is known, then the answer to this discrepancy is that two different
values for displacement are being measured. As discussed earlier in this chapter, the amount of
missing section determined from correlation of an electric log of a well is not equal to throw, but
rather, equal to vertical separation (unless, of course, the strata in the area have zero dip or the
fault is vertical). Then the values for throw and vertical separation are the same. What is com-
monly measured on a seismic line for fault displacement is not vertical separation but throw and,
more commonly, apparent throw, which is a measurement of fault displacement dependent upon
the orientation of the seismic line in relation to both the strata and fault dip. The only case where
we can actually measure the true throw of a fault, at a particular depth, is where the seismic line
is oriented perpendicular to the strike of the fault surface (not perpendicular to the fault trace as
seen on a structural map). This means that if several seismic lines, oriented at different angles to
the strike of a fault surface, are used to estimate the throw of that fault, each line will yield a dif-
ferent value for the apparent throw for the same fault. Even if it were possible to shoot all seis-
mic lines perpendicular to the strike of each fault surface being mapped, the throw of the fault is
not the fault component used in the construction of fault and integrated structure maps, and it is
not the displacement value determined from well logs. The added work and expense does not
seem justified.
Fault displacement problems from seismic data can be easily reconciled by estimating the
vertical separation of a fault from seismic lines, instead of throw or apparent throw. If a seismic
line has good correlations across a fault, it is very simple to measure the vertical separation for
that fault. Remember, the amount of missing section for a fault in a well log is the measurement
of the fault’s vertical separation. In order not to confuse components in our fault displacement
304 Chap. 7 / Fault Maps
Figure 7-37 Seismic line offshore Gulf of Mexico. The figure inserts illustrate the correct method for
estimating the vertical separation of a fault from a seismic line where the events are dipping (Horizon
R) and where the events are horizontal (Horizon G). In order to structure contour the correct dis-
placement, based on seismic data, across a fault, the vertical separation must be estimated. Throw
is not used in subsurface mapping. (Modified from Tearpock and Harris 1987. Published by permis-
sion of TGS Offshore Geophysical Company and Tenneco Oil Company.)
measurements and constructed maps, the vertical separation for a fault is a required measurement
on a seismic section. These depth-converted vertical separation values can then be used to tie the
fault data from wells and aid in the construction of the fault and structure maps.
Figure 7-37 illustrates the proper method for measuring the vertical separation from a seis-
mic line. The line is an east-west line shot in the central Gulf of Mexico Basin by TGS Offshore
Geophysical Company. It is an excellent line to use to define the procedure for measuring verti-
cal separation because the dip of the structure changes from east to west. At the eastern end of
the line, the structure is essentially flat, with zero dip; to the west, the dip increases. Therefore,
we can show the procedure for measuring the vertical separation in flat-lying beds as well as
those with significant dip. We also show the comparison of the values for vertical separation and
apparent throw to reinforce the understanding that these fault components are not the same and
that the use of throw in place of vertical separation can cause significant mapping errors.
We begin with the displacement for Fault A at Horizon R. The procedure is more easily fol-
lowed by using the insert blowup in the upper left corner of the figure. The R Horizon is dipping
at about 16 deg (depth-corrected) to the west. To measure the displacement of the fault in terms
of vertical separation, project the dip of the horizon from the upthrown fault block through the
fault, as if the fault were not there, into the downthrown block until the projection is vertically
over the intersection of the R Horizon with the fault in the downthrown block. Then convert time
Fault Data Determined from Seismic Information 305
to depth. The vertical difference in depth from the projection, at –10,900 ft, to the downthrown
intersection of the horizon and fault, at –11,760 ft, is 860 ft. This is the measurement of the ver-
tical separation of the fault at this horizon. If a well were drilled through the fault at shotpoint
7865, so that the R Horizon is faulted out of the well, the missing section in the well log would
be approximately 860 ft.
The measurement of the apparent throw of the fault at this horizon is the vertical difference
in depth from the intersection of Horizon R with the fault in the upthrown fault block (–11,290
ft) to the intersection of the horizon with the fault in the downthrown block (–11,760 ft), and this
is 470 feet. The difference between the apparent throw and the vertical separation of the fault at
the R Horizon is a significant 390 ft.
Turning to the right side of the figure, we conduct the same procedure and measure the dis-
placement of Horizon G by Fault B, in an area in which the structural bed dip is essentially zero
deg. Refer to the insert blowup in the lower right corner of the figure. First, measure the vertical
separation by projecting the dip of the horizon in the upthrown block through the fault into the
downthrown block until the projection is vertically directly over the intersection of Horizon G
with the fault in the downthrown block. The vertical distance between the depth of Horizon G at
the fault intersection in the upthrown block at –8530 ft, and the depth of the horizon/fault inter-
section in the downthrown block, of –9120 ft, is 590 ft.
The apparent throw, measured as the vertical distance from the intersection of the horizon
with the fault in the upthrown block to the intersection of the same horizon and fault in the down-
thrown block, is 590 ft. This is the same value obtained for the vertical separation. Since the stra-
ta dip at this location is zero deg, we should expect these values to be the same, based on the def-
initions and discussion presented so far in this chapter.
If necessary, go back and review these procedures again until you thoroughly understand
how to measure the vertical separation and clearly see how it differs from apparent throw. It is
essential to remember that whether the fault data are derived from well logs or seismic lines, the
same fault component for displacement must be used in mapping faults and structures, and that
component is vertical separation. Of course, vertical separation changes laterally along a fault
and varies with depth, so you need to determine it on an appropriate number of seismic sections,
at depths suitable for the mappable horizons.
Figure 7-38 illustrates the accuracy of the technique outlined above in reconciling well log
data with seismic data. The seismic line is a Tenneco Oil Company line in South Marsh Island
(SMI), offshore Gulf of Mexico. The seismic line intersects the Signal Well No. 1 in SMI Block
67. A fault in the Signal Well produces a missing section equal to 720 ft at a log depth of 10,940
ft, and it faults out Horizon A. Based on this seismic line, the apparent throw of the fault at
Horizon A, marked on the line, is 505 ft, and the vertical separation is 695 ft. Notice that the
measurement for the vertical separation agrees very closely with the missing section of 720 ft
obtained in the well. The estimate of the apparent throw, however, differs by nearly 200 ft from
the estimate of fault displacement by well log correlation. By this example, we again illustrate
the importance of measuring the same fault component to obtain the value for fault displacement,
which will later be used for fault and structure mapping, whether that value is derived from well
log or seismic data.
Seismic and Well Log Data Integration – Fault Surface Map Construction
Since the measurement of fault displacement data can be made from seismic and well log data,
both can be used in the construction of fault surface maps. Figure 7-39 illustrates the integration
306
Chap. 7 / Fault Maps
Figure 7-38 Seismic section illustrates the accuracy with which the vertical separation can be estimated, as compared to fault data
from a well log directly on the section line. Missing section is 720 ft in the Signal No. 1 Well by correlation with surrounding wells.
This compares favorably with the vertical separation of 695 ft estimated from the seismic section. The apparent throw of the fault is
only 505 ft. (From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
Fault Data Determined from Seismic Information 307
Figure 7-39 Fault surface map prepared using data from well logs and seismic
sections. (Modified from Tearpock and Harris 1987. Published by permission of
Tenneco Oil Company.)
of fault data from two well logs and two seismic lines to construct a fault surface map. The fault
data obtained from the well logs are posted next to the appropriate well, and the fault data from
the two seismic lines are posted next to the shotpoints at which the data were obtained. Unlike
fault data from a well log, which provides fault information from a single depth at a specific loca-
tion, a seismic line presents continuous fault data over a series of shotpoints along the profile of
the line. This permits the fault surface to be mapped over a greater area than would be possible
with well log data alone. The seismic also provides a tie to each well to confirm whether the fault
cut observed in the well is the same fault surface seen on the seismic line.
Fault-Displacement Mapping. A review of each fault surface map presented in this text reveals
that the vertical separation and subsea depth for each well log or seismic fault pick is posted on
the base maps (e.g., 310 ft/–6621 ft). So far we have used only the subsea depth data to prepare
fault surface contour maps. What about the posted vertical separation data? Can these data be
used for mapping and, if so, what is their importance? The vertical separation of the fault can and
should be contoured where sufficient data are available. Such a map, called a vertical separation
map, or “throw map” by some, provides additional information on the growth history of the fault
and additional data for the construction of integrated structure maps (see Chapter 8).
In Fig. 7-40a, vertical separation contours are superimposed on the fault surface contours.
This vertical separation is contoured in a dashed pattern with a contour interval of 200 ft. Notice
that the vertical separation contours are perpendicular to the fault surface contours at depths
greater than –5000 ft. This type pattern indicates that the fault is post-depositional and is there-
fore termed a structural (nongrowth) fault. In other words, the fault is not a syndepositional
growth fault; it does not show an increase in displacement with depth. However, the deviation in
trend of the vertical separation contours, at a depth from about –3000 ft to –5000 ft, shows an
increase in vertical separation with depth and indicates the time during which the fault was active.
308 Chap. 7 / Fault Maps
(a)
(b)
Figure 7-40 (a) Vertical separation contours added to fault surface contour map pro-
vide additional information that is helpful during fault and structure map integration
(see Chapter 8). (Modified from Tearpock and Harris 1987. Published by permission
of Tenneco Oil Company.) (b) Seismic lines A and B shown in map view in Fig. 7-40a.
(From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
This information has application in evaluating a fault trap because it graphically illustrates
the depth of the geologic interval affected by post-depositional faulting. A late trap formed by a
post-depositional fault can be evaluated as a prospect as long as migration of hydrocarbons
occurred during or after the initiation of fault movement. Displacement mapping for growth
faults is also very important and is covered later in this chapter.
Seismic Pitfalls
As we have already mentioned, this text is not intended to instruct in all areas of seismic inter-
pretation, nor is there any intent to discuss in detail the pitfalls of seismic interpretation. However,
Fault Data Determined from Seismic Information 309
several seismic pitfalls can be avoided if fault surface and vertical separation mapping are under-
taken. We address these subjects to emphasize the importance of constructing fault surface maps
as part of any subsurface geologic study.
In extensional basins, do faults die with depth? The answer to this question is obviously
“Yes,” but, when a fault is interpreted as dying with depth, there has to be a geologic explanation
for this occurrence (Chapter 10). We have seen faults interpreted on a single seismic section or a
limited number of sections as dying with depth, where they actually do not. A fault may appear
to die with depth on a seismic profile, but this does not mean that it actually dies with depth in
the subsurface. What it does indicate is that an interpretation based on one line or a few selected
lines can be suspect, requiring additional subsurface work that may include evaluating addition-
al seismic or well control, if it is available.
Seismic line B in Fig. 7-40b illustrates something commonly seen on seismic – a fault
“dying with depth.” Is it real or just an illusion? Line A shows the same fault; however, on this
line the fault continues with depth rather than appearing to die at about 2.1 sec as it does on line
B. The fault shown on these two seismic lines is the fault contoured in Fig. 7-40a. Referring to
this figure, we see that the fault does not die with depth. The fault does, however, die laterally to
the northwest. Notice that Line B is oriented north-south in the general direction in which the
fault is laterally dying. At shotpoint 50 on line B, the vertical separation is 175 ft, it is 45 ft at
shotpoint 70, and 0 ft (fault dies) at about shotpoint 77. Therefore, on line B the fault appears to
die with depth. The key word is appears. This effect is simply a function of the orientation of the
line in relation to the fault surface of this laterally dying fault.
In this particular case, the dilemma of incorrectly interpreting a fault dying with depth can be
avoided if a detailed evaluation of the fault surface is conducted. First, Lines A and B can be tied
at shotpoint 20. This fault tie confirms that the fault continues with depth. Second, the construc-
tion of a fault surface map, incorporating the well and seismic data, shows that the fault continues
with depth. Third, as mentioned earlier in this section, whenever possible contour the vertical sep-
aration on the fault map as part of the geologic evaluation of that fault. In this case, the vertical
separation contoured in Fig. 7-40a, using the seismic and well log data, clearly shows that the fault
continues with depth but dies laterally to the northwest. These three steps illustrate how correct
and detailed fault surface mapping can present a better and more complete geologic interpretation
of a fault and prevent incorrect subsurface interpretations that could later prove to be costly.
Figure 7-41 is a seismic line that illustrates an authentic example of a pitfall that could have
been avoided with fault surface and vertical separation mapping. The profile shows two faults,
both coincidentally lining up to look as if one fault trace could be drawn; however, there
actually are two faults. This line was a regional line used to illustrate a prospect. As shown in
Fig. 7-41b, a fault correlation was forced where no fault exists (dashed line). This is significant,
because by forcing a one-fault interpretation, a footwall trap is inferred between 2.5 and 2.9 sec
that really does not exist. Figure 7-42a is a structure map interpreted from the forced correlation
on that line, as well as the interpretation on three other lines of what the geophysicist considered
to be the same fault. The result is a curved fault that forms a trap. Notice that the structure map
identifies 500 ft of closure in the footwall fault block. Figure 7-42b is the fault surface map
implied from the structure map. However, the prospect was generated without this fault surface
map being made and evaluated. The prospect was later presented with the intent of placing a
future bid on the offshore sale block in which it is located.
The prospect was reviewed using the seismic line in Fig. 7-41 and the additional lines in the
surrounding area. First, based on the structure map in Fig. 7-42a, the implied fault surface map
310 Chap. 7 / Fault Maps
(a)
Figure 7-41 (a) The orientation of this seismic line is such that Faults A and B coincidentally line up and appear
as if a one-fault trace could be drawn. (b) A one-fault interpretation forced through continuous reflectors (see
dashed line). (From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
Fault Data Determined from Seismic Information 311
(b)
in Fig. 7-42b was prepared. Second, using all the available lines, the fault and structural inter-
pretations in Fig. 7-43a and b were made, which show two faults. The shallower fault is inter-
preted as a fault that is dying laterally away from an anticline to the west. It appears to die with
depth on the seismic line (Fig. 7-41a), but actually the fault surface is simply migrating out of the
plane of the line. The deeper fault is a buried growth fault that is also dying laterally, but in a gen-
erally opposite direction to that of the shallower fault. Seismic line 120A just happens to be in
the correct orientation to make the two fault traces appear to line up on the profile. Figure 7-43
illustrates the correct fault surface and structural interpretation based on all the available data.
Notice that based on this new and correct interpretation, there is no fault trap and therefore no
prospect.
The generation of this nonexistent prospect would not have occurred had the procedures out-
lined in this chapter been followed. First, a fault surface map should always be prepared. Had a
fault surface map been prepared based on the one-fault interpretation and map, the sharp convex
bend (in the downthrown direction) of the fault surface (Fig. 7-42b) should have been suspect and
(a)
further review of the fault undertaken. Second, the construction of a vertical separation map
would have shown a fault losing displacement with depth, then abruptly gaining displacement
again, a very questionable and unlikely situation.
No prospect should be based on a single seismic line or even a few lines where more data
are available. Too many illusions of geometry can occur on any given line. A gross under-use of
costly 3D seismic data, frequently done to “save time,” is to use every fifth or tenth line to gen-
erate an interpretation and prospect. This means that only 10 to 20 percent of the seismic data are
used. Companies spend millions of dollars to acquire and process 3D seismic surveys. The pur-
pose is to provide a dense grid of data to improve the interpretation process. The use of every fifth
or tenth line to generate an interpretation is at times a costly, unjustified shortcut that should not
be accepted.
In Chapter 1, we stated that the only certainty in a subsurface interpretation, prepared from
seismic and well log data, is that it is most likely incorrect. The best geoscientist is the one who
comes up with the most accurate and geologically reasonable subsurface interpretation.
(a)
Figure 7-43 (a) Fault surface map for Faults A and B. (b)
Completed structure map constructed by integration of fault
and structure maps shows no hydrocarbon trap. (From
Tearpock and Harris 1987. Published by permission of
Tenneco Oil Company.)
(b)
314 Chap. 7 / Fault Maps
Remember the two significant causes of errors in geologic interpretations and maps: the failure
to use all the data available and the incorrect application of the data. If we fail to make fault sur-
face and vertical separation maps, where the data are available, how can we expect our interpre-
tation and resultant maps to be accurate, geologically reasonable, and representative of the best
interpretation of the subsurface?
GROWTH FAULTS
One of the principal petroleum-related settings for normal faults is that of major large-scale grav-
ity slide structures. The normal faults in this category decrease in dip with depth to low-angle
faults, commonly becoming bedding-plane faults. These faults slide along a decollement and are
called detachment faults (Chapter 11). Such fault systems are well known in areas such as the
U.S. Gulf of Mexico, Brunei, and the Niger Delta. These faults are called listric growth normal
faults, based on the Greek word listron, or shovel, because of their curved appearance in profile.
The major listric fault on a structure is commonly referred to as the master fault, if it has the
largest relative displacement. Numerous synthetic and antithetic (secondary) faults are often
associated with the master fault.
These listric growth faults are syndepositional, commonly exhibiting a significant increase
in stratigraphic thickening in the downthrown (hanging wall) block and an increase in displace-
ment with depth. This relationship indicates that there was movement along the fault surface
while the sediments were being deposited. Growth faults also are called syndepositional and con-
temporaneous faults because fault movement was occurring while the adjacent sediments were
being deposited.
Growth faults are commonly associated with hanging wall anticlines, called rollovers, which
develop as a result of the hanging wall block bending as it conforms to the curved fault surface
(Suppe 1985). These rollover anticlines are one of the most important hydrocarbon traps associ-
ated with listric growth faults.
The Restored Top Method. The thickness of the stratigraphic section in the footwall
(upthrown) block of a growth fault is thinner than the equivalent stratigraphic section in the hang-
ing wall (downthrown) block. Therefore, if a faulted well is correlated with a well in the
upthrown block, the missing section due to the growth fault will be estimated to be smaller than
if the correlation is made with a well in the downthrown block. So the procedure for estimating
the vertical separation for a growth fault is not as straightforward as that for nongrowth faults. In
most cases, at least two correlation well logs are needed: one located in the upthrown fault block
and the other in the downthrown block. With the available electric well logs, the vertical separa-
tion at any horizon faulted out of a well can be estimated by the Restored Top Method, which is
described in detail in Chapter 4.
1. Choose a horizon, faulted out of the subject well, for which you wish to estimate the
vertical separation.
Growth Faults 315
2. The log of a well in the upthrown block is then correlated with the deeper log section (the
interval below the fault) in the faulted well in order to obtain an estimate of the upthrown
restored top for the chosen horizon.
3. The log of a well in the downthrown block is correlated with the upper log section (the
interval above the fault) in the faulted well in order to estimate the downthrown restored
top for the horizon.
4. The estimate of vertical separation for the growth fault at the horizon is the difference in
measured depths of the upthrown and downthrown restored tops for the horizon in the
faulted well.
Using the cross section in Fig. 7-44, we illustrate the use of the restored top method. Fault
1 is a growth fault downthrown to the east. In Well No. 1, Beds A, B, and most of C are faulted
out. Bed B is a horizon of interest to be mapped; therefore, we want to estimate the vertical sep-
aration for the growth fault at the level of Bed B. Three wells are on the cross section: Well No.
2 in the upthrown fault block, Well No. 3 in the downthrown fault block, and Well No. 1, which
has the horizon of interest faulted out. If necessary, go back to Chapter 4 and review the section
on estimating restored tops before continuing this section.
First, estimate the missing section due to the fault in Well No. 1 using the other two wells
(Fig. 7-44). By correlation with Well No. 2 in the upthrown fault block, the missing section at the
fault in Well No 1 is 190 ft, but by correlation with Well No. 3 in the downthrown block, the miss-
ing section at the fault is 315 ft. Since these two estimates for the missing section are different,
which one should be used as vertical separation in mapping Bed B? The answer is neither esti-
mate. The 190-ft value is too small since it is based on correlation with a well in the stratigraph-
ically thin upthrown block. The 315-ft value is too large because it was obtained by correlation
with a well in the stratigraphically thick downthrown block.
How do we obtain an estimate of the vertical separation for the growth fault to be used in the
structure mapping of Bed B? The restored top method outlined earlier must be used. The
upthrown restored top for Bed B in Well No. 1, by correlation with Well No. 2 in the upthrown
fault block, is at 7705 ft measured log depth (Fig. 7-44). The downthrown restored top in Well
No. 1 for Bed B, by correlation with Well No. 3 in the downthrown Block, is at 7980 ft measured
log depth. Using the restored top method, the vertical separation for the growth fault at Bed B is
estimated as the difference in restored top depths for Bed B (7980 ft – 7705 ft = 275 ft).
Therefore, in the preparation of a structure map on Bed B, the vertical separation value of 275 ft
is used to contour across Growth Fault 1. Also, in using this method, two additional control points
for mapping are estimated, these being the upthrown and downthrown restored tops in Well
No. 1. Remember, these tops must be corrected to subsea depths before they are used in the struc-
ture mapping of the horizon.
The estimated vertical separation of 275 ft for the fault cut for Bed B lies between the low
value of 190 ft and the high value of 315 ft for missing section derived by correlation with wells
on either side of the fault. The restored top method is the most accepted procedure for estimating
the vertical separation (missing section) on a growth fault at any particular horizon.
The Single Well Method. Given certain limitations, it is possible to use only one reference well
to closely approximate the vertical separation for a growth fault. This method is referred to as the
Single Well Method. It is applicable only if the given horizon is close to the top or to the base of
the stratigraphic section missing in the faulted well. This method is illustrated in cross-sectional
view in Fig. 7-45.
316 Chap. 7 / Fault Maps
Figure 7-44 Cross section illustrates the restored top method for estimating the vertical separation
for a growth fault. The restored tops are estimates using the procedures discussed in Chapter 4.
Three wells are in the cross section in Fig. 7-45: Well No. 3 is in the upthrown fault block,
Well No. 2 is in the downthrown fault block, and Well No. 5 is the well of interest. The growth
fault has faulted out a significant stratigraphic section from Bed B through Bed F in Well No. 5.
For mapping purposes, we are interested in the vertical separation for the growth fault at Beds B,
D, and F.
First, we estimate the missing section in Well No. 5 with the other two wells (Fig. 7-45). By
correlation with Well No. 3 in the upthrown block, the estimated missing section due to the
growth fault is 200 ft, and with Well No. 2 in the expanded downthrown fault block the estimat-
ed missing section is 400 ft. There is a two-to-one difference in the estimated missing section at
the fault by correlation with these two wells. Which one is correct, or are they both correct?
Which one, if any, should be used for vertical separation in the structure mapping of the various
horizons? We know that this is a growth fault, which means by definition that the vertical sepa-
ration increases with depth. Therefore, it is reasonable to assume that vertical separation increas-
es from Bed B to Bed F. But by how much, and what is the vertical separation at each horizon?
The vertical separation at any horizon can be estimated by using the Restored Top Method,
whereas the vertical separation at only two horizons can be estimated by using the Single Well
Method. We will begin by using restored tops to estimate the vertical separation at Bed B (Fig.
7-45). The upthrown restored top for Bed B is at 8050 ft, and the downthrown restored top is at
8250 ft. The difference between these two restored top values is 200 ft, which is an estimate of
the vertical separation for the fault at Bed B. This is recorded within the insert in Fig. 7-45, which
contains a summary of the data pertinent to Beds B, D, and F.
Growth Faults
Figure 7-45 The vertical separation is estimated for this growth fault using both the restored top
317
and single well methods. The displacement is determined at the levels of Beds B, D, and F.
318
(a)
(b)
Figure 7-46 Segments of the integrated structure maps for Beds B, D, and F. Notice how the
fault gap width changes with depth in response to the increase in vertical separation. Well depths
are converted to subsea depths (KB = 50 ft for each well).
Growth Faults 319
The single well method can be used to estimate the vertical separation for a growth fault, but
it can be applied only under certain restraints. Notice in the cross section (Fig. 7-45) that Bed B
is at the top of the section missing in Well No. 5; therefore, we can say that most of the interval
faulted out of Well No. 5 can be represented by the thin upthrown section seen in Well No. 3.
Therefore, the vertical separation at Bed B can be closely estimated simply by correlating Well
No. 5 only with the thin upthrown Well No. 3. By making this correlation, we obtain an estimate
of 200 ft for the missing section due to the fault. Observe the 200 ft of section faulted out of Well
No. 5, as illustrated on the well log stick for Well No. 3. We can check the accuracy of this sim-
plified technique by comparing the result with that obtained from the restored top method (Fig.
7-45). They are the same. Using the two methods independently, the vertical separation of the
growth fault at Bed B is estimated to be 200 ft.
Now estimate the vertical separation at Horizon F. First, using the restored top method, the
estimated vertical separation at this horizon is 400 ft. Again, the results are summarized in the
insert in Fig. 7-45. Reviewing the log sticks in cross section, notice that Bed F is at the base of
the missing section in Well No. 5. Therefore, most of the section faulted out of Well No. 5 is rep-
resented by the thick downthrown section seen in Well No. 2. So, we can use the simplified sin-
gle well method and closely estimate the vertical separation at Bed F by correlating Well No. 5
with just one well, the thick downthrown Well No. 2. With this correlation, the vertical separa-
tion at Bed F is estimated to be 400 ft. Again, the estimate for the vertical separation is the same
using the more detailed restored top method or the simplified single well method.
The last horizon requiring an estimated vertical separation is Bed D. Since Bed D is not near
the top or bottom of the missing section in Well No. 5, the simplified one-well method cannot be
used. Therefore, the restored top method for estimating the vertical separation at Bed D must be
used, and it is illustrated in Fig. 7-45. The upthrown restored top is at 8150 ft, and the down-
thrown restored top is at 8450 ft, resulting in an estimated vertical separation of 300 ft at Bed D.
Figure 7-46 shows a portion of the structure map for each of the three horizons, Beds B, D,
and F. For mapping purposes, well depths are converted to subsea depths by using a KB = 50 ft
for each well. The beds are dipping and their apparent dip is horizontal in the cross section in Fig.
7-45. The maps were accurately constructed by the integration of the growth-fault surface map
with each horizon map, as described in Chapter 8. Note how the restored tops for Well No. 5 are
honored by the structure contours. The maps collectively demonstrate that the fault gap widens
with the increase in vertical separation with depth, as is typical for any growth fault delineated
on a set of horizon maps. Also, from map to map, the position of Well No. 5, within the fault gap,
changes relative to the upthrown and downthrown traces of the fault. That also is typical for any
well in which mapped horizons are faulted out by the same fault. Compare the three maps with
the cross section in Fig. 7-45 to visualize why the position of Well No. 5 changes relative to the
traces of the fault. If the horizon being mapped is at or near the top of the missing section in the
faulted well, the fault size vertical separation can be closely approximated by using the one-well
method and correlating the faulted well with a nearby well in the thin upthrown fault block. See
Bed B in our example in Fig. 7-44. In map view this is represented in Fig. 7-45a, which is a por-
tion of the structure map on Bed B.
If the horizon being mapped is at or near the bottom of the missing section in the faulted
well, the missing section fault size can be closely approximated by using the one-well method
and correlating the faulted well with a nearby well in the thick downthrown fault block. This was
done for Bed F, as shown in Fig. 7-46. In map view this is represented in Fig. 7-47c, which is a
portion of the structure map on Bed F. The faulted Well No. 5 is near the upthrown trace of the
growth fault. If the horizon being mapped falls clearly within the missing section in the well log
320 Chap. 7 / Fault Maps
(D Horizon, Fig. 7-44), then the restored top method must be used. The restored top method is
always preferable over the single well method if time is not a factor. In addition to providing an
estimate of vertical separation for the fault at any horizon, it also contributes two control points
for contouring (the upthrown and downthrown restored tops).
The faults used to illustrate the methods for estimating the vertical separation for a growth
fault do possess linear expansion, and the beds are gently dipping in Fig. 7-46. However, these
methods are just as applicable for growth faults with nonlinear expansion and in areas with
steeply dipping beds.
(a)
Figure 7-47 (a) Fault surface map of a growth fault. Vertical separation contours are placed on the
fault map, adding valuable mapping information pertaining to the growth fault. (b) Vertical separation
values are not contoured, but strategically placed on the fault surface map, which is a shortcut method.
(From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
Growth Faults 321
(b)
Once the vertical separation data are gathered, there are at least two ways to map the data. The
vertical separation values can be contoured on the fault surface map as shown in Fig. 7-47a, or
they can be recorded at strategic locations on the fault surface map as shown in Fig. 7-47b. The
actual contouring of the vertical separation is by far the more accurate method and the one we
recommend.
Looking at the fault surface and vertical separation contour map in Fig. 7-47a (which is only
a segment of the entire fault), we can quickly observe a lot of valuable information about this
growth fault. The sources of the fault surface and vertical separation data are both well logs and
seismic sections. The fault surface contour spacing indicates that the fault has a listric shape with
a maximum dip rate of 60 deg at around –5000 ft, decreasing to 40 deg at about –11,000 ft. Both
the vertical separation values next to each well and seismic shotpoint and the vertical separation
contours indicate that the vertical separation increases with depth from 250 ft at –4688 ft in Well
No. 12 to 590 ft at –10,847 ft in Well No. 3. Based on the spacing of the vertical separation con-
tours, the fault growth appears to be generally uniform over this area of the fault. Also, the con-
tours show that the vertical separation increases laterally from the northeast to the southwest.
Taking the –9000 fault contour as an example, in the southernmost portion of the map the verti-
cal separation at this depth is 500 ft. Following along this –9000-ft contour to the northeast, this
displacement decreases steadily to only 400 ft at the northern end of the map. Lastly, the contours
point to the direction of maximum growth fault activity (major sedimentary depocenter for this
fault) as being to the southwest. All this information is important for future integration of this
fault interpretation with structure and stratigraphic data to prepare detailed structure and isochore
maps for use in the exploration and exploitation of localized hydrocarbon accumulations.
The pattern of fault surface and vertical separation contours is different for nongrowth
(structural) faults than it is for growth faults. Refer once again to Fig. 7-40a and the discussion.
Compare this map with that for the growth fault in Fig. 7-47a. Identify the differences observed
and establish criteria that can be used to distinguish a growth fault from a post-depositional fault
using the fault maps.
A-2
A-1 A-5
A-6
A-4
(a)
(b)
Figure 7-48 (a) Straight-line method of plotting directional wells in map view. (b) Detailed plot of directional
survey data indicating the location and subsea depth of each wellbore along its entire length.
324 Chap. 7 / Fault Maps
1000 ft), of all directional wells on a base map. The benefits to structure and isochore mapping
are covered in the appropriate chapters.
An example will best show the importance of detailed wellbore plots in the interpretation
and construction of fault surface maps. The wellbore base map segments in Figs. 7-49 and 7-50
are from an offshore oil and gas field. Table 7-1 contains a simplified tabulation of fault data from
wells in the area of Fault C. For each well, the fault data provided are (1) subsea depth, (2) miss-
ing section (vertical separation), and (3) rectangular coordinates (obtained from directional sur-
veys and measured in feet). The rectangular coordinates provide, in map view, the location of the
fault in each well as measured from the surface location.
Using the simplified base map construction in Fig. 7-49, the location of each wellbore fault
pick must be plotted by using the rectangular coordinates to measure the position of the fault pick
from the surface or platform location. An “x” is used to represent the fault pick for each individ-
ual well. Notice that the fault position in map view for several wells, such as Well No. C-1, does
not coincide with the dashed directional path of the wellbore, because this simplified directional
wellbore path considers only the surface and bottomhole locations of the wellbore. Directional
wells are seldom drilled in a straight line from surface to bottomhole, because of either wellbore
design or natural wellbore drift. Therefore, the simplified wellbore plot for directional wells in
many cases does not follow the true and accurate path of the wellbore. Since each wellbore fault
pick in Fig. 7-49, represented by an “x,” was located using the rectangular coordinates calculat-
ed from the actual directional survey for each well, we assume the location for the fault picks to
be accurate (see plot for Well No. C-1).
In Fig. 7-50, we see the same base map area with the actual path and subsea depth for each
directional well plotted in 500-ft subsea increments. Such a plot is prepared once and can be used
for all future subsurface fault, structure, and isochore mapping.
Take another look at the data for Fault C using the detailed wellbore base map in Fig. 7-50.
With this base, the only data needed to post the fault in each well is the subsea depth for the fault.
For example, the subsea depth for the fault in Well No. C-1 is –10,905 ft. This datum point is
posted on the map in Fig. 7-50, using the detailed directional well plot, and is as accurate as the
same point plotted in Fig. 7-49. By using the detailed base map, however, it was not necessary to
calculate the rectangular coordinates from the directional survey for each fault pick nor to use
these coordinates to measure the location of the pick from the platform location. The elimination
of these steps saves mapping time. The example shown here is relatively simple, with only one
fault per well and only ten wells. But consider a field with one hundred wells and ten faults result-
ing in two or three fault picks per well. This adds up to some 200 to 300 fault data points for
which the rectangular coordinates must be calculated and then used to measure the location for
each fault. These calculations and measurements, which can be time-consuming, are unnecessary
if the base map is prepared using the actual directional survey data to plot accurately the path and
subsea depth for each wellbore along its entire length (Fig. 7-50). If you are using a workstation,
the proper directional well paths should be loaded into the project database. Be sure to check the
wellbore plots on the printed base map against each directional survey to verify the plots. Errors
are common, and they may result in costly incorrect interpretations. Also, if you work a 3D data
set over an old field, check the well files thoroughly to insure that directional surveys for all wells
are loaded in the database.
If saving time were the only benefit derived from preparing a base map with the detailed
directional wellbore plots, it would be worth doing, but other benefits may be even more impor-
tant. Look at Well No. C-4 in Fig. 7-50. No fault pick is assigned to this well for Fault C. Should
there be a fault in this well? Was it missed during correlation? Is the well deviated in such a way
Directional Surveys and Fault Surface Maps 325
Figure 7-49 Simplified straight-line plots for directionally drilled wells provide no information about
the position and depth of the wellbore between the surface and bottomhole locations.
that it does not intersect the fault, or is there no fault in the well? The answers to these questions
are very important and play a vital role in the interpretation of Fault C. Where do we look for
answers to these questions? If the base map has the detailed directional wellbore plots like that
in Fig. 7-50, it might be the first place to start looking for answers, or at least it might provide a
direction for further investigation.
Well No. C-4 in Fig. 7-50 does not have an assigned fault. Starting with the interpreted fault
surface map for Fault C and the detailed wellbore plot for C-4, can we address any of the ques-
tions asked with regard to the absence of a fault pick in the well? The answer is yes. Well No. C-4
is drilled to a subsea depth of –12,346 ft and is an S shape well. Look at the fault surface map
along the wellbore path from –11,000 ft to –11,500 ft. At a fault surface depth of –11,000 ft, the
wellbore is at a depth of –6300 ft, or 4700 ft vertically above the fault surface. At a fault surface
depth of about –11,500 ft, the wellbore is at a total depth of –12,346 ft, or 846 ft below the fault
surface. Therefore, if the fault surface interpretation is correct and the wellbore plot accurate, then
somewhere between –11,000 ft and –11,500 ft the wellbore should intersect Fault C (Fig. 7-51).
Since the wellbore is vertical below about –10,000 ft, the wellbore should penetrate the fault just
above –11,500 ft. Maybe the missing section due to the fault was unrecognized or questionable
during the initial correlation work.
With the new information obtained from the review of the fault map and detailed wellbore
plot, the correlations in Well No. C-4 can be further reviewed. If there is a fault in the well, it will
326 Chap. 7 / Fault Maps
Figure 7-50 Directional survey points plotted accurately over the entire length of the
wellbore are helpful in the interpretation, construction, and evaluation of fault, structure,
and isochore maps.
Table 7-1
Fault Data for Fault C
Directional Surveys and Fault Surface Maps 327
add another datum point to the fault surface interpretation. If no section is missing and thus no
fault, then we may conclude that the present interpretation is incorrect and another fault inter-
pretation is required, using the negative control from Well No. C-4. Negative control in this case
means no fault pick in the well. If there is no fault, then the fault surface interpretation must be
changed to honor the data. The possibility also exists that the fault passes below the well or the
fault may die before it reaches the well. All possibilities must be considered. In this case, how-
ever, further correlation of Well No. C-4 confirms a 90-ft fault at –11,480 ft, which was originally
documented on the electric log as a questionable short section. This fault information can now be
added to the fault map in Fig. 7-50.
The D-1 well, marked with an asterisk in Fig. 7-50, is similar to Well No. C-4 in that no fault
pick is assigned to this well for Fault C. Assuming that the fault surface interpretation is correct
and the wellbore plot is accurate, should Well No. D-1 intersect Fault C? And if so, at what sub-
sea depth?
We have shown that the detailed plot of a directional well, including its position and subsea
depth at regularly spaced increments on a base map, can save mapping time, be very beneficial
in helping evaluate a fault interpretation, and provide another check for questionable fault corre-
lations. Finally, supervisors, managers, and prospect evaluators can use the directional well data
and techniques discussed here to quickly review or evaluate fault maps. Where directional survey
data are available, we strongly recommend that the deviated well paths be plotted on a base map
to show the position and subsea depth of the borehole throughout its entire length.
tions allowing a maximum deviation before an actual directional survey is required by law.
The fault surface maps in Fig. 7-52a and b illustrate the kinds of problems that can arise
when wells are assumed to be vertical or when no directional well data are available. The fault
surface map in Fig. 7-52a was prepared using the available well data shown on the map. Each
depth value represents the subsea depth at which this mapped fault was encountered in the adja-
cent well. The map does not look too bad, but it does have some unusual variations in contour
spacing (dip of fault) and indicates a problem in the northeast part of the map. According to the
available data on the base map and electric logs, all the wells are identified as being vertical.
However, looking at the fault surface map, it is apparent that something is wrong. Either the fault
picks are incorrect, the wells are spotted at the wrong surface location, or some of these vertical
wells are not vertical.
The answer to the problem is shown in Fig. 7-52b. Upon additional review of well files, five
wells were found to be deviated, as shown in the figure. By replotting the wellbores as deviated
wells and respotting the fault picks to correspond to the true wellbore deviation, the fault surface
map shown in this figure was constructed.
Well No. 6 is of special interest. The only available directional data on this well was a Totco
survey, which measures angle but not direction. In the case of Well No. 6, the assumption was
made that the deviation was all in the same direction. Therefore, a circle whose radius is equal to
the distance the well could have deviated from the vertical to a depth of –7200 ft was plotted. If
the wellbore deviation were all in the same direction, the fault pick should fall somewhere on this
(a)
circle. The direction of deviation that best fit the fault was used to spot the pick, as shown in the
figure. Notice that the contour spacing has evened out and the major contouring problem in the
northeast was resolved.
Keep this example in mind, especially when mapping in areas with very old wells. A prob-
lem that might at first appear to be that of wrong correlations or an incorrect interpretation might
turn out to be nothing more than a problem of a vertical well not being vertical.
Figure 7-53 Directionally drilled Well No. 212 intersects Fault H four times, resulting in two missing
sections and two repeated sections. The well changes fault blocks four times. (Published by permission
of Texaco, USA.)
Vertical Separation – Correction Factor and Documentation 331
Figure 7-54 Portion of the fault map for Fault H. Note the four fault data points
from Well No. 212. (Published by permission of Texaco, USA.)
CHAPTER 8
STRUCTURE MAPS
INTRODUCTION
The subsurface structure map is one of the primary vehicles used by geoscientists to find and pro-
duce hydrocarbons from the initial stage of exploration through the complete development of a
field. Each subsurface structure map is a geologic or geophysical interpretation based upon lim-
ited data, technical proficiency, creative imagination, three-dimensional visualization, and expe-
rience. We consider the construction of a structure map to be an interpretive and creative process.
No two geoscientists will construct a map exactly the same, even with the same data, because
each uses the factors just mentioned in addition to educational background and field experience
to develop his or her interpretation. However, the interpretation must incorporate sound geolog-
ic principles, correct and accurate mapping techniques, and be valid in three dimensions.
The importance and reliability of subsurface structure mapping increase with advancing
stages of field development and depletion. Many management decisions are based on the inter-
pretations presented on subsurface structure maps. These decisions involve investment capital to
purchase leases, permit and drill wells, and to work over or recomplete wells, to name a few
examples. A geoscientist must employ the best and most accurate methods to find and develop
hydrocarbons at the lowest cost per net equivalent barrel.
Since faulted structures play such a significant role in the trapping of hydrocarbons, we
devote a considerable portion of this chapter to the correct and accurate subsurface mapping tech-
niques required to integrate fault surface map interpretations into the structural interpretation to
construct completed structure maps. A reasonable structural interpretation, in most faulted areas,
begins with an accurate fault picture developed from the interpretation of faults and the con-
struction of fault surface maps using fault data from well logs and seismic sections (Chapter 7),
followed by the integration of these fault surface maps into the structural interpretation.
332
Introduction 333
Many petroleum provinces involve multiple faults that result in extremely complicated struc-
tural relationships. The attempted reconstruction of a complex structure with isolated fault data
from well logs or seismic sections can result in erroneous geologic interpretations. Too often,
subsurface interpretations and the accompanying structure maps are prepared without giving
much consideration to the three-dimensional geometric validity of the interpretation. The most
accurate and sound structural interpretation in a faulted area requires (1) the interpretation and
construction of fault surface maps for all important structure-forming and trapping faults, (2) the
integration of the fault surface maps with the structural horizon maps, and (3) mapping of multi-
ple horizons at various depths (shallow, intermediate, and deep) to justify and support the integri-
ty of any structural interpretation (Tearpock and Harris 1987).
The exploration for and exploitation of hydrocarbons is interpretive and creative work. Most
of the time a geoscientist is dealing with geologic structures that are not visible on the surface.
The formidable challenge of interpreting these unseen structures can be accomplished only with
a clear understanding of basic geologic principles, familiarity with the geometry of structural and
fault surface relationships, analysis of all available data, use of all technical capabilities, applica-
tion of technical knowledge and skills, and imagination.
In this chapter, we concentrate on the technical knowledge and skills necessary to develop a
geologically reasonable structural interpretation. Technical knowledge and skills fall into two cat-
egories: (1) a good understanding of the tectonic setting being worked, and (2) understanding and
application of correct interpretation and mapping techniques. The primary focus of this chapter
is on the broad range of important structural mapping techniques; however, since the application
of many techniques depends upon the tectonic style (type of structure and trap), we discuss and
illustrate techniques as they apply to different tectonic settings and review a number of real-world
examples.
Subsurface structure maps usually are constructed for specific stratigraphic horizons to
show, in plan view, the geometric shapes of these horizons. These maps are constructed using cor-
relation data from well logs, interpretations of seismic sections, and in some cases, outcrop data.
Remember that accurate correlations are paramount for reliable subsurface interpretation and
mapping. Subsurface structure maps are no more reliable than the correlations used in their con-
struction. Incorrect correlations will find their way, at some point, into the final interpretation.
They may be incorporated into the fault, structure, isochore, or isopach maps and result in seri-
ous mapping problems. Therefore, it is essential that utmost care be taken in correlating logs and
interpreting seismic sections.
Not every horizon within a stratigraphic sequence is suitable for structure mapping. A hori-
zon that is not correlatable over a large area or one that is limited in areal extent may not be suit-
able. Maps on stratigraphic horizons of limited extent, if important, can be prepared after the
overall structural interpretation has been developed from fieldwide or regional correlations and
structure maps.
A structure map is a form of contour map. As discussed in Chapter 4, marine shales exhibit
distinctive characteristics over large areas. Therefore, they serve as good correlatable horizons for
fieldwide or regional structure mapping. A structure contour map presents a 2D interpretation of
the 3D shape of a specific stratigraphic horizon. Each contour connects points of equal elevation
above or below sea level for a given stratigraphic horizon. A good structural interpretation
requires 3D thinking, as illustrated in the simplified block diagram in Fig. 8-1.
Broadly interrelated assemblages of geologic structures constitute the fundamental structur-
al styles of petroleum provinces. These assemblages generally are repeated in regions of similar
334 Chap. 8 / Structure Maps
Figure 8-1 A three-dimensional view of an anticlinal structure 7000 ft below sea level.
deformation, and the associated types of hydrocarbon traps can be anticipated (Harding and
Lowell 1979). There are a number of petroleum-related tectonic habitats around the world; each
results, to varying degrees, in different kinds of hydrocarbon traps that may require modified or
different mapping techniques. In the first part of this chapter, we discuss numerous subsurface
structure mapping techniques. These techniques are then reviewed as they apply to the following
tectonic habitats and their associated hydrocarbon traps.
1. Extensional terranes, including normal faulting and detached listric growth fault systems.
2. Compressional terranes, including reverse faulting and fold and thrust belts.
3. Diapiric salt terranes.
4. Strike-slip fault terranes.
GUIDELINES TO CONTOURING
Review the five basic rules of contouring presented in Chapter 2. In addition to these basic rules,
the following guidelines to contouring make a map easier to construct, read, and understand; they
also help to ensure the technical accuracy and correctness of the completed map. Some guide-
lines covered in Chapter 2 are repeated here; many have been expanded, and additional guide-
lines are presented.
1. All contour maps must have a chosen reference to which the contour values are com-
pared. A structure contour map usually uses sea level as the chosen reference. Therefore, the ele-
vations on the map can be referenced above or below mean sea level. A negative sign in front of
a depth value indicates that the elevation is below sea level.
2. The contour interval on a map should be constant. The use of a constant contour inter-
val makes a map easier to read and visualize in 3D because the distance between successive con-
tour lines has a direct relationship to the steepness of slope. Steep dips are represented by close-
ly spaced contours, gentler dips by contours with a wider spacing. Figure 8-2 illustrates the con-
fusion and difficulty involved in trying to visualize a contoured surface in 3D where the contour
interval is not constant over the mapped area. From fault block to fault block, the contour inter-
val changes from 100-ft to 50-ft contours with no consistency. Notice upthrown to Fault A that
the contour interval is 50 ft, and downthrown it is 100 ft, yet the contour spacing is about the
Guidelines to Contouring 335
Figure 8-2 This structure map has an inconsistent contour interval randomly changing from a 50-ft to 100-
ft contour interval from fault block to fault block. Such inconsistency in the contour interval makes a map dif-
ficult to visualize in three dimensions. Observe the change in contour interval upthrown and downthrown to
Fault A, and even in the fault block upthrown to Fault A. (From Tearpock and Harris 1987. Published by per-
mission of Tenneco Oil Company.)
same. This indicates that the area downthrown to Fault A has a much steeper dip than the area
upthrown. However, when we look at this area of the map, the contour spacing gives the illusion
that the dip rate in both fault blocks is about the same. The contour interval changes even within
the fault block northeast of Fault A.
The choice of a contour interval is an important decision. Factors to be considered include
the density of data, the practical limits of data accuracy, the steepness of dip, the scale, and the
purpose of the map.
3. Contour spacing depends upon the dip of the structure being mapped. For any given
structure, the spacing of contours will vary at different locations unless the equal spacing method
of contouring is used. Several graphs are designed for convenient use to compute contour spac-
ing when the dip is known; likewise, the dip on a completed map can be determined by measur-
ing the contour spacing. Figure 8-3 is a graph that relates the dip of beds to the horizontal dis-
336 Chap. 8 / Structure Maps
Figure 8-3 Graph of dip versus horizontal distance (in feet) between 100-ft contours.
tance between 100-ft contours. It can be used for determining dip or contour spacing for fault
maps as well as structure maps.
4. All maps should include a graphic scale (Fig. 8-4). A graphic scale gives the viewer an
idea of the areal extent of the map and the magnitude of the features shown. Maps are common-
ly reduced or enlarged for various reasons; without a graphic scale, the values shown on the map
become meaningless.
5. Every fifth contour is an index contour. It should be bolder or thicker than the other con-
tours and labeled with its value (Fig. 8-4).
6. Hachured lines should be used to indicate a closed depression (Fig. 8-5).
Guidelines to Contouring 337
7. Contouring should be started in areas with the maximum number of control points
(Chapter 2, Fig. 2-8). The area or areas of maximum control commonly occur around structural
highs or lows.
8. Construct contours in groups of several lines rather than one single contour at a time
(Fig. 2-8). This method provides better visualization of the surface being contoured and results
in more consistent contouring.
9. Initially, choose the simplest contour solution that honors the control and provides a
realistic interpretation. The simplest solution may be the best (Occam’s Razor), and it is usually
easy to test. If problems arise with this solution, a more complex interpretation can be prepared.
10. Use a smooth rather than undulating style of contouring unless the data indicate other-
wise. Some geoscientists argue that a smoothly contoured structure is not likely to occur in nature
(Chapter 2, Fig. 2-9). This may be true; however, it is better to keep the structure simple with
smooth contours until the data indicate otherwise. It is possible to present a significant misinter-
pretation by placing unjustified wiggles in contours (Silver 1982).
11. Initially, a hand-contoured map should be contoured in pencil with lines lightly drawn
so that they can be erased as the map requires revision.
12. Establish regional dip whenever possible. Regional dip is the general direction of dip
for any given area. Regional dip may not be constant over a large area, but changes should be
gradual. In areas of regional dip, contour lines have a certain degree of parallelism along region-
al strike. Any change in the dip rate may be an indication of local structures. In areas of minor or
localized structures, contours extend away from regional dip. Such indications are important
because in many petroleum provinces minor anomalous highs that break regional dip commonly
are productive of hydrocarbons (Fig. 8-4).
If regional dip is interrupted by a localized structural high, reentrants occur on each side of
the minor uplift (Figs. 8-4 or 8-5). If the axis of the localized uplift parallels regional strike, the
magnitude of the reentrants may be small compared to reentrants adjacent to a high that is per-
pendicular to regional strike (Fig. 8-5).
Any flattening or reversal of normal dip is a possible clue to local structures. Therefore,
changes of this kind are extremely important. Local uplifts may have their axes perpendicular or
parallel to the regional strike. When the axis of a local fold is perpendicular to strike, contours
flare outward in a down-dip direction and the distance between contour lines increases as the rate
Figure 8-4 An example of a localized structural high indicated by a change in regional dip.
338 Chap. 8 / Structure Maps
Figure 8-5 A diagrammatic structure contour map of a basin illustrating several important contour-
ing guidelines. (From Bishop 1960. Published by permission of author.)
of dip decreases. A nosing or U-shaped projection results in the bottom of the U pointing basin-
ward (Fig. 8-5). Nosings are flanked by reentrants, the axes of which may be perpendicular or
oblique to the regional strike. The reentrants begin where the contours start to widen out and are
less pronounced down-dip until eventually they disappear.
If the axis of a structurally high area is parallel to regional strike, reentrants are also paral-
lel to strike. As the direction of regional dip reverses at the axis of the reentrants, a high area
results down-dip, as shown in Fig. 8-5 (Bishop 1960).
13. Contouring can be optimistic or pessimistic depending upon your experience, corporate
guidelines, and exploration philosophy. All contouring, however, must be governed by sound
geologic principles and the general tectonic style, and optimism must be kept within geological-
ly reasonable limits. Pessimistic contouring can condemn potentially prospective areas to the
point that no exploratory drilling is undertaken. A good mapping philosophy to follow is to map
neither optimistically nor pessimistically, but instead to use all of your technical expertise to map
realistically.
14. In areas of either limited subsurface control or vertical faults, it is important to contour
the limited data to reflect as simple a geologic interpretation as possible, rather than just to con-
nect points of equal elevation. Therefore, any radical change that occurs in the strike of the con-
tours may suggest faulting even though no fault has been recognized by well control. Figure 8-6a
depicts such a situation. In these cases, all available data need to be reviewed, including produc-
tion and pressure data to help resolve the geologic problem. In the example shown in Fig. 8-6,
notice a significant change in contour strike in the area marked as a possible fault, although no
fault is recognized in the wells. An interpretation that fits all the geologic and hydrocarbon data
includes a vertical fault not intersected by the wells (Fig. 8-6b).
An abrupt increase in the rate of dip is a good indication of faulting. An increase in the rate
of dip accompanied by an abrupt change in strike is very strong evidence of faulting (Bishop
1960). Increased dip might, alternatively, result from folding, but in most cases the increase is
more abrupt where faulting is responsible.
Guidelines to Contouring 339
15. A change or reversal in the direction of dip suggests the crossing of a fold axis (Fig. 8-5
or Fig. 8-7). Reversal of dip may occur over the crest of an anticline or the trough of a syncline.
The amount of dip reversal is often a guide to whether the reversal is due to a regional change or
local structure. An excessively steep dip may indicate the presence of a fault or steep fold, while
a dip that flattens may be indicative of the crest of a fold or the bottom of a syncline.
16. Structures may or may not have structural attitude compatibility (contour compatibili-
ty) across a fault. The compatibility of structural attitude on opposite sides of a fault depends pri-
marily upon the size and type of fault. For example, within many, if not most, structures, struc-
tural compatibility exists across normal and reverse nongrowth faults (Fig. 8-8a). In contrast,
many large listric normal faults (such as growth faults) and thrust faults, with significant dis-
placements, result in structures that are not compatible across the large fault (Fig. 8-8b). The
method of structure contouring across a fault depends upon whether there is a compatibility of
structural attitude on both sides of the fault.
(a)
(b)
Figure 8-6 (a) An abrupt change in the strike direction of contours suggests the possibility of faulting.
(b) Another interpretation of Fig. 8-6a that fits the geologic and hydrocarbon data includes a vertical
fault not intersected by the wells. (Modified after Bishop 1960. Published by permission of author.)
340 Chap. 8 / Structure Maps
Figure 8-7 Structural highs and lows can sometimes be recognized by their effect on contour spac-
ing. (From Bishop 1960. Published by permission of author.)
17. Structural highs, regardless of origin, usually tend to flatten across the axis with gentle
dips across the top of the structure (Fig. 8-7). Exceptions do occur, such as the structure in the
inner core of a fault propagation fold (Chapter 10). Contour line spacing widens across the crest
of the structure compared to spacing on the flanks. Synclines, like anticlines, also tend to flatten
across their axes. The widening of contours is often even more pronounced in a syncline. If the
data indicate a continuously steep slope up to the crestal high with little if any flattening of dip,
this may indicate that the surface is one that has been affected by erosion (the presence of an
unconformity).
18. Closed structural lows are not common. If possible, avoid closing lows with contours
unless the data require it. The presence of closed lows often suggests an eroded surface or the
presence of faulting. If the closed low is elongated, faulting is likely, and the greater the size of
the closed low, the greater the probability of faulting.
19. Contour license refers to the geologist’s right to contour a structure in a way that best
fits the geologic, geophysical, and engineering data, and that best represents the types of struc-
tures present in the tectonic setting. The interpretive method of contouring as defined by Bishop
(1960) best corresponds to what we call contour license. The use of geologic license requires a
solid educational background and extensive experience.
20. Specific structural highs may be in the form of domes, anticlines, and noses. Domal
structures are usually the result of local positive features, such as diapirs, that provide relative
uplift. On the flanks of domes, the direction of dip is away from the central high and the dip rates
are commonly constant along strike, at least within each major fault block around the structure.
Therefore, the contour spacing is commonly uniform along strike. In the dip direction, the struc-
turally highest areas on the flank typically have the steeper dips, with a gradual decrease in dip
away from the uplift. Contour spacing is close high on the flank of the structure, but widens with
distance down-dip.
Anticlinal structures generally appear as elongated domes. Their origin can be the result of
compressional forces (e.g., fault-propagation folds associated with reverse faults), extensional
forces (e.g., rollover anticlines associated with growth normal faults), or strike-slip forces. In
general, the direction of dip is away from the crestal area in two opposing directions. Since anti-
clines are commonly asymmetric, with inclined axial surfaces, the dip rate and resulting contour
spacing may vary around the anticline.
Structural noses that trend off local structures show dip away from the crest in three direc-
tions. Contour lines widen, indicating flatter dips in the area of a nose or an associated reentrant.
Summary of the Methods of Contouring by Hand 341
(a)
Contours become closer together immediately down-dip of the local high until regional dip is
attained again.
The best test of the 3D geometric validity of any structure contour map is its predictability.
How well does the interpretation hold together with additional data from the drilling of new wells
or the shooting of new seismic lines? If the interpretation and maps require major revision each
time new data are obtained, there should be serious concern regarding the validity of the inter-
pretation and accompanying maps. On the other hand, if only minor adjustments are required and
hydrocarbon traps predicted by the mapping are successfully found, the interpretation may be
considered reasonable. Remember, we always work with a limited amount of subsurface data to
be interpreted. Each geoscientist must have imagination, an understanding of local structures, the
ability to visualize in three dimensions, a sound geologic education, field experience, and tech-
nical knowledge and skills to evaluate any number of possible (alternative) interpretations.
Finally, a geoscientist must decide which interpretation, in his or her judgment, is the most rea-
sonable. The 20 guidelines presented in this section should help you construct more accurate and
reasonable structure contour maps.
1. Mechanical Contouring. In using this method of contouring, the assumption is made that
the slope or angle of dip of the surface being contoured is uniform between points of
control and that any change occurs at the control points. With this approach, the spacing of
the contours is mathematically (mechanically) proportioned between adjacent control
points. Mechanical contouring allows for little, if any, geologic interpretation. Even though
the map is mechanically correct, the result may be a map that is geologically unreasonable,
especially in areas of sparse control (Fig. 8-9a).
342 Chap. 8 / Structure Maps
2. Parallel Contouring. With this method, the contour lines are drawn parallel or nearly
parallel to each other. This method does not assume uniformity of slope or angle of dip;
therefore, the contour spacing can vary. Like the previous method, if honored exactly,
parallel contouring yields an unrealistic geologic picture. It allows for some geologic
license to draw a map a little closer to the real world, because there is no assumption of
uniform dip (Fig. 8-9b).
3. Equal-Spaced Contouring. This method assumes uniform slope or angle of dip over an
entire area or at least over an individual flank of a structure. Sometimes this method is
referred to as a special version of parallel contouring. The advantage to this method, in the
early stages of mapping, is that it can indicate the maximum number of structural highs and
lows expected in an area of study (Fig. 8-9c).
4. Interpretive Contouring. With this method, the geoscientist has extreme geologic license
to prepare a map to reflect the best interpretation of the area of study, while honoring the
available control. No assumptions, such as constant bed dip or parallelism of contours, are
made. Therefore, the geoscientist can use his or her experience, imagination, ability to
think in three dimensions, and an understanding and knowledge of the structural and
stratigraphic styles of the geologic region to develop an accurate and realistic
interpretation. This method is the most acceptable and the most commonly used method of
contouring (Fig. 8-9d).
The specific method or combination of methods chosen for hand-contouring may be dictat-
ed by such factors as the number of control points, the areal extent of these points, and the pur-
pose of the map. No individual can develop an exact interpretation of the subsurface with the
same accuracy as that displayed on a topographic map. What is important is to develop the most
reasonable and realistic interpretation of the subsurface with the available data.
1. delineates the position of the footwall (upthrown) and hanging wall (downthrown) cutoffs
or traces of the fault in map view;
2. provides for the proper contouring of the mapped horizon across the fault;
3. depicts the vertical separation of the fault for any particular mapped horizon; and
4. defines the limits of fault-bounded reservoirs.
Fault surface mapping is covered in detail in Chapter 7. In this section, we present the prop-
er techniques for integrating a fault surface map with a structure contour map. Included in this
section are (1) techniques for positioning the upthrown and downthrown traces of a fault on a
structure map; (2) construction of the fault gap or overlap; (3) the mapping of vertical separation
versus throw; (4) the use of restored tops in structure mapping; (5) the application of contour
compatibility across faults; and (6) the exceptions to contour compatibility.
Contouring Faulted Surfaces 343
(a)
(b)
Figure 8-9 (a) Mechanical contouring method. (b) Equal-spaced contouring method. (c)
Parallel contouring method. (d) Interpretive contouring method. (Modified after Bishop
1960. Published by permission of author.)
344 Chap. 8 / Structure Maps
(c)
(d)
Figure 8-10 Integrated fault and structure map for the 6000-ft Horizon. The darkened circles
delineate the intersection of each structure contour with the fault contour of the same elevation.
(a)
(b)
Figure 8-11 (a) Faulted structure map on the 8000-ft Horizon. The structure is cut by a
400-ft fault. The correct method for contouring vertical separation (missing section) is illus-
trated by the dashed contour lines. (b) Fault surface map for Fault 1, which strikes east-west
and dips at 50 deg to the south. (Published by permission of D. Tearpock.)
Contouring Faulted Surfaces 347
Contours may be projected for some distance within the fault gap. Notice how the –8300-ft
contour is projected from the upthrown trace of the fault for some distance through the fault gap
before it intersects the downthrown trace and enters the downthrown block as an –8700-ft con-
tour. The mechanics of projecting contours, such as the –8300-ft contour, through the fault as
shown in this figure is the correct technique for contouring across a normal fault, using the ver-
tical separation as obtained from well log correlation (missing section) or seismic sections. The
application of this technique leads to correctly delineating the position of the upthrown and
downthrown traces of the fault, thus establishing the fault gap. It also assures that the correct dis-
placement has been mapped across the fault (Tearpock and Harris 1987; Tearpock and Bischke
1990).
Some of you may still be asking yourselves why throw was not contoured across the fault.
One reason is that no throw data are available for mapping, and throw is not the correct vertical
displacement we want to map across the fault (refer again to Chapter 7). However, if we want to
know the fault throw and heave, their values can be determined by simple measurements once the
structure contour map has been prepared, as shown in Fig. 8-11a or by use of Eq. (7-1).
Throw is the difference in the vertical depth between where the fault intersects a specific
horizon in the upthrown block and where it intersects that same datum in the downthrown block,
measured perpendicular to the strike of the fault surface (not perpendicular to the strike of the
fault trace). Therefore a fault surface contour map must be available in order to calculate the
throw from a completed structure contour map, as shown in Fig. 8-11a. The fault shown in Fig.
8-11b strikes east-west; therefore, the throw is determined in a north-south direction (see arrows
in fault gap through Well No. 7). Using the points A and B on the map (Fig. 8-11a), the upthrown
depth at point A is –8460 ft and the downthrown depth at point B is –8940 ft. The throw of the
fault at this location is the difference between these two depths, or 475 ft. Mathematically, by
applying Eq. (7-1), the throw is estimated to be 496 ft using an average bed dip of 13 deg and a
fault dip of 50 deg. Considering the accuracy of graphical measurements on a contoured map,
these two estimates for throw are in excellent agreement. We can see for this particular example
that the throw is about 75 to 96 ft greater than the vertical separation.
Heave, which is the horizontal distance across the fault gap from the upthrown to down-
thrown traces, measured perpendicular to the strike of the fault surface (not perpendicular to the
fault trace), is 390 ft. Therefore a fault surface map must be available to determine heave as well.
For subsurface petroleum-related structure mapping, the measurements of throw or heave are
usually measured for academic purposes and have little application in the actual contouring of a
structure map. However, the throw and heave can be used to check a completed structure map
using the graphical and mathematical methods described in this chapter and in Chapter 7. If the
estimates for throw or heave determined by both methods (graphical and mathematical) are rea-
sonably close, you can conclude that the map construction is reasonable. We point out here that
the mapping of throw and heave are important in subsurface mining, the mapping of subsurface
mineral deposits and outcrop mapping.
To further illustrate the proper construction of contours across a fault, we review Fig. 8-12,
using the same data given in Fig. 8-11, with one exception. In this case, we map a horizon about
2000 ft shallower, and at this level the fault trace falls on the northern flank of the structure. Here
the fault is dipping in the opposite direction as the horizon. The horizon, dipping generally to the
north, is displaced by the south-dipping fault. The fault has a vertical separation of 400 ft. From
the data points available, the structure contours were first established in the upthrown block
and extended to the upthrown trace of Fault 1. As shown in the figure, to contour across the fault
it is necessary to project the contours from the upthrown block through the fault gap to the
348 Chap. 8 / Structure Maps
Figure 8-12 Portion of the structure map for the 6000-ft Horizon, showing the
method for contouring vertical separation across a fault. (Published by permis-
sion of D. Tearpock.)
downthrown trace of the fault, and then into the downthrown block. Ask yourself where the con-
tour would be drawn had the structure not been faulted. The construction from one block to the
other is shown by a series of dashed contour lines placed from the upthrown trace, through the
gap, to the downthrown trace. Once the contours have been projected into the downthrown fault
block, they are adjusted in depth from the upthrown contour values by the amount of vertical sep-
aration, which in this case is 400 ft. As an example, the –6000-ft contour in the upthrown block
projected into the downthrown block becomes a –6400-ft contour. As in Fig. 8-11, the same tech-
nique is followed for construction of all structure contour lines.
Now that the structure contours have been established in both blocks, the fault throw and
heave can be calculated and measured. We graphically determine the throw of the fault by esti-
mating the upthrown and downthrown structural depths using points A and B on the map. Notice
that the values for throw and heave are different than those estimated in Fig. 8-11a. The value for
throw is now 325 ft, as compared with 475 ft in Fig. 8-11a. The value for heave is 280 ft, as com-
pared with 390 ft in Fig. 8-11a. It is important to note that although the vertical separation does
not change, there is a difference in the values for throw and heave. The missing section due to the
fault has not changed, as it is equal to vertical separation and not throw. Remember, throw and
heave are dependent fault slip variables that change with variation in the apparent attitude of the
fault or horizon. In this case (Fig. 8-12), the fault intersects the horizon where the horizon dip is
generally in the opposite direction to that of the fault. At the map level shown in Fig. 8-11a, the
fault and horizon are dipping in the same general direction. It is this change in relative dip direc-
tion (fault strike is unchanged) that causes the different values for throw and heave shown in the
two figures, even though the vertical separation has not changed. Using Eq. (7-1) and the data on
Contouring Faulted Surfaces 349
the map near Well No. 2 (average bed dip is 14.5 deg), estimate the throw across the fault at points
A and B and compare this estimate with that obtained graphically from the map.
Mapping Throw Across a Fault. Previously, we mentioned that missing section and repeated
section due to a fault are equal to vertical separation. Let us assume for a moment, however, that
a geoscientist incorrectly considers the fault data as throw and contours across the fault as if the
missing section were throw. In Fig. 8-13, the fault and structural data are exactly the same as that
in Fig. 8-11; therefore, we can compare the results of this (throw-contoured) map with the map
in Fig. 8-11a.
Using the available data, the contours are first established in the upthrown fault block and
extended to the upthrown trace of Fault F-1. When contouring throw across a fault, the strike
direction of a contour changes at this point and becomes perpendicular to the strike of the fault
surface (Fig. 8-13). The contour is then projected through the fault gap to the downthrown trace
of the fault. The strike direction of the contour is again changed to conform to its trend in the
upthrown block.
Follow the –8500-ft contour through its construction in order to gain a good understanding
of the technique. The strike direction of the –8500-ft contour in the upthrown block is established
by the surrounding well control. At the intersection with the upthrown trace of the fault, the strike
direction of the contour line changes abruptly to a strike direction that is perpendicular to the
strike of the fault surface. In this example, as in the one in Fig. 8-11a, the strike direction of the
fault surface is east-west. Therefore, treating the vertical separation as if it were throw, the con-
tour is projected through the fault gap in a north-south direction, perpendicular to the strike of the
Figure 8-13 Fault and structure data are the same as shown in Fig. 8-11a. For this
interpretation, the missing section is contoured across the fault incorrectly as if it
were throw. Compare the downthrown fault block with that shown in Fig. 8-11a.
(Published by permission of D. Tearpock.)
350 Chap. 8 / Structure Maps
fault surface. At the intersection of the contour and the downthrown trace of the fault, the con-
tour strike direction changes again to conform to the strike direction in the upthrown fault block.
Once the contour is projected into the downthrown fault block, its depth is adjusted from the
upthrown contour value by the amount of missing section (or in this case, assumed throw) of the
fault (400 ft), so it becomes a –8900-ft contour.
A proposed well location is in the downthrown fault block in Figs. 8-11a and 8-13.
Considering the correctly contoured map (Fig. 8-11a), the depth at which the proposed well is
estimated to penetrate this horizon is –8720 ft, whereas at the same location on the incorrectly
contoured map (Fig. 8-13), the well is estimated to penetrate the horizon at –8640 ft. The depth
to the horizon is mapped 80 ft shallower on the incorrect map. Contoured depth differences of 80
ft can make the difference between a successful well and a dry hole, or result in a well that is not
drilled in the optimum position on the structure.
Based upon the nature of this erroneous contouring technique (mapping vertical separation
incorrectly as throw), the magnitude of error becomes greater near or on the crest of a structure,
as well as becoming larger with increasing structural dip. This is very critical since hydrocar-
bons are often trapped near the crest of structures and we commonly map these areas.
Over the years, we have reviewed hundreds of maps that have been contoured incorrectly as
shown here. This is a common error based on a misconception of the meaning of throw. This error
has caused petroleum companies hundreds of millions of dollars in dry holes or poorly positioned
wells. The information presented here should provide you the basic knowledge needed to avoid
similar or costly errors (Tearpock et al. 1994). Furthermore, good field training in which maps
are made on faulted structures using outcrop and well log data can provide the three-dimension-
al understanding of these various fault components, their important differences, and the impact
of mapping vertical separation incorrectly as if it were throw.
Error Analysis – Procedure for Checking Structure Maps. The values for the vertical sep-
aration, fault dip, and bed dip can be determined from normal mapping parameters. With these
data, Eq. (7-2) can be used to calculate the throw at any point along a fault. If the calculated value
for throw agrees fairly well with the value determined graphically from the map, we can conclude
that the interpretation is reasonable. If the missing or repeated section were mapped as throw
instead of vertical separation, the value for throw determined mathematically will not compare
favorably with that determined graphically, indicating that the map is in error.
The graph presented in Fig. 8-14, which is derived from Eq. (7-1), can be used during daily
operations to check the consistency of structure contour maps and to ensure that a structure map
has been contoured correctly across any existing faults. Equation (7-1) is accurate for the situa-
tion in which the bed dip direction is more or less perpendicular to the strike of the fault surface
(regardless of whether the beds dip toward the fault or in the same direction as the fault). So the
graph in Fig. 8-14, and also the graph in Fig. 8-16, should be applied to that structural condition.
That relationship is common in prospective closures against faults, and prospects are where you
most want to check the accuracy of structure maps.
In Fig. 8-14, the fault and bed dips are taken to be clockwise. If the bed dips exceed the fault
dip and if the beds and fault dip are in the same direction, the ratio AE/AC (vertical
separation/throw) must be added to 1.0 in the figure. This is the case of a repeated section due to
a normal fault.
As a practice exercise, use the data in Fig. 8-11a, near the proposed location, and the graph
(Fig. 8-14) to verify the dip of Fault 1 shown in Fig. 8-11b.
Contouring Faulted Surfaces 351
An example of how to use Fig. 8-14 for steeply dipping beds encountered around a salt
dome, is as follows. On Fig. 8-14 the fault dip θ is on the y-axis, the bed dip φ is in the central
portion of the graph, and the absolute value of (AE/AC–1.0) is plotted on the x-axis. Assume a
fault has a dip of 50 deg. If the vertical separation AE is 400 ft and the throw AC is 200 ft, then
the ratio of AE/AC = 2.0. The absolute value of (AE/AC–1.0) = 1.0. Next construct a vertical line
from the abscissa of Fig. 8-14 at the value of 1.0 and a horizontal line across the graph from the
ordinate for a fault that dips at θ = 50 deg. The two lines cross the steeply dipping bed-dip curves
at about φ = 50 deg. Thus, a correctly contoured map is consistent with a bed that dips at 50 deg
into a fault that also dips at 50 deg (in opposite directions), for a vertical separation of 400 ft and
a throw of 200 ft.
The relationship shown in Fig. 8-15 is used to conduct error analyses on incorrectly con-
toured maps (i.e., how much error is introduced into a map that assumes the missing or repeated
section in a well is throw). By this time it should be readily apparent that if vertical separation is
mapped as if it were throw, the errors involved could result in a well that totally misses its target.
The graph in Fig. 8-16, derived from the relationship in Fig. 8-15, can be used to analyze the
magnitude of error (Tearpock and Bischke 1990). Figure 8-15 illustrates a faulted horizon, for
Figure 8-14 Graph used to check contouring across a fault. See text for explanation.
(Published by permission of D. Tearpock and R. Bischke.)
352 Chap. 8 / Structure Maps
which the footwall position is determined incorrectly by mapping vertical separation as throw.
From Fig. 8-15:
AE = A′C
Therefore, from the Law of Sines,
BA = TA / sin φ a
TA sin φ a cos θ (8-1)
=
A′C sin (θ – φ a )
π
φ+
2
Figure 8-15 The relationship shown here can be used to con-
π
duct error analysis on incorrectly contoured structure maps.
−θ Based on the bed dip and intersection of Bed A and the fault
2
in the downthrown block (hanging wall), observe the difference
between the correct versus incorrect depth of intersection of
Bed A and the fault in the upthrown block (footwall). Bed dip
and fault dip are virtually in the same direction. (Published by
permission of D. Tearpock and R. Bischke.)
The distance TA is the error in depth between the correct (B) and incorrect (A' ) points for
the horizon projected onto the footwall. The ratio TA/A' C provides a percent error relative to the
vertical separation value AE, which was incorrectly used as throw. Thus, Fig. 8-16 can be used
to measure the error that is introduced by improper contouring techniques.
On Fig. 8-16 the percent error TA/A' C is plotted on the y-axis against the fault dip q on the
x-axis. The bed dips f plot across the body of the graph as the gently to steeply dipping curved
lines. The plot is symmetric across a horizontal line at which percent error = 0, to distinguish
between beds that dip in the same direction of the fault (plotted on lower half of Fig. 8-16) from
beds that dip in the opposite direction of the fault (upper half of plot). On the plot the errors range
from 0 percent to 500 percent.
An examination of Fig. 8-16 shows that unacceptably large errors of greater than 50 percent
are introduced for faults dipping at angles of less than 30 deg (or greater than 150 deg) and for
beds dipping at less than 5 deg (or greater than 175 deg, recognizing that 180 deg is a lower dip
than 175 deg). For faults dipping at 45 deg, bed dips of 10 deg or more can introduce unaccept-
able errors. Notice that if the beds are dipping in the same direction as the fault, then the errors
encountered are correspondingly larger (see lower half of the graph) than a situation in which the
beds are dipping in the direction opposite to the fault. If the bed dip is about equal to the fault
Contouring Faulted Surfaces 353
dip, as is often the case along the flanks of salt domes, then the errors involved in mapping ver-
tical separation as throw can readily exceed several hundred percent.
θ = Fault Dip
Figure 8-16 Graph used to measure the percent error on an incorrectly contoured structure map. Fault and
bed dips are taken to be clockwise from 0 deg to 180 deg. (Published by permission of D. Tearpock and
R. Bischke.)
In Fig. 8-16, the error is relative to the depth of an incorrectly contoured horizon or a hori-
zon mapped as throw. To calculate the error encountered when mapping vertical separation as
throw, consider a bed dipping at 135 deg (or 45 deg to the west) into a fault that dips at 45 deg
to the east. On the plot, a 45-deg fault dipping to the east is located along the 0 percent error line
at 45 deg. From this value, project a vertical line from the x-axis into the upper portion of the
plot. This portion of the plot defines beds that dip in the direction opposite to the fault. A bed
dipping at 135 deg is located between the gently sloping 130-deg to 140-deg dipping lines (see
X on plot). The percent error in incorrectly mapping a vertical separation as throw can now be
determined by projecting the bed and fault dip intersection at X over to the y-axis.
Thus, a bed that dips at 135 deg (or 45 deg to the west) into a fault dipping at 45 deg (to the
east) will have an error of 50 percent, which means that the correct depth to the bed under con-
sideration is 50 percent, or one-half the vertical distance away from the incorrect level. Thus, rel-
ative to the correctly contoured bed, the error relative to the depth that should have been correct-
ly contoured as vertical separation is
or
(1 – 0.5)
= a 100% error
0.5
From this discussion, we can conclude that the errors involved in substituting throw for ver-
tical separation are larger than most interpreters may realize or be willing to accept. Equations
(7-1) and (8-1), and the graphs in Figs. 8-14 and 8-16, are powerful tools that can be used on a
daily basis to help generate correctly contoured maps and to evaluate the maps of others. Such
analysis can be routinely conducted when evaluating prospect maps. Use Fig. 8-16 to analyze the
magnitude of error in the incorrectly contoured map in Fig. 8-13, in the area of the proposed loca-
tion.
For petroleum geologic mapping, you might consider this discussion on the correct tech-
nique for mapping throw as more of an academic exercise than one having some practical impor-
tance. But remember that in mining geology and outcrop mapping, where the actual fault surface
can be touched and throw values physically measured, this mapping technique is valid. You
should, however, consider one important point. If the fault data consist of both actual throw
measurements from a mine or an outcrop, and well log fault data, they cannot be used together in
structure mapping because the well log fault data represent vertical separation. Equation (7-1),
however, can be used to convert the throw data to vertical separation, or vice versa, which can
then be used for mapping.
Only if the dip of the horizon being mapped is zero or nearly so, or if the fault is vertical,
are the values for vertical separation and throw essentially the same. In these cases, they can be
used together in fault and integrated structure mapping.
Later in this chapter we will examine a petroleum-related generic case study, which further
illustrates the importance of mapping the missing section correctly as vertical separation rather
than as throw in petroleum subsurface mapping. It can make the difference between drilling a
successful well and a dry hole.
across a reverse fault. Reverse faults and overthrusts, however, produce a fault overlap rather than
a fault gap. Figure 8-19 shows the technique for contouring across a reverse fault. The 4500-ft
Horizon, which dips generally to the west-northwest, is displaced by a southwest-dipping reverse
fault. The vertical separation or repeated section determined from well log correlation is 450 ft.
The technique for contouring across a reverse fault is usually easier than that for a normal
fault because there is no projection of contours through a fault gap. With a reverse fault, the hang-
ing wall is thrust up and over the footwall, resulting in an overlap of structural horizons.
Therefore, the hanging wall is contoured right up to the hanging wall cutoff at the upthrown fault
trace. Likewise, the footwall is contoured up to the footwall cutoff at the downthrown trace of the
fault. Since the fault blocks overlap, the strike direction of the contours established in the block
that is contoured first (the block with the most control) serves as a guide to the contouring of the
other fault block. As with the normal fault example, to guide the strike direction of the contours,
consider how the structure would be contoured if the fault were not there. Wells positioned in the
fault overlap serve as a guide to the contouring of both fault blocks.
Referring back to Fig. 7-29 in Chapter 7, the thickness of the repeated section or vertical sep-
aration resulting from a reverse fault can be calculated by measuring the vertical distance from
the top of the mapped horizon in the hanging wall to the top of the same horizon in the footwall.
Notice that Well No. 2 in Fig. 8-19, in the fault overlap, has penetrated the top of this horizon
twice: the first time at 4700 ft and the second time at 5150 ft. The vertical difference between
these two tops is equivalent to the repeated section or vertical separation, which is equal to 450
ft. The vertical separation can therefore be seen directly in the fault overlap.
There is one problem with the construction of a reverse fault overlap; it results in significant
clutter, which can be confusing. Some confusion is eliminated by dashing the contours on the
footwall beneath the fault, within the zone of fault overlap, as illustrated in Fig. 8-19. Another
method of eliminating clutter is to pull the fault blocks apart and present each fault block sepa-
rately. This is a good way to construct a structure map with a reverse fault, especially if hydro-
carbons are present and isochore maps are required. The method is shown later in this chapter.
Normal Faults
The step-by-step method of integrating a fault surface map of a normal fault with a structural
horizon map is illustrated in Figs. 8-20 and 8-21. Figure 8-20 is a fault surface map constructed
from fault data from the wells shown on the figure. The fault strikes generally north-south, is
slightly convex to the east with a dip of 45 deg, and has a vertical separation of 400 ft. The stra-
tum being contoured is called the 7000-ft Horizon. The subsea tops for this horizon in each well
are shown on the partly completed structure map in Fig. 8-21a. In Well No. 3, for example, the
top of the 7000-ft Horizon is at a depth of –7045 ft.
The 7000-ft Horizon is cut by Fault A. A review of the depth of the fault picks and the struc-
ture tops for each well indicates that five of the seven wells are in the upthrown block and that
only Wells No. 4 and 5 are in the downthrown fault block. Employing the general contouring
guideline of beginning the structure contouring in the area or fault block with the most control,
the contouring of the 7000-ft Horizon originates in the upthrown fault block. The initial con-
touring indicates an anticlinal structure with a slightly elongated east-west axis (Fig. 8-21a). The
contours are not continued across the entire map area because at some point the contours in this
upthrown fault block intersect Fault A and terminate at the upthrown fault trace. By underlaying
the fault surface map beneath the structure map, the approximate location of this intersection can
be determined and the contouring stopped in the vicinity of this intersection.
When doing this method by hand, the next step is to overlay the structure map (Fig. 8-21a)
onto the fault map (Fig. 8-20), as shown in Fig. 8-21b, to determine the upthrown trace of the
fault. The upthrown trace occurs where structure contours in the upthrown block intersect fault
contours of the same elevation. These intersections are highlighted on the map by a small mark
placed at the end of each contour line, as shown in the figure. Because the fault surface map has
a contour interval of 1000 ft and the structure contour map has a 200-ft contour interval, the pre-
cise position of the intersections for each contour line should be located by interpolation, using
10-point proportional dividers or a scale. Dividers or a scale allow subdivision of the 1000-ft fault
contours into 200-ft intervals anywhere on the map without the clutter of actually drawing the
extra contours (see Fig. 2-11). For a small portion of the fault map, 200-ft contours are shown
between the –7000-ft and –8000-ft depths. These contours serve two purposes: (1) to illustrate
the accuracy of the intersection of the fault and structure contours for the –7000-ft, –7200 ft,
–7400-ft, –7600-ft, –7800-ft, and –8000-ft elevations; and (2) to show that the construction of
200-ft fault contours over the entire map would result in unnecessary clutter. Two-hundred-foot
(200-ft) contours can, however, be placed as marks in localized areas, as shown in the figure, to
determine the intersections precisely. Once used for this purpose, the additional contours are then
358 Chap. 8 / Structure Maps
Figure 8-20 Fault surface map for Fault A constructed from well log fault data from seven
wells. The fault map has a 1000-ft contour interval.
(a)
Figure 8-21 (a) Partially completed structure map on the 7000-ft Horizon. (b) Integration of
the fault and structure maps to identify the intersection of the fault surface with the upthrown
fault block of the 7000-ft Horizon. (c) Structure map shows the delineation of the upthrown
trace of Fault A and the projection of form contours into the downthrown fault block. (d)
Integration of the fault and structure maps to identify the intersection of the fault surface with
the downthrown fault block of the 7000-ft Horizon. (e) The final, integrated structure map for
the 7000-ft Horizon.
Manual Integration of Fault and Structure Maps 359
(b)
PREFAULTED MODEL
(c)
(d)
(e)
erased. After all the intersections have been identified, the upthrown trace of the fault is accu-
rately constructed simply by connecting all the marks with a smooth line, as shown in Fig. 8-21c.
The next step is to project the structure contours through the upthrown trace of the fault into
the downthrown fault block. There are two points of control in the downthrown fault block that
must be honored: the horizon elevation in Well No. 4 is –8420 ft, and in Well No. 5 it is –8130
ft. Earlier in this chapter we presented the technique for projecting contours from one fault block
across a fault into the adjacent fault block. This technique is used to complete the construction of
the anticline in the downthrown fault block and delineate the downthrown trace of the fault. An
easy way to remember the technique to guide your contouring through the fault is to consider
how the structure would be contoured if the fault were not there. One method is to restore the
faulted block to an unfaulted condition and relabel the depth values in the wells to account for
the restoration. The unfaulted structure can then be mapped. Once completed, the map can be
refaulted or restored again to its faulted condition. The structure contours must be adjusted again
in consideration of the vertical separation. This is a time-consuming task when done by hand,
whereas computers are more capable of using such methodology. For simplicity, often it is best
to project the contours into the opposing fault block as form contours (contours without values)
in order to complete the structural picture, as shown in Fig. 8-21c.
Once contours are projected through the fault into the downthrown block, their values are
adjusted from those in the upthrown block by the amount of vertical separation, which in this case
is 400 ft. Thus, the projection of the –7400-ft contour from the upthrown block become a –7800-
ft contour in the downthrown block. Continue this procedure for all contours projected through
the fault. In Fig. 8-21d, the downthrown contours have been assigned structural elevation values
and now the downthrown trace of the fault can be determined. Again, the fault surface map is
placed under the structure map and the intersections of the structure contours in the downthrown
block with the fault contours of the same elevation are identified and indicated by small marks.
Finally, connect all the marks with a heavy smooth line to accurately delineate the down-
thrown trace of the fault. Place a symbol on the downthrown trace to show the direction of fault
dip, erase the contours in the fault gap, and the integrated structure contour map for the faulted
anticlinal structure is complete (Fig. 8-21e). By correctly integrating the fault and structure maps,
we have (1) accurately delineated the position of the fault trace on the structure for a particular
horizon; (2) precisely constructed the upthrown and downthrown traces of the fault; (3) estab-
lished the actual width of the fault gap; and (4) projected the structure contours correctly across
the fault. An understanding of this technique is paramount to the correct and precise integra-
tion of a fault map and a structure map; therefore, take the time to review and master this
process.
For the example shown in Fig. 8-21, the process was relatively easy because the structural
pattern is a simple anticline cut by only one fault. But the procedure is basically the same for a
more complex structure and pattern of faults. Figure 8-22 shows the complexly faulted anticlinal
structure of an oil and gas field. A fault surface map was prepared for each of the 13 individual
normal faults, and the integration technique was used to prepare the completed structure map.
With a complex structure such as the one shown here, the logistics are more involved and it takes
more time to prepare the fault maps and integrated structural interpretation, but the methodology
is basically the same.
With respect to the workstation environment, a number of the individual software packages
do not allow a geoscientist to integrate faulted surfaces and horizons as shown here. This type of
accuracy is necessary, however, if you wish to generate a valid integrated interpretation, minimize
the drilling of dry holes or positioning of wells in the wrong subsurface location, and accurately
362 Chap. 8 / Structure Maps
Figure 8-22 An integrated structure map of a very complexly faulted anticlinal structure. Each
fault was integrated with the structural interpretation as shown in Fig. 8-21.
estimate hydrocarbon volumes. The details of how to accomplish this method of integration with
the various workstation programs are beyond the scope of this book. Chapter 9 on 3D seismic
interpretation provides some guidance for this method in a general sense. Each geoscientist must
learn the individual mapping program being used and determine the best method for combining
the use of applications within each package in order to achieve this type of subsurface integra-
tion accuracy.
(DT) is –6300 ft; for Well No. 7 the UT is –6235 ft and the DT is –6535 ft. The vertical differ-
ence between DT and UT restored tops should be equal or nearly equal to the section missing in
the wells due to the fault. In this case, the difference is 300 ft, which is equal to the missing sec-
tion determined by log correlation; consequently, the restored tops appear consistent with the
available data.
Procedures for projecting contours through a fault and the integration of a fault map with a
structure map have been discussed, so we can review the structure map in Fig. 8-10 and deter-
mine the amount of missing section. A darkened circle marks the intersection of each structure
contour with the fault surface contour at the same elevation. By connecting these marks, the
upthrown and downthrown fault traces are delineated as shown on the map. Any contour may be
taken from its intersection with the upthrown fault trace and projected through the fault to the
downthrown trace to estimate the vertical separation for the fault at this horizon. For example, the
–6100-ft contour in the upthrown block projected through the fault gap intersects the –6400-ft
contour in the downthrown block, indicating the fault displacement (vertical separation) of 300
ft, which was based on the restored tops and used in the mapping. The restored tops in Wells No.
2 and 7 provide important control points for contouring the structure map.
Restored tops are located in a vertical well itself (Fig. 4-42), or vertically above and below
the fault pick in a deviated well (Fig. 4-43). They are not placed at the upthrown and downthrown
traces of a fault. Using Fig. 8-10 as an example, the restored tops for Wells No. 2 and 7 are placed
right at the well locations. The dashed lines through each of the two wells illustrate the honoring
of the restored tops in the wellbores. In Well No. 7, an interpolated contour of –6235 ft projects
from the upthrown fault block into the well, and an interpolated contour of –6535 ft projects from
the downthrown fault block into the well. Similarly for Well No. 2, the –6000-ft contour in the
upthrown block and the –6300-ft contour in the downthrown block project into the well.
How are these restored tops used as an aid in structure mapping? Depths for restored tops
are honored in the same way as any other well control point during structure mapping. For exam-
ple, the UT restored top for Well No. 2 is –6000 ft; therefore, the –6000-ft structure contour in
the upthrown block honors this control point at Well No. 2. Likewise, the –6300-ft contour in the
downthrown fault block is projected through the fault gap to intersect with Well No. 2, whose DT
restored top is –6300 ft. Each UT and DT restored top provides two additional points of control
to aid in structure contouring in and around a fault. The contouring of the structure map in Fig.
8-10 confirms that the four restored tops were used as control points in the construction of the
final structure map.
In areas of limited well control, restored tops in faulted wells provide significant structural
information which is often necessary for the preparation of a realistic and accurate structure map.
Considering the well control in Fig. 8-23a, how would you contour these data points? The data
can be contoured in a number of ways. Figure 8-23b shows one interpretation which appears to
be reasonable. The map does honor all the established well control and was used to propose the
two drilling locations shown on the map. The first location is upthrown to Fault C in Reservoir
C-3. An oil show in Well No. 14 establishes the down-dip limit of oil at –9245 ft. The proposed
well is designed to penetrate the reservoir in the optimum position for maximizing the well’s
drainage efficiency. Based on this interpretation, the volume of anticipated recoverable oil up-dip
of the oil/water contact is 6480 acre-feet. Considering a reasonable recovery factor for this area
of 450 barrels per acre-foot, this prospect represents 2,916,000 barrels of potentially recoverable
oil. Downthrown to Fault C, a development location is proposed to maximize the drainage effi-
ciency for the east-west elongated Reservoir C-5. Based on this structural interpretation, the vol-
ume of estimated reserves is 1,107,000 barrels of recoverable oil.
364 Chap. 8 / Structure Maps
(a)
(b)
(c)
Figure 8-23 (a) Base map with posted data for the 9000-ft Horizon. How would you contour the data?
(b) Structural interpretation of the 9000-ft Horizon using all the well data except for the restored tops in
Wells No. 3 and 11. Two development wells are proposed based on this interpretation. (c) Revised struc-
tural interpretation of the 9000-ft Horizon using all the available well data including the restored tops for
Wells No. 3 and 11. Compare this interpretation with that shown in Fig. 8-23b.
Manual Integration of Fault and Structure Maps 365
The interpretation appears reasonable except for the two wells within the fault gap in which
the 9000-ft Horizon has been faulted out. For whatever reason, no restored tops were estimated
for these wells to incorporate into the structural interpretation. Figure 8-23c is a structure contour
map for the same two reservoirs using the UT and DT restored tops for the 9000-ft Horizon in
Wells No. 3 and 11 in the interpretation. It is obvious that the four restored tops are very impor-
tant data needed to develop a more accurate interpretation of this structure. The addition of the
restored top data has a significant impact on the interpretation of the overall geometry of the
structure and the proposed well locations. For the upthrown Reservoir C-3, the volume of recov-
erable hydrocarbons has been reduced to 3480 acre-feet or 1,566,000 barrels of recoverable oil,
a reduction in volume of 46 percent. The map for Reservoir C-5 in the downthrown block, using
the restored tops, indicates that the volume of Reservoir C-5 is smaller by 41 percent compared
to the map in Fig. 8-23b. Potentially recoverable hydrocarbons decrease from 1,107,000 barrels
to 648,000 barrels. More importantly, the proposed development location for Reservoir C-5 is
actually down-dip of the oil/water contact, and if drilled the well would result in a dry hole.
Information provided by the UT and DT restored tops was critical in this example. The tops
are valuable mapping data that should not be ignored, particularly in areas of limited well control.
These extra data points guide the structure contours into and through the fault gap. Figure 8-23
makes this point very clear. Upthrown and downthrown restored tops improve the accuracy of a
structure contour map, and in this case reduced the size of the two prospective reservoirs.
Restored tops are also helpful in generating prospects where none previously existed. Figure
8-24a shows a portion of a structure contour map showing a potential oil reservoir. Notice Well
No. 22 in the western portion of the map, upthrown to Fault A, has an oil show with an oil/water
contact at 6458 ft TVD (true vertical depth). Well No. 25 to the east is wet and the mapping hori-
zon is faulted out in Well No. 65 by Fault A. Based on the structural interpretation shown in this
figure, the volume of potentially recoverable oil is insufficient to justify the drilling of an up-dip
development well into Reservoir A-1. The structural closure is too small, and the target area
between the oil/water contact and the upthrown trace of the fault too small to risk the drilling of
a well. During the preparation of this structure map, the trace of Fault A was constructed based
on isolated fault data without the use of a fault surface map and without estimated restored tops
for the faulted out Well No. 65.
The structure map in Fig. 8-24b was constructed using a fault surface map for Fault A to
integrate with the structure and to contour across the fault. The UT and DT restored tops for this
horizon in Well No. 65 were calculated and used, and maps on shallower and deeper horizons, in
which the tops in Well No. 65 were present, were also made. These additional mapping data
change both the configuration of Reservoir A-1 and the volume of potentially recoverable hydro-
carbons. This accurately integrated structure contour map delineates a reservoir of sufficient size
(569,000 barrels of oil) and closure to be reevaluated as a potential development location. The
UT and DT restored tops are used in Fig. 8-24b to guide the strike direction of the contours into
the fault and to honor the vertical separation, thus changing the structural configuration of the
reservoir.
We cannot overemphasize the importance of using all available data when preparing sub-
surface maps of any type. Geoscientists always work with limited data, so to ignore or not use
valuable data, for any reason, is unthinkable. Our job is to develop the most reasonable and accu-
rate interpretation possible from available data. The use of upthrown and downthrown restored
tops is a necessary part of preparing a structure contour map. The time required to estimate these
tops and incorporate them into the integrated structural picture is minimal, but it can have a sig-
nificant impact on the final interpretation, as shown in the examples presented.
366 Chap. 8 / Structure Maps
Figure 8-24 (a) Structure map on the 6450-ft Horizon shows Reservoir A-1 of insufficient size to justify a
development well. This map was constructed without the integration of a fault map, and the restored tops
for Well No. 65 were not used. (b) Revised structure map on the 6450-ft Horizon, Reservoir A-1. The revised
interpretation was prepared using the restored top data from Well No. 65 and the integration of the fault
surface map for Fault A with the structural interpretation. Also, shallower and deeper structure maps were
used to guide the structure contouring. Reservoir A-1 is now of sufficient size to propose a development
well location.
Reverse Faults
The method of manually integrating a reverse fault surface map with a structure contour map is
illustrated in Figs. 8-25 and 8-26. We do not describe the procedure in as much detail as we did
for the integration of normal faults, since the procedure is basically the same. There are seven
wells from which fault and horizon data were obtained. Figure 8-25 shows the reverse fault sur-
face map constructed using the data from the seven wells. The fault strikes generally north-south
with a dip of 35 deg to the west and has a displacement (vertical separation) of 300 ft, as shown
by the values of repeated section in the well data.
The technique for integrating a reverse fault surface map with a horizon map is often easier
than that for a normal fault because there is no projection of contours through a fault gap. In the
case of a reverse fault, the hanging wall is thrust up and over the footwall, resulting in an over-
lap of structural horizons. Therefore, the hanging wall is contoured up to the hanging wall cutoff
at the upthrown trace of the fault, and the footwall is contoured up to the footwall cutoff at the
downthrown trace of the fault. Since the fault blocks overlap, the strike direction of the contours
established in one fault block can often be used to guide the direction of the contours in the
opposing block.
The faulted structure for this example (Fig. 8-26) is the nose of a plunging anticline that is
penetrated by seven wells. Wells No. 2, 3, and 6 penetrated the footwall, Wells No. 5 and 7 pen-
etrated the hanging wall, and Well No. 1 is in the fault overlap, thereby penetrating the horizon
twice – first at 4760 ft in the hanging wall, and second at 4460 ft in the footwall. Notice that the
contour values are positive, indicating this structure is above sea level.
Manual Integration of Fault and Structure Maps 367
We begin contouring this structure in the hanging wall fault block. Place the fault surface
map under the structure map as shown in Fig. 8-26a. The fault trace on the hanging wall occurs
where the structure contours in the hanging wall intersect fault surface contours of the same ele-
vation. These intersections are highlighted on the map as small marks placed at the end of each
contour. Likewise, as we contour the footwall horizon, the fault trace is located where the struc-
ture contours in the footwall intersect fault contours of the same value. Again, these intersections
are highlighted by small marks placed at the end of each contour (Fig. 8-26a). For accuracy,
10-point dividers or a scale may be used to locate the exact points of intersection between struc-
ture and fault contours.
The difference in elevation between contours in the footwall and those in the hanging wall
equals 300 ft, which is the value of the vertical separation or repeated section determined from
well log correlation. For example, the 4000-ft contour in the footwall within the zone of overlap
is positioned directly under the 4300-ft contour in the hanging wall. This same 300-ft difference
in contour elevation is maintained for each of the structure contours in the mapped area.
Finally, connect all the marks representing the intersections of the fault surface with the foot-
wall and hanging wall structure contours with smooth lines to accurately delineate the fault traces
(Fig. 8-26b). Place a symbol on the hanging wall fault trace to show the direction of fault dip and
dash the structure contours in the footwall, under the fault overlap for clarity. Having completed
these various steps, the integrated structure contour map for the faulted plunging anticlinal nose
is complete.
By correctly integrating the fault and structure maps, the following is accomplished: (1) the
position of the fault on the mapped horizon is accurately delineated; (2) the traces for the fault
are precisely constructed; (3) the actual width of the fault overlap is established; and (4) the con-
tours are carried correctly in the fault overlap. Although the actual fault surface map is striking
north-south, the trace of the fault on this horizon strikes northwest-southeast on the northern flank
368 Chap. 8 / Structure Maps
(a) (b)
(c)
Figure 8-26 (a) The fault map is placed under the structural interpretation to obtain the intersec-
tion of the fault surface with both the footwall and hanging wall fault blocks. (b) Final integrated
structure map for the 5000-ft Horizon. (c) A pull-apart map of the final integrated structural inter-
pretation. This type map separates the footwall and hanging wall fault blocks to show them indi-
vidually, without the clutter of overlapping contours.
Fault Traces and Gaps – Shortcuts and Pitfalls 369
of the structure and southwest-northeast on the southern flank of the structure. The curved shape
and position of the trace for Fault 1 on this structure map could not be intuitively constructed
without the integration of the fault and structure maps.
We mentioned earlier that a structure map with a reverse fault overlap is cluttered and can
be confusing because of the double set of contours in the overlap. One way to minimize the clut-
ter and confusion when contouring by hand is to dash the contours in the footwall in the zone of
overlap (Fig. 8-26b). Another solution to the clutter and confusion is to prepare a pull-apart map,
as shown in Fig. 8-26c. With this map, the two separate fault blocks are pulled apart or mapped
separately to show each block separately without any overlap. Such a map should be construct-
ed only after the fault and structural integration is complete in order to eliminate any possibility
of contouring errors. In this example, the hanging wall is productive of hydrocarbons. Eventually,
net sand and net oil isochore maps will be constructed for the hanging wall reservoir. These maps
can be prepared more easily if the fault blocks are separated as shown in Fig. 8-26c.
see the same effect. The fault cuts this horizon north of the crest and is dipping in the opposite
direction to the mapped horizon; therefore, the fault gap is thinner in this area than on the flank
of the structure. On the flanks, where the fault cuts this horizon south of the structural axis, the
fault is dipping at almost a right angle to the mapped horizon, resulting in a wider fault gap. If a
deeper horizon were mapped, one in which the fault were entirely to the south of the crest of the
structure, the width of the fault gap near the crest would be greater. As a simple exercise, project
the well control to a horizon 1000 ft deeper (e.g., change the depth in Well No. 3 from –5300 ft
Figure 8-27 A fault surface map superimposed on a completed structure map to show that the fault traces
have been determined by the integration of the fault and structure maps. Observe the change in the width of
the fault gap across the structure. The change gives the appearance that the fault is smaller near the crest of
the structure. (Prepared by J. Bollick. Published by permission of Tenneco Oil Company.)
to –6300 ft). Using the contours in Fig. 8-27 as a guide, contour that deeper horizon on tracing
paper and then integrate the fault map with that horizon map. What observations can you make
about the change in width of the fault gap and direction of the fault trace?
In Fig. 8-10, the curvature of the fault trace and width of the fault gap change across the
structure. The gap changes from 350 ft on the flanks to 250 ft near the crest, whereas the fault
has a constant vertical separation of 300 ft. Figures 8-11a and 8-12 show a significant difference
in the width of the fault gap, which ranges from an average 420 ft in Fig. 8-11a to 290 ft in Fig.
8-12, despite the fact that the fault on each map is the same fault, with a vertical separation of
400 ft. Finally, examine Fig. 8-27, which shows a map of a fault surface, striking east-west and
Fault Traces and Gaps – Shortcuts and Pitfalls 371
dipping to the south at 55 deg, superimposed on the structure map of the 5500-ft Horizon. The
upthrown and downthrown fault traces, and therefore the fault gap, were constructed using the
integration technique. Notice how the fault gap thins near the crest of the structure, causing
the vertical separation of the fault to appear smaller on the crest than on the flanks. The vertical
separation of the fault only appears to change across the structure because the width of the fault
gap changes.
The width of a fault gap depends upon the amount of vertical separation and the angles and
directions of the fault dip and the horizon dip. Therefore, the width of a fault gap for any given
fault on a structure map cannot be intuitively determined. It is extremely difficult at best, regard-
less of vertical separation and our ability to see in three dimensions, to predict the position of
fault traces and the width of a fault gap by looking at fault data, or even at a fault surface map
and a structure map independently. Examine Fig. 8-27 again and ask yourself whether you could
have predicted the curvature of the overall fault trace and the variation in width of the fault gap
on this final structure map simply by reviewing the fault surface map and structure map data sep-
arately without integration. Most likely, your answer is no. We emphasize that if the construction
of the upthrown and downthrown fault traces is done properly by integrating the fault and struc-
ture maps, the width of the fault gap is automatically constructed without any need for guesses
or estimations (Tearpock and Harris 1987).
Rule of 45
Figure 8-28 shows a special condition in which the fault traces can be delineated with reasonable
accuracy, without integrating a fault and structure map, and where the fault gap is equal in width
to the amount of missing section in the well. The technique used for this situation is known as
the Rule of 45. If a fault is inclined at 45 deg and the beds are approximately horizontal (zero
deg dip), the position of the traces can be identified with reasonable accuracy without integration.
The width of the fault gap is equal to the amount of vertical separation for the fault, measured off
using the map scale as reference (see map insert in Fig. 8-28).
Consider Fig. 8-29a. It is a prospect map illustrating a hydrocarbon trap upthrown to a fault
and limited down-dip by a water contact in Well No. 3. The interpretation is based on both well
log and seismic data. The completed structure map shows the horizon tops, fault data from the
wells, and the subsea depths of the fault at various shotpoints on the seismic lines. The prospect
map with a proposed location looks very attractive at first glance.
Without detailed knowledge of how the interpretation was generated, management might
approve the drilling of such a location. However, upon detailed investigation, it is learned that the
completed trace for Fault B was prepared using the position of the fault on the horizon from
the two seismic lines and the Rule of 45 to position the trace relative to the well locations. Does
the application of this method provide a reasonable interpretation in this example? The answer is
no. One major structural feature not considered in using the Rule of 45 was the structural dip: the
horizon is far from flat. As we have shown, bed dip is one of the structural features that affects
the position of a fault on a completed structure map.
The structure map shown in Fig. 8-29b was prepared using the same structural and fault data.
In this case, however, a fault surface map was first constructed (Fig. 8-29c) and then integrated
(cross-contoured) with the structure map to arrive at the correct positioning of the upthrown and
downthrown fault traces.
The correct integration illustrates that the original prospect map has an enormous error
regarding the position of the fault and the prospect potential. This error would result in the pro-
posed well being drilled downthrown to Fault B, missing the oil reservoir entirely. How often can
a petroleum company afford such mistakes? A number of prospect maps reviewed worldwide
have shown similar mistakes resulting in costly dry holes. And in one particular case, one com-
pany drilled two such dry holes in one year. The company’s CEO stated that no additional
prospects would be drilled without constructing fault surface maps and integrating the fault with
the structural interpretation.
1. The horizontal distance between the well and the fault trace on the horizon equals the
cotangent of the dip angle of the fault multiplied by the vertical distance between the fault
pick and the horizon in the well (Fig. 8-30b).
2. The well serves as the center point of a circle, the radius of which equals the distance
between the fault and the well as determined in step 1 (Fig. 8-30c).
3. A tangent common to two or more circles determines a fault trace. The exterior tangent is
the fault trace for wells on the same side of the fault; the interior tangent is the fault trace
for wells on the opposite sides of the fault (Fig. 8-30c). Refer again to Chapter 7, Fig. 7-9,
and the accompanying text for determining in which fault block a well is located for any
mapped horizon.
4. Finally, contour the horizon tops assuming continuity of structural attitude across the faults
(Fig. 8-30d).
If the beds are horizontal, the fault gap width can be determined as shown in Fig 8-30b. Then
the fault gap can be drawn on the structure map rather than representing the fault as one line, as
Fault Traces and Gaps – Shortcuts and Pitfalls 373
PROSPECT
Depth of Fault on ACREAGE
Seismic Line
Seismic
-6000'
eA
-8205'
4 -6000'
450'/-6350'
Lin
-8000'
Line B
mic
-7000'
-7800' -8025'
500'/-6760'
-7600'
Seis
3
-7400'
-7000'
1
-8220' -7200'
475'/-7840'
-8000' -7000'
-8000'
P.L.
-7 0'
-78 00'
-7
80
-76 0'
60
-80
-7
0
-74 00'
0'
6
40
'
-7
00
0'
20
14
'
-6870'
00
0'
-7
-72
00
0'
2
-68
00
-6920' '
10
-660
8 -6740'
0'
7 5
-6722'
GAS -6855'
9
-7010' OIL
0 1000'
Scale
(a)
Figure 8-29 (a) Prospect map based on well log and seismic data. Position of the fault trace based on the
Rule of 45 relative to the well locations and on the fault position as interpreted on the seismic profiles.
(b) Reinterpretation of prospect map in (a), based on integration of the fault surface map with the horizon map.
Fault trace moves about 900 ft north. (c) Fault surface map. (Published by permission of D. Tearpock.)
shown in Fig. 8-30d. This method of estimating the fault trace is useful if the dip of the fault sur-
face can be determined or at least assumed with some degree of certainty and if the horizon being
mapped is approximately horizontal. Compare the mapping results in Fig. 8-30d with those
shown in Fig. 8-30e and f. Figure 8-30e is a fault map for the two faults and Fig. 8-30f shows the
374 Chap. 8 / Structure Maps
Seismic
-6000'
A
4
Line
-8205'
-8000'
-6000'
Line B
450'/-6350'
mic
-7000' -7800' 3 -8025'
-7600'
Seis
500'/-6760'
-8220' -7000'
475'/-7840' 1
-8000'
P.L.
-8000'
'
00
-8 '
-7 '
00
-7 0'
-7
-7 '
-7
4
-7 0'
80
00
60
-8
-7 0'
40
20
2
80
0'
0'
-8
60
0
0'
0
0'
40
20
-7
0 1000'
(b) Scale
0 1000
Scale
(c)
fault traces constructed by the integration of the fault and structure maps. The position of the fault
in Fig. 8-30d compares very favorably with the position of the fault in Fig. 8-30f, showing that
in an area of relatively flat beds, the tangent method can provide a good approximation for posi-
tioning the fault on a structure map. However, we caution that this technique does not work in
areas of dipping beds.
In Fig. 8-30c or d, the fault traces are constructed as straight-line segments with sharp angu-
lar changes in strike direction. In order to prepare a more accurate map, the change in strike direc-
tion can be smoothed to best fit the overall pattern of the faults and the data in the area being
mapped. The fault surface map for Fig. 8-30e and the integrated structure map in Fig. 8-30f were
Fault Traces and Gaps – Shortcuts and Pitfalls 375
(a) (b)
Figure 8-30 (a) Base map with the depth to fault picks (F-3100 ft) and sand tops (ss –3320 ft) posted.
(Modified after Bishop 1960; published by permission of author.) (b) Determination of the horizontal distance
from a well to the nearest fault trace using the “Cotangent Rule” and the width of the fault gap. (c) Assumed
angle of fault = 60 deg. The position of the fault trace is determined by the “Circle Method.” (Modified after
Bishop 1960. Published by permission of author.) (d) Structure contour map. Position of the fault determined
by the “Circle Method.” (e) Integration of mapped horizon and fault surface maps to determine the position and
width of the fault gaps. (f) Final integrated structure map. Compare position of fault traces with those shown in
Fig. 8-30d. (From Bishop 1960. Published by permission of author.)
also constructed as straight line segments to make the results comparable with those in Fig.
8-30d.
The tangent method is not accurate if the horizon being mapped has significant dip (more
than 5 deg) because the wellbore is then not perpendicular to the bedding. The correct construc-
tion of the fault traces and fault gap for faults on a dipping structure is especially critical because
the fault position has a significant impact on any proposed location for an exploration or devel-
opment well. Figure 8-31a shows a portion of a structure map on the flank of a piercement salt
dome. The depth to the mapped horizon and the missing section and depth of the faults are shown
next to each well. Since the fault is a perfect plane with a dip of 45 deg, the structure map in Fig.
8-31a was prepared (incorrectly) using the Rule of 45 to determine the position of the fault traces
and the width of the fault gap. As an example, look at Well No. 6 shown in the figure insert. The
fault cuts the well at –9510 ft and the top of the horizon intersects the well at –9395 ft. From the
figure insert we see that Fault A crosses the well 115 ft below the mapped horizon. Therefore,
based on the 45-degree trigonometric relationship of 1 to 1, the downthrown trace of the fault was
located 115 ft west of the location of Well No. 6 as the horizon in the well is in the downthrown
376 Chap. 8 / Structure Maps
(c) (d)
block. Also, since the fault has a dip of 45 deg, the width of the fault gap was scaled to equal the
missing section of the fault (400 ft). Using this method, the fault cut in Well No. 3 does not fit as
part of the same fault identified in Wells No. 2, 5, 6, and 10; therefore, a second fault, Fault B,
was postulated as shown on the figure, resulting in an untested fault trap up-dip of Well No. 5.
The fault in this example has a constant dip of 45 deg and, in this respect, the above meth-
ods were appropriate. However, the beds are not horizontal; they dip quite steeply (about 40 deg
in the up-structure position). Therefore, neither the rule of 45 nor the tangent method is appro-
priate for the preparation of the structure contour map in Fig. 8-31a. Figure 8-31b shows the fault
surface map and the correctly contoured structure map for this horizon. The fault is a peripheral
fault paralleling the face of the salt; the beds are dipping at about 40 deg at –7000 ft and flatten
to about 32 deg between –11,000 ft and –12,000 ft. Since the beds are steeply dipping, the fault
map was required for integration of the fault and structure maps to delineate accurately the fault
traces on the structure map. The upthrown and downthrown traces of the fault, and therefore the
resulting fault gap, were constructed using the integration technique discussed earlier in the chap-
ter.
Observe that although the fault has a constant vertical separation of 400 ft (as determined
from well log correlation), and the fault is a perfect plane dipping at 45 deg, the trace of the fault
(a) (b)
Figure 8-31 (a) Portion of a structure map on a steeply dipping horizon on the flank of a piercement salt
dome. The map was incorrectly prepared using the Rule of 45 to position the fault traces. Although the
fault is dipping at 45 deg, the Rule of 45 is not applicable because the beds are not horizontal. (b) Correctly
contoured structure map based on the integration of the structural interpretation with the fault surface
map. Compare this map to that shown in Fig. 8-30a.
378 Chap. 8 / Structure Maps
is curved and nowhere along its trace is the gap 400 ft wide. Measuring across the fault gap at the
location where the –8000-ft structure contour in the upthrown block intersects the upthrown trace
of the fault, the width of the fault gap is approximately 1700 ft. At the intersection of the
–10,000-ft structure contour and the upthrown trace of the fault, the width of the fault gap has
decreased to 1200 ft. If a well were drilled, on the basis of the map in Fig. 8-31a, in the “untest-
ed fault block,” then the mapped horizon would probably be faulted out of that well (as shown by
the correct map in Fig. 8-31b). Study this figure very carefully, for it shows how important it is
to integrate the fault and structure maps on steeply dipping structures to accurately position the
fault traces and delineate the fault gap.
As a general rule of thumb, we consider any structure with a dip greater than 5 deg as steeply
dipping for purposes of fault and structure map integration. However, to prepare a structure map
as accurately as possible regardless of the structural dip, the integration technique should always
be used to delineate the position of the fault traces and the width of the fault gap.
Figure 8-32 shows three different sets of integrated fault and structure contour maps that
result in significantly different fault traces and gaps. For example, the set on the left side of the
figure consists of completed structure maps A(1), B(1), and C(1), all of which use the same
oblique down-dip fault with 100 ft of vertical separation to integrate with the structure. A(1)
shows a horizon with a dip that is constant, B(1) shows a steeply dipping horizon which increas-
es in dip in the upstructure direction, and C(1) is similar to B(1) except that the horizon dip is
less. Despite the constant fault surface in all three cases, each integrated structure map results in
a strikingly different fault trace pattern and fault gap width. Take a few minutes to review the dif-
ferent fault traces and gaps for sets 2 and 3 in Fig. 8-32.
The nine separate patterns shown in Fig. 8-32 and those in Figs. 8-10, 8-26, 8-27, and 8-31
reinforce our statements that the final fault traces and fault gap for any particular set of fault and
structural conditions are not intuitively predictable and, in fact, are often impossible to predict
without actually integrating the two surfaces. The precise positioning of the fault traces and delin-
eation of the width of the gap or overlap (in the case of a reverse fault) can mean the difference
between generating and not generating a prospect, the difference between drilling a successful
well versus a dry hole, or the difference between a reservoir whose calculated volume matches
its performance versus a reservoir that appears to either have overproduced or underproduced the
volumetric estimate of reserves. Taking shortcuts to identify the positions of fault traces and to
determine the widths of fault gaps is a risky business unless the beds are relatively flat-lying and
the fault dips are very close to 45 deg.
(A)
(B)
(C)
Figure 8-32 Three different sets of integrated fault and structure maps illustrating significantly differ-
ent fault traces. (Published by permission of Tenneco Oil Company.)
380 Chap. 8 / Structure Maps
sin ( π / 2 – θ) sin (θ + π – φ a )
=
DX AX
and
cos ( π – φ a ) = BD / DX
Therefore,
sin ( π / 2 – θ) sin (θ + π – φ a )
=
BD / cos ( π – φ a ) AX
Fault Traces and Gaps – Shortcuts and Pitfalls 381
or
AX cos ( π – φ a ) sin ( π / 2 – φ a )
BD =
sin (θ + π – φ a )
Therefore,
– AX cos φ a cos θ
Radius = (8-2)
sin (θ – φ a )
where
– AX cos ( π ) cos θ
Radius =
sin θ
AX cos θ
=
sin θ
Radius* = AX cotan θ
Example
θ = 30 deg east
φa = 20 deg west or 180 deg – 20 deg = 160 deg
AX = 1000 ft
*This is the equation used for the tangent method where the beds are horizontal.
382 Chap. 8 / Structure Maps
sin ( π / 2 – φ a ) sin (θ + π – φa )
=
DE AE
and
cos ( π – φ a ) = BD / DE
BD
DE =
cos ( π – φ a )
Therefore,
AE sin ( π / 2 – θ) cos ( π – φ a )
BD (Heave) =
sin (θ – φ a )
Example
θ = 60 deg
φa = 45 deg
AE = 1000 ft
BD = 353.5 / 0.2588
BD = 1366 ft
Structure Map – Generic Case Study 383
In order to use these equations, the strike direction of the fault surface must be known. These
equations should be used only as a quick look or check, and do not take the place of the
fault/structure integration technique.
correctly the displacement resulting from a fault. The displacement to which we refer is vertical
separation, as determined from well log correlations or seismic data. In the same section, the
incorrect technique of contouring displacement as throw is also presented. Now we look at a
generic case study which illustrates the impact of preparing a map with the incorrect, as opposed
to correct, contouring technique.
This field example shows an anticline cut by a large 600-ft down-to-the-south normal Fault
A and a smaller 100-ft antithetic (compensating) normal Fault B (Fig. 8-36). The structure map
in Fig. 8-37a was prepared with the incorrect assumption that the missing section in the wellbores
is equal to throw (note the dashed contours in the fault gap drawn perpendicular to the strike of
the fault). Based on this interpretation, two development wells are proposed. Location X,
upthrown to Fault A in Reservoir A, is estimated to penetrate the reservoir up-dip and to the east
of Well No. 12 at a depth of –4960 ft. The purpose of the well is to improve the drainage effi-
ciency of this reservoir. Location Y, upthrown to Fault B, is estimated to penetrate Reservoir B at
Figure 8-36 Fault surface map (Faults A and B) for the generic case study. (Published by per-
mission of Tenneco Oil Company.)
Structure Map – Generic Case Study 385
a depth of –5360 ft and is designed to be drilled in the optimum structural position to drain the
attic reserves in this reservoir.
The structure map in Fig. 8-37b was constructed correctly, using the missing section in the
wellbores as vertical separation. Observe the dashed contours in the fault gap showing the trend
projection of the contours from one fault block to the next. This correct interpretation shows a
different structural picture. First, Well X is estimated to penetrate the formation at a subsea depth
of –5030 ft, which is the depth of the oil/water contact in Reservoir A. Therefore, Well X, if
drilled, would probably be a dry hole. The penetration point for Well Y, in Reservoir B, is esti-
mated at –5415 ft, or 55 ft deeper than that shown in Fig. 8-37a. Based on the correctly contoured
map, the proposed well is about 800 ft west of the optimum position to efficiently drain the
remaining reserves in Reservoir B.
The correctly contoured map has the following impact on these two reservoirs and proposed
wells.
The magnitude of error created by incorrectly mapping vertical separation as if it were throw
is greatest near or on the crest of a structure, as illustrated in this generic case. On both the east
and west flanks of the structure, where the strike direction of the contours is nearly perpendicu-
lar to the strike of the fault, the projection of contours through the fault is almost identical for
both structure map interpretations. Where the strike direction of the structure contours is perpen-
dicular to the strike of the fault, lines parallel to the structure contour strike, or a cross section in
this orientation, give the appearance that the formation has an apparent dip of zero degrees (hor-
izontal). Therefore, whether the contours are projected through the fault as if the fault were not
there (mapping vertical separation), or the strike direction of the contours is projected through the
fault perpendicular to the strike of the fault (mapping throw), the direction of contour projection
through the fault is similar. However, as the crest of the structure is approached, the projection of
the contours is radically different for each structure map. Review the –5000-ft contour starting in
the upthrown block of Fault A and follow the projection across the fault into the downthrown fault
block on both maps; notice the significant difference in contour projections. Here again we
emphasize that it is very important to contour correctly over the entire map, but it is especially
critical in and around the crest of a structure where hydrocarbon accumulations are expected to
occur.
Three cross section lines are shown on each structure map. The B-B' cross section for each
map is shown in the upper right corner of Fig. 8-37a and b. The B-B' section in Fig. 8-36a shows
that if the throw of the fault were 600 ft as mapped, then the vertical separation in cross section
B-B' would have to be 725 ft; therefore, the missing section in the wells would also have to be
725 ft. The B-B' cross section in Fig. 8-37b shows a vertical separation of 600 ft, which agrees
with the missing section in each well and a throw across the fault at this location of 500 ft. This
386 Chap. 8 / Structure Maps
(a)
(b)
Figure 8-37 (a) Generic case study – structure map prepared using missing section (vertical
separation) incorrectly as if it were throw. (b) Generic case study – structure map prepared using
vertical separation to correctly contour across Faults A and B. (Published by permission of
Tenneco Oil Company.)
The Additive Property of Faults 387
cross section geometry is supported by the well log data and the completed structure map. As a
cross section exercise, complete sections A-A′ and C-C′ for both maps and evaluate the results.
If you are interested, use Eq. (7-1) again to test the validity of the structure contour map in Fig.
8-37b.
(a)
(b)
(c)
There appear to be three separate problems with this structure map. (1) Although this is a
compensating fault pattern, the map is prepared incorrectly, depicting a bifurcating pattern where
the individual vertical separations for Faults A and B are added. The 400 ft for Fault A added to
the 100 ft for Fault B equal 500 ft. Notice that the vertical separations do not even add correctly,
which is another problem related to the completed map. (2) It is obvious that this map was pre-
pared without being integrated with fault maps. When fault and structure maps are integrated, the
fault traces and gaps are properly constructed. Thus, an error of constructing fault traces that
depict the addition of fault vertical separation rather than their subtraction, such as in the case of
this compensating fault pattern, would not occur. (3) The contouring across Fault A is incorrect
at the intersection. The contours on opposite sides of Fault A east of the intersection should have
an elevation difference of 350 ft rather than the 450 ft contoured on the map, since this is a com-
pensating fault pattern.
The Additive Property of Faults 389
(a)
(b)
Figure 8-39 (a) Vertical separation is not conserved around the intersection of Faults A and B. This mapping
bust appears to be the result of incorrect contouring and the failure to integrate the fault and structure maps.
(b) The failure to conserve the vertical separation around the intersection of Faults A-1 and A-2 in this example
is significant. This mapping bust is probably due to a seismic correlation mis-tie across one or both of the inter-
secting faults. (Modified from Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
390 Chap. 8 / Structure Maps
Using tracing paper, recontour this map (Fig. 8-39a) to reflect an acceptable structural inter-
pretation around the fault intersection. Finally, examine the structure map further for mapping
busts in contouring across Fault A-2 and the intersection of Fault A-2 with A-1. See if you can
identify and correct these problems.
Figure 8-39b shows another example of a completed structure map, generated from seismic
data, that fails to honor the conservation of vertical separation. Unlike the cited error in Fig. 8-39a,
which is minor as errors go, the error in Fig. 8-39b is quite significant. A test of the additive prop-
erty of faults around the intersection of Faults A, A-1, and A-2 indicates an error on the order of
1800 ft. Such a significant error is not due to incorrect contouring, but is most likely the result of
a seismic correlation mis-tie across one or more of the intersecting faults. Look at the structural
dip direction for the westernmost fault block versus the fault block upthrown to Faults A and
A-1. What are your thoughts?
391
map. Depths are in feet. (Prepared by C. Harmon. Modified from Tearpock and Harris 1987. Published by
permission of Tenneco Oil Company.)
392 Chap. 8 / Structure Maps
the line of section. For Fault B, the upthrown elevation of the intersection of the horizon with the
fault is –9460 ft, and the downthrown elevation at the fault/horizon intersection is –9815 ft.
Finally, the vertical separation data estimated across the fault are posted and used to aid in the
mapping (e.g., 450 ft for Fault B on Line D).
When using seismic data, be aware of the difference between the apparent throw values and
the vertical separation, and label the latter on the map. The estimate for apparent throw is the dif-
ference in elevation between the upthrown and downthrown intersections of a horizon with the
fault along the line of section. Notice the changing values for apparent throw along fault A; these
changing values are a function of the orientation of the line of section crossing the fault (lines B,
D, and E). However, the values for the vertical separation along this portion of the fault do not
change. For example, based on the upthrown and downthrown values measured along lines E, B,
and D, the apparent throw of Fault A is 520 ft at line E, 425 ft at line B, and 375 ft across line D.
Considering these apparent throw values, the fault could be incorrectly interpreted as decreasing
in size or dying to the east. But the estimate of the vertical separation at each line shows that ver-
tical separation for Fault A is a constant 500 ft in the area east of the intersection with Fault B.
Contrast this with Fault B, which shows a decreasing vertical separation as the fault dies lateral-
ly to the east; however, the apparent throw, as measured from line D, is 355 ft, and it is 325 ft
from line B, suggesting a fault of almost constant apparent throw. It is most common to use the
vertical separations for mapping displacement across faults, whether the data are from well logs
or seismic sections.
The contouring techniques presented at the beginning of this chapter apply to the use of seis-
mically derived data as well as log data. Seismic data offer some additional advantages that are
lacking in well data, such as the ability to post upthrown and downthrown mapping control points
at fault/horizon intersections. These additional data give two more control points to help in con-
touring the final structure map.
In summary, seismic data allow us to extend the interpretation beyond the limits of well con-
trol. It helps in the construction of a continuous subsurface picture of the horizon being mapped
and provides additional, more complete information about the characteristics of the faults that
intersect these horizons.
or horizon. Unconformity maps are not normally made to stand alone; instead, they are integrat-
ed with a structural horizon map to delineate the intersection of the unconformity with the hori-
zon being mapped. This intersection marks the termination of the horizon at the unconformity. If
a hydrocarbon trap is present beneath the unconformity, the termination of the reservoir against
the unconformity defines a boundary of the reservoir (Fig. 8-41).
Although there are various kinds of unconformities, for petroleum exploration and exploita-
tion we are primarily interested in two types: (1) angular unconformity and (2) disconformity. An
angular unconformity is one in which the rocks above and below the unconformity are not par-
allel with one another, with the rocks below the unconformable surface typically dipping at a
steeper angle than those above. Figure 8-42 is a north-south electric log cross section showing a
good example of an angular unconformity determined by log correlation (Lock and Voorhies
1988). Good, easily traceable resistivity markers in the section below the unconformity are seen
to be progressively cut out at the unconformity surface. Notice, in the interval above the uncon-
formity, that by correlating down the logs, the section is seen to be interrupted in each well at the
same stratigraphic marker. From the Tribal Oil No. 1 to the Sinclair No. 5, as much as 150 ft of
section is missing as a result of the unconformity.
A disconformity is defined as an unconformity in which the rocks above and below the
unconformity are approximately parallel. In the subsurface, disconformities are particularly dif-
ficult to recognize. In some cases, paleontological data may be helpful, in the absence of evidence
based on well log correlation or seismic data.
Unconformities occur on a variety of scales ranging from very local in extent, such as those
associated with meandering stream channels across a flood plain, to those covering many square
miles. A major unconformity in the area around the western flank of the Sabine Uplift (located
along the Louisiana-Texas border in the United States) serves as the trap for the giant (5-billion
barrel) East Texas Oil Field (Fig. 8-43). Actually, this giant field results from the intersection of
two unconformities creating the up-dip oil trap in the Woodbine Sandstones. Unconformities are
very common around steeply dipping structures, especially salt domes. An excellent example of
a major angular unconformity associated with a salt dome is shown in Fig. 4-45.
Mapping Techniques. Figure 8-44a is a map of a deformed unconformity surface based on log
correlations and seismic data. Figure 8-44b shows the integration of the top of Stratigraphic Unit
A, which is oil-bearing, with the unconformity, thus delineating the western limit of the horizon
against the unconformity. The termination of the mapped horizon against the unconformity is
394
Chap. 8 / Structure Maps
Figure 8-42 Example of an angular unconformity recognized by electric log
correlation. (From Lock and Voorhies 1988. Published by permission of the Gulf
Coast Association of Geological Societies.)
Other Mapping Techniques 395
Figure 8-43 Structure map on top of the Woodbine Sand in the East Texas pool. As shown
in the cross section insert, the intersection of two unconformity surfaces marks the eastern
boundary of this unconformity trap. (From Geology of Petroleum, first ed. By A. I. Levorsen,
Copyright 1954 by W. H. Freeman and Company. Reprinted by permission.)
396 Chap. 8 / Structure Maps
(a) (b)
(c)
Figure 8-44 Construction of a map of the top of a reservoir trapped beneath an unconformity. (a) Contour map on an
unconformity surface based on both electric log and seismic correlation data. (b) The termination of the top of Unit A is
defined by the intersection of the unconformity with the top of Unit A. (c) The intersection of the base of Unit A with the
unconformity surface defines the western limit of this oil reservoir Unit A. (d) Map of the top of the reservoir. An uncon-
formity wedge zone is defined as the area between the unconformity traces on the top and base of the sand unit. The map
of the wedge zone is a copy of the map of the unconformity in that area. The maps of the wedge zone and the top of Unit
A are combined to form the top of reservoir map.
Other Mapping Techniques 397
determined by overlaying the structure map on the top of Unit A onto the unconformity map. The
intersection of each structure contour for the top of the unit with each unconformity contour at
the same elevation delineates the position of the unconformity on the top of Unit A (Fig. 8-44b).
This integration is basically the same as that for fault and structure maps.
Depending upon the scale of the map, thickness of the stratigraphic unit, dip of the beds, and
other considerations, a map on the top of a particular stratigraphic unit may not be sufficient to
accurately show the effect of the unconformity on that unit. At times, a structure map on the base
of the stratigraphic unit is required in addition to a map showing the unconformity wedge
between the top and base of the unit. Figure 8-44c is a structure map on the base of Unit A, and
it delineates the position of the intersection of the unconformity with the mapped horizon. The
procedure for determining the position of the unconformity on the base of the Unit A is the same
as that just shown for the top of the unit.
Notice that the unconformity traces on the top and base of Unit A are not in the same posi-
tion, indicating that there is an unconformity wedge zone between the two traces. This wedge
zone can be mapped in combination with the structure map on the top of Unit A in order to depict
the top of the reservoir, as shown in Fig. 8-44d. This map is prepared by first overlaying the struc-
ture map on the top of the unit onto the structure map on the base of the unit and transferring the
trace of the unconformity on the base to the structure map on the top of the unit. There are now
two unconformity traces on the top of structure map and they delineate the wedge zone. Between
these two traces, the reservoir varies from full thickness to zero thickness. Since the structure in
the wedge zone is actually that of the unconformity itself, next overlay the top of the structure
map onto the unconformity map and trace the structure contours in the wedge zone from the
unconformity map. For example, the –6200-ft contour on the top of Unit A, at the trace of the
unconformity, continues as the –6200-ft contour on the unconformity in the wedge zone until it
intersects the western limit of the wedge zone. Notice that the contour changes strike direction in
the wedge. This procedure is followed for all the contours on the map. The result is a structure
map on the top of the reservoir. Observe that the unconformity wedge zone for this interval com-
prises a significant area of the oil reservoir; therefore, in such situations the wedge zone must be
contoured.
Figure 8-45 Cross section shows that the values for both vertical separation and throw
are the same across a vertical fault.
One of the biggest problems with a vertical fault, besides recognizing the presence of the
fault, is estimating the vertical separation of the fault for contouring. The fault cannot be inter-
sected by a vertical well. An estimate of the vertical separation must come from intersections of
the fault with deviated wells, from seismic data, or from sufficient well control within each fault
block to contour both the upthrown and downthrown blocks separately and use the offset con-
tours to determine the vertical separation for the fault.
Figure 8-46 Completed structure contour map on the 3600-ft Horizon cut by two vertical
faults. The structure contours are drawn from one fault block to the next assuming struc-
tural continuity across the faults.
It is common for the upper portion of a particular stratigraphic unit to be composed of non-
reservoir-quality rock. This nonreservoir-quality rock is often referred to as a tight zone or tight
streak. Although the top of the unit may represent an actual stratigraphically equivalent horizon,
it is underlain everywhere by impermeable rock. Therefore, the structure maps prepared to inter-
pret the true structure commonly cannot be used to evaluate the reservoir itself.
Once a structure map is completed, the next step is to prepare a top of porosity map for accu-
rate delineation of the reservoir, and for later use in the construction of net hydrocarbon isochore
maps. Two parameters are considered in evaluating the importance of separately mapping the top
of porosity: (1) the thickness of the tight zone, and (2) the relief of the structure. A thick tight
400 Chap. 8 / Structure Maps
Figure 8-47 Electric logs from three wells. The upper stratigraphic marker conforms to true
structure and is used to construct a map representing the true structural framework of the
area. The top of the thick productive sand member does not conform to structure, but it rep-
resents a porosity top. It must be mapped separately to delineate the actual reservoir con-
figuration.
zone has a greater effect than one that is thin. Low-relief structures introduce greater error in
delineating the limits of a reservoir than steeply dipping structures, particularly if the low-relief
structure contains a bottom water reservoir.
Figure 8-48a shows a structure map and cross section for the 6000-ft Reservoir. This unit
consists of nonreservoir-quality rock in the upper 75 ft. The same reservoir is mapped on the top
of the porous rock or porosity top in Fig. 8-48b. Notice in cross section A-A' that by mapping on
the top of the unit, in which the upper 75 ft consists of nonreservoir-quality rock, the limit of the
reservoir (gas/water contact) is extended beyond the true gas/water contact as mapped on the top
of porosity (also see Fig. 8-48c). Even though no net pay is assigned to the tight zone, the pro-
ductive area of the reservoir mapped on the top of the nonproductive portion of the unit is larger.
In turn, the volume of the reservoir is also larger than that mapped on the porosity top. In this
case, the volume, based on net gas isochore maps, is larger by 32 percent. This added reservoir
area (Fig. 8-47c) created by mapping on the top of the stratigraphic unit does not contain hydro-
carbons and therefore is not productive; consequently, the volume of recoverable hydrocarbons
based on this map is overestimated.
The decision to prepare a separate map on the top of porosity, where the upper portion of a
unit is not productive, needs to be made on a reservoir-by-reservoir basis. Depending upon the
geometry of the reservoir and thickness of the tight zone, the difference in volume between a map
on the top of the unit and a map on the top of porosity may be too insignificant to warrant addi-
tional mapping. In Chapter 14 we discuss in greater detail the impact of nonproductive zones in
isochore mapping.
next horizon to be mapped. The simplest way to do this is to lay the completed map under the
base map for the next horizon to be mapped, and use it as a guide for contouring the new hori-
zon map. This technique usually ensures that consecutively mapped horizons possess structural
compatibility or contour compatibility.
Figure 8-49 illustrates the use of this technique and also shows the problems that can arise
if vertically consecutive horizon maps are not overlaid during structure contouring. In Fig. 8-49a,
the completed structure map for Horizon A was used to guide the contouring of Horizon B.
Notice the compatibility in the structural configuration of the two horizons. Figure 8-49b illus-
trates the kind of contouring errors that can arise when a completed map is not used to guide the
structure contouring of another, vertically close horizon. Notice that in the area of good well con-
trol, the contours are similar to those in Fig. 8-49a. The major problem occurs in the areas of lim-
ited control on Horizon B. Observe in Fig. 8-49b that several contours on Horizon A actually
cross contours of the same value on Horizon B (for example, the –5950-ft contour), which incor-
rectly indicates a stratigraphic thickness between Horizons A and B of zero feet, as well as a
crossing of the horizons. We know, however, from the available well data, that there are about
100 ft gross between these two horizons.
(a)
Figure 8-48 (a) Structure map on top of the 6000-ft Unit, with a gas/water contact at a depth of –6216 ft, and
cross section A-A′ illustrating (1) the top of the unit, (2) top of porosity, and (3) base of unit. (b) Structure map
on the top of porosity for the 6000-ft Unit, with the gas/water contact at a depth of –6216 ft, and cross section
A-A′. (c) Mapping on top of structure versus top of porosity results in a 32 percent increase in volume.
402 Chap. 8 / Structure Maps
(b)
(c)
(a) (b)
Figure 8-49 (a) Structure contour maps on Horizons A and B showing structural compat-
ibility. The configuration of Horizon A was used to guide the contouring of Horizon B, espe-
cially in areas of limited or no well control. (b) An alternate interpretation of the structure
contour map on Horizon B drawn independently and not taking advantage of the configu-
ration of Horizon A, which has additional subsurface well control. A geologically impossible
situation is shown; a contour of a given value (e.g., –5950) on Horizon A crosses a contour
of the same value on Horizon B.
One contouring rule states that contours of the same value cannot cross on separately
mapped horizons. If contours of the same value, mapped on different horizons, intersect or cross,
then the interval between the two horizons would have zero or negative thickness, which is a
physical impossibility. This is a serious mapping error that occurs because completed structure
maps are not used to guide the contouring of other, vertically close horizons.
A major mapping bust is shown in Fig. 6-11a and b in Chapter 6. In the situation shown, the
two horizons are not vertically close (separated by 800 ft of section in the off-structure position)
and yet the structure contour maps for the P5 and P7 Horizons cross in the up-structure position.
To avoid such errors, we recommend that completed structure maps always be used to guide the
contouring of other horizons, whether they are above or below the completed map.
The use of seismic data is very helpful in determining whether there is structural (contour)
compatibility across faults in any area being worked (Figs. 7-37 and 7-38). To be a good inter-
preter and mapper, you must have a good understanding of the tectonic setting being worked and
detailed knowledge of the individual structures being mapped. It is very important to know where
to apply a particular technique, such as mapping across faults, and where the structure being
mapped is an exception to the rule, thus requiring a different technique.
A normal growth fault, depending upon its vertical separation and the local amount of con-
temporaneous deposition, may or may not possess structural (contour) compatibility across the
fault. Usually, the larger the fault, the less chance of compatibility. For example, Fig. 7-37 shows
a growth fault (Fault A) with compatible beds across the fault. However, Fig. 8-50 illustrates a
large growth fault (Fault A) that does not have compatibility from the upthrown to downthrown
fault blocks of the main fault. Figure 8-51 is a cross section from the Gulf of Mexico that shows
a fault with both characteristics. The structural attitude is compatible across the fault in the shal-
low section, gradually becoming incompatible with depth. There are several reasons for this
change in structural attitude, including the activity of the large Miocene growth fault shown to
the far left in the cross section. The seismic section in Fig. 11-12 illustrates a faulted rollover anti-
cline. Structural compatibility is lacking across the large fault on the left, which is the fault that
Application of Contour Compatibility Across Faults
Figure 8-50 Observe the differing bed attitude upthrown and downthrown to the major growth Fault A. The
405
rollover anticline, containing all the hydrocarbon accumulations, displays a compatibility of bed attitude across
Faults B, C, and D. (Modified from Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
406
Chap. 8 / Structure Maps
Figure 8-51 The cross section from an offshore Gulf of Mexico field exhibits structural compatibility in the shal-
low section and structural incompatibility in the deeper section across the same normal fault. (Published by per-
mission of the New Orleans Geological Society.)
Mapping Techniques for Various Tectonic Habitats 407
formed the structure. On the other hand, all faults within the structure do exhibit structural com-
patibility.
Figure 8-26, discussed earlier, is an example of an area cut by a reverse fault that exhibits
structural (contour) compatibility across the fault, so the vertical separation (the repeated section
in this case) is used to map across the fault. In Fig. 6-25, however, there are significantly differ-
ent structural attitudes on opposite sides of the blind thrust fault; in this case, the technique of
contouring the vertical separation across the fault cannot be used.
Vertical separation cannot easily be used to contour across a fault where there is no compat-
ibility of structural attitude across the fault. For such structures, it is necessary to map each fault
block separately and independently using the available well or seismic control. Each contoured
fault block must be individually integrated with the fault surface map for the displacing fault to
determine the upthrown and downthrown fault traces (Tearpock and Harris 1987).
Observe in Fig. 8-50 that although the vertical separation mapping technique cannot be used
across the major growth Fault A, the faulted hanging wall anticline that contains all the reservoirs
does exhibit structural compatibility across all faults. Vertical separation must be used to map
across all the faults that cut the rollover anticline in the hanging wall block of the major growth
fault. This type of situation is quite common in both extensional and compressional areas. Major
growth normal, thrust, or strike-slip faults may downdrop, thrust, or slide one fault block with
respect to the other to such a degree that there is no compatibility of structural attitude across
these major faults. However, the associated structures formed by these faults commonly possess
structural compatibility across the faults within the structure. This is shown clearly in the cross
section in Fig. 8-50. Figure 8-22 is a structure map of a similar anticline downthrown to a very
large growth fault (which is not shown on the map). Note the structural compatibility across the
large fault (on the west), even though the vertical separation is about 1100 ft.
A good example of an area exhibiting both compatible and incompatible structural attitudes
across faults, within the same field, is seen in the Cactus-Nispero Field in the Southern Zone,
Mexico. Figure 8-52a is a structure map on the middle Cretaceous showing this salt-cored struc-
ture cut by both normal and reverse faults. Figure 8-52b and c are cross sections through the field,
both of which are shown in plan view (F-F′ and G-G′) on the structure map. A careful review of
the structure map and cross sections shows that the structure exhibits contour compatibility
across most of the faults, and therefore the vertical separation should be used to map across these
faults. Two areas provide major exceptions. In cross section F-F′, structural compatibility is lost
across the reverse fault cutting Well No. C-87. On the structure map, the large normal fault just
north of the salt body does not exhibit contour compatibility from the upthrown to downthrown
blocks. To contour these two areas accurately, each block must be contoured separately and then
integrated with the fault map for each associated fault.
All these examples stress the importance of understanding the three-dimensional geometry
of any structure being interpreted and mapped. The accuracy of the mapping depends upon the
use of the correct mapping techniques, which often depends upon the geometry of the structure
itself. Different techniques may be needed within the same field.
(a)
Figure 8-52 (a) Structure map of the Middle Cretaceous, Cactus-Nispero Field, Southern Zone, Mexico. The
structure is a salt-cored structure cut by normal and reverse faults. (b) Geologic cross section F-F′ through
the Cactus-Nispero Field. (c) Geologic cross section G-G′ through the Cactus Field. (From Santiago 1980;
AAPG©1980, reprinted by permission of the AAPG whose permission is required for further use.)
contour construction of fault traces for each of the fault patterns observed in these settings, and
present a field example where possible.
Extensional Tectonics
For extensional tectonics, we include nongrowth normal fault systems and detached listric growth
fault systems, since both commonly occur together within this setting. In many areas of the
world, hydrocarbons are trapped upthrown to normal faults. Important exceptions to this are
found in the Gulf of Mexico and the Niger Delta, where hydrocarbons are commonly trapped
downthrown as well as upthrown to normal faults. Also, reservoir-bearing rollover anticlines
commonly occur in blocks downthrown to listric growth normal faults.
Mapping Techniques for Various Tectonic Habitats 409
(b)
(c)
Faults are the primary trapping mechanism for hydrocarbons in extensional habitats,
although four-way closures, unconformities, and stratigraphic traps are also important. Our pri-
mary discussion centers around five different types of fault patterns: (1) single nongrowth (struc-
tural) faults, (2) compensating, (3) bifurcating, (4) intersecting, and (5) growth fault systems.
The construction techniques for all fault patterns that exhibit structural compatibility across
a fault are the same as those used for the working model (Fig. 8-21) to arrive at an integrated
structure map. For the examples of each fault pattern, we show at least two mapped horizons that
bound a stratigraphic interval of constant thickness. For the purpose of detailing the correct map-
ping techniques, the data for the examples are somewhat idealized. However, the techniques are
applicable in both simple and complex petroleum geological mapping. Where possible, actual
examples for each of the fault patterns are presented.
Compensating Fault Pattern. Figure 8-53a shows the fault surface contour map for Faults A
and B, which together represent a typical compensating fault pattern. If necessary, refer again to
Chapter 7 and review the major aspects of this fault pattern. Several key points must be re-empha-
sized here. First, the line of termination must be shown on the fault surface map; its importance
will become clear as we review the structure maps on the 6000-ft and 7000-ft Horizons. Second,
vertical separation is conserved across fault intersections. Observe that throughout the area where
two faults are present (i.e., the area south of the line of termination of Fault B at Fault A in our
example), Fault A has 100 ft of vertical separation and Fault B has a vertical separation of 200 ft.
Only Fault A is present north of the termination line, and there its vertical separation is 300 ft.
Figure 8-53b is a completed structure contour map on the 6000-ft Horizon. The structure is
a faulted anticline. Figure 8-53c shows the fault surface map for Faults A and B superimposed
onto the completed structure map for the 6000-ft Horizon. Using this figure, we review the con-
struction of the upthrown and downthrown traces for each fault, the construction of the intersec-
tion of the two faults, and the proper horizon contour construction across each fault using the
vertical separation (missing section from each wellbore).
Begin with the northern upthrown fault block and the –6200-ft contour, and follow its com-
plete construction over the entire structure. The intersection of this structure contour with the
fault contour of the same value defines the upthrown trace of Fault A at point A. The contour is
projected through the fault, as if the fault were not there, and adjusted in the downthrown block
by the amount of the vertical separation (300 ft) to become the –6500-ft contour. The intersec-
tion of this –6500-ft structure contour with the fault contour of the same value defines the down-
thrown trace of Fault A at point B. The –6500-ft contour downthrown to Fault A crosses the axis
of the structure and intersects with the –6500-ft contour for Fault B at point C, defining the down-
thrown trace of Fault B at this location. The contour is now projected through Fault B, as if the
fault were not there, and then adjusted by 200 ft (representing the vertical separation for Fault B)
in the upthrown block of Fault B to become the –6300-ft contour. This –6300-ft structure contour
intersects the –6300-ft fault surface contour at point D, defining the upthrown trace of Fault B at
this location. This contour once again crosses the axis of the structure, this time on the eastern
side of the anticline, and intersects with the –6300-ft contour of Fault A at point E, defining the
downthrown trace of Fault A. Project the contour through the fault gap of Fault A and adjust its
value by 100 ft for the vertical separation at point F. Finally, this contour continues in the
upthrown block of Fault A until it intersects with point A, which was our starting point. This tech-
nique is applied for each structure contour on the map to arrive at the integrated structure map
shown.
Mapping Techniques for Various Tectonic Habitats 411
The heavy dashed line (Fig. 8-53c) striking southwest-northeast is the line of termination of
Fault B at Fault A. The intersection of the fault traces on the completed structure map for both
faults falls on this line of termination. For this map, there are two trace intersections: (1) the inter-
section of the downthrown traces of Faults A and B shown at point G, and (2) the intersection of
the upthrown trace of Fault B with the downthrown trace of Fault A shown at point H. Since all
trace intersections must fall on the line of termination, it is very important to place this line on
the fault surface map. By following the line, we can predict the location of the fault intersections
for other horizons either shallower or deeper in the subsurface, because all intersections must fall
on this line. Observe that north of the line of termination, only Fault A is present, with a vertical
separation of 300 ft, whereas south of the line, both Faults A and B are present, with vertical sep-
arations of 100 ft and 200 ft respectively. Take a minute and test the additive property of faults
around the fault intersection on the completed structure map.
This construction technique of integrating the fault and structure maps removes all the guess-
work with regard to: (1) fault trace construction; (2) the location of all faults on the structure map;
(3) the location of all fault intersections; and (4) the width of the fault gaps. The integration of
the two maps constructs all these features automatically without the need for guesses or esti-
mates. Take one or two additional structure contour lines and go through the same procedure just
described to make sure you fully understand the technique for integrating a fault map with a
structure map.
Figure 8-53d is a structure map on the deeper 7000-ft Horizon, and it shows that the faults
have migrated with depth. The technique of construction is exactly the same as presented for the
6000-ft Horizon. In Fig. 8-53e, we again superimpose the fault map onto the completed structure
map. Choose several structure contours and follow each across the mapped area to review the
construction techniques as discussed for the preparation of the completed structure map for the
6000-ft Horizon.
Finally, the completed fault traces for both the 6000-ft and 7000-ft Horizons, and the gener-
al trend of the structure contours is superimposed onto the fault map for Faults A and B in Fig.
8-53f. Looking at each set of fault traces (those for the 6000-ft and 7000-ft Horizons), can you
detect any difference between them? Use 10-point proportional dividers or an engineer’s scale to
measure the width of the fault gap for Fault A on both horizons. Notice that the gap is wider for
Fault A on both sides of the line of termination on the structure map for the 7000-ft Horizon,
when compared with the fault gap on the structure map for the 6000-ft Horizon. Remember the
general rule of fault gap width presented earlier. A fault dipping up-dip (in the opposite direction
of bed dip) will have a narrower fault gap than a fault of the same or similar vertical separation
dipping down-dip (in the same direction as the bed dip). Notice that Fault A on the 6000-ft
Horizon is on the northern side of the structural axis and is dipping generally in the opposite
direction to the formation. On the deeper structure map for the 7000-ft Horizon, Fault A has
crossed the axis of the structure and is now dipping more or less in the same general direction as
the formation. Therefore, the fault gap is wider as mapped on the structure map for the 7000-ft
Horizon.
With the technique of integrating the fault and structure maps, this change in fault gap width
is automatically built into the integration. It is unlikely that such a change in gap width could
have been intuitively incorporated into the completed structure contour maps had the fault and
structure map integration technique not been used. For this example, the change in gap width is
significant, but depending upon the configuration of the structure and the geometry of the fault,
the change in width can be minor. Let the integration technique be the tool used to establish fault
gap widths.
412 Chap. 8 / Structure Maps
(a)
(b)
Figure 8-53 (a) Fault surface map for a compensating fault system including Faults A and B. (b) Integrated structure map
on the 6000-ft Horizon. (c) Fault and structure maps (6000-ft Horizon) superimposed to illustrate the details of the inte-
gration technique. (d) Integrated structure map on the 7000-ft Horizon. Upthrown and downthrown restored tops in Wells
No. 9 and 14 were used to aid in the structural interpretation. (e) Fault and structure maps (7000-ft Horizon) superimposed
to show the accurate construction of the faulted 7000-ft Horizon map using the integration technique. (f) Overlay of fault
traces for both the 6000-ft and 7000-ft Horizons. The figure shows that the intersections of all fault traces fall on the line
of termination. The location of the fault intersections on any structural horizon can be predicted from the strike of the line
of termination.
Mapping Techniques for Various Tectonic Habitats 413
(c)
(d)
(e)
(f)
The completed structure map in Fig. 8-22 illustrates two compensating faults to the large
down-to-the-east normal fault. Fault surface maps were prepared for each fault, and they were
integrated with the structure map to arrive at the completed map.
Look at the completed structure map in Fig. 8-54. What type of fault pattern is represented
by the intersection of Faults A and D? Is it a compensating fault pattern, or is it a mapping error?
Can you find any other errors on this map?
Bifurcating Fault Pattern. The fault map in Fig. 8-55a shows a fault contour map for Faults A
and A-1, which together form a bifurcating fault pattern. Refer again to Chapter 7, if necessary,
to refresh your memory on the characteristics of a bifurcating fault pattern. Figure 8-55b is the
completed structure map on the 6000-ft Horizon. Figure 8-55c shows the fault surface map for
Faults A and A-1 superimposed onto the structure map for the 6000-ft Horizon. Choose several
structure contours upthrown to Fault A and go through the integration technique to check the
position of the fault traces and the fault intersections, the width of the fault gaps, and the contour
construction for the 6000-ft Horizon across the two faults.
Fault traces are relatively easy to delineate when there are a number of structure contours
intersecting fault contours, as in the case shown in Fig. 8-55; however, in areas where fault con-
tours are semiparallel to the structure contours, it is more difficult to delineate the fault traces. In
these situations, you must interpolate depth values to precisely delineate the position of the fault
traces.
Figure 8-55d is the completed structure map on the 7000-ft Horizon. Figure 8-55e shows the
fault surface map superimposed onto the structure map. Observe that the trace intersections for
416 Chap. 8 / Structure Maps
Faults A and A-1 fall directly on the line of bifurcation. Figure 8-55f shows the fault traces, as
seen on the completed structure maps for both the 6000-ft and the 7000-ft Horizons with the fault
surface map superimposed. Notice how the fault intersection migrates from north to south along
the path of the line of bifurcation. Also observe that the width of the fault gap for Fault A is wider
on the structure map for the 7000-ft Horizon than it is on the 6000-ft Horizon structure map. This
occurs here for the same reason discussed for the compensating fault example.
Figure 8-56 is a portion of the structure map on the Nodosaria Marker, and Fig. 8-57 is a
portion of the U Sand structure map from the offshore Gulf of Mexico. Both figures show good
examples of the mapping of a bifurcating fault system. The fault system in Fig. 8-56 actually has
two bifurcations. Take a minute to check the construction of these completed maps in the area
around the bifurcations by applying the additive property of faults.
Now we revisit Fig. 8-54. Both of the faults that merge dip in the same direction, so this is
a bifurcating system. The gap for the fault that extends to the east of the point of merger would
be expected to be larger than the gap on either of the two faults, thus reflecting the additive prop-
erty of faults. However, the gap is much narrower than that of either fault. Actually, the gap is of
the size expected if this were a compensating fault system, so the interpretation is incorrect. The
only way that the gap could be correctly constructed to be so narrow is if the fault were to abrupt-
ly become almost vertical right at the point of merger of the two faults, and that is certainly unrea-
sonable. In further review of the map, we can use the implied strike technique on the northern
(a)
Figure 8-55 (a) Fault surface map for a bifurcating fault system including Faults A and A-1. (b) Integrated
structure map on the 6000-ft Horizon. (c) Fault and structure maps (6000-ft Horizon) superimposed to show
the accuracy of the integration technique: to position the fault traces and intersections and to determine the
width of the fault gap. (d) Integrated structure map for the 7000-ft Horizon. (e) The fault map superimposed
onto the 7000-ft structure map. (f) Overlay of fault traces on both the 6000-ft and 7000-ft Horizons. As with
the compensating fault pattern, all intersections of fault traces fall on the line of bifurcation.
Mapping Techniques for Various Tectonic Habitats 417
(b)
(c)
(d)
(e)
(f)
fault, whose trace bends sharply. Using the points where the –14,900-ft contours meet the fault
trace, we see that the implied fault strike abruptly changes. Is such a sharp bend reasonable for a
normal fault? Finally, structural compatibility is lacking across the northern fault. Notice that the
vertical separation decreases rapidly from 600 ft to 300 ft in an easterly direction. The poor con-
tour compatibility is apparently due to the interpreter honoring that displacement. Again, do you
think that this abrupt change in vertical separation is reasonable along this single fault, or could
something else be happening? Notice that it occurs near the sharp bend in the implied fault sur-
face. Could Fault D actually be two faults?
Intersecting Fault Pattern. Figure 8-58a illustrates a typical intersecting fault pattern com-
posed of Faults A and B. Cross section A-A' is shown in Fig. 8-58b. In the section in Chapter 7
dealing with the intersecting fault pattern, the fault surface maps were contoured with the
assumption that neither fault surface was offset (Fig. 8-58b). In other words, the two faults were
considered contemporaneous rather than of two different ages. The fault traces for the structure
maps shown in Fig. 8-58 were constructed with this assumption. Notice, however, in Fig. 8-58c
through g, that all the fault traces are offset. Figure 8-58d shows, for example, that although the
fault surfaces themselves are not offset, the fault traces on the completed structure maps show
significant offsets for all the fault traces at the fault intersections. The positions of the offset
traces are accurate for the constructed fault surface maps. If the faults are of two different ages,
420 Chap. 8 / Structure Maps
Figure 8-56 Portion of the structure map on the Nodosaria Marker illustrating a bifur-
cating fault system. (Published by permission of the Lafayette Geological Society.)
Figure 8-57 Example of a bifurcating fault system from the offshore Gulf of Mexico.
(Published by permission of the New Orleans Geological Society.)
Mapping Techniques for Various Tectonic Habitats 421
however, then the completed maps are subject to error around the intersection. If the faulting is
contemporaneous, then the traces are accurate as shown on the completed structure maps.
Compare the completed structure maps in Fig. 8-58d and f. There is a pronounced change in
the configuration of the fault traces shown on these two maps. The change is the result of the
intersecting geometries of the two horizons with Faults A and B at different depths. Without the
integration of the fault and structure maps, the accurate delineation of these fault traces, particu-
larly for the 7000-ft Horizon, would be difficult, if not impossible.
Finally, observe that all the fault trace intersections fall on the line of intersection. In this
case of two intersecting faults, there are four fault trace intersections (Fig. 8-58e and g). This con-
struction again emphasizes the necessity of placing the line of intersection on the completed fault
maps.
Where faults of two different ages intersect, the construction of accurate fault surface maps
showing the offset of an earlier fault by a later one may require seismic data to determine the rel-
ative ages of the faults. Also, a graphical solution to the construction of the fault surfaces and their
effects on stratigraphic horizons can be undertaken. A graphical technique is presented by
Dickinson (1954). The method is not detailed here; however, if it is absolutely necessary to pre-
pare precise fault and structure maps reflecting the accurate geometry of intersecting faults of dif-
ferent ages, then the Dickinson method should be considered.
Figure 8-59, from Dickinson’s paper, shows a completed structure map for a horizon above
a deep-seated salt dome in Texas. Such a complex intersecting fault pattern is not uncommon in
areas of salt tectonics. In this example, faults 3 and 5 are intersecting. A test was planned for the
upthrown block to the east. In order to test the block at an optimum structural position, it was
necessary to accurately locate the position of the intersection of faults 3 and 5, so the graphical
solution outlined by Dickinson was used to prepare the fault and structure maps.
In addition to the intersecting normal faults, some areas around the world, such as western
Texas, California, southern Oklahoma, the North Sea, and the southern zone of Mexico, contain
intersecting normal and reverse faults. The Dickinson method is also very helpful in developing
a structural picture of these intersecting faults. For many situations, however, particularly those
involving small faults, the shortcut method of assuming contemporaneous faulting, shown earli-
er in this section, is often adequate for the construction of the fault surface maps and completed
structure maps.
We cannot leave this subject of intersecting faults without covering an important misunder-
standing with regard to intersections of fault traces on completed structure maps. There seems to
be a popular belief that where two faults of different ages intersect, the fault trace of the inter-
sected or older fault is not offset on a completed structure map, as shown in Fig. 8-60. We strong-
ly emphasize that this idea is incorrect for most cases. The traces of each of the intersecting
faults on any mapped horizon always offset, except in two special cases. These exceptions,
which are covered in detail by Dickinson (1954), occur where (1) the beds are horizontal, the
faults intersect at right angles, and the movement on both faults is entirely dip-slip (Fig. 8-60);
or (2) the intersection of the (older) intersected fault and the mapped horizon is parallel to the
direction of slip on the younger fault. Obviously, these two exceptional situations are rare. In both
cases, the trace of the (older) intersected fault will not be offset, but the trace of the (younger)
intersecting fault will be. So, if neither of the faults traces is offset, then the structure map is
incorrect (no exceptions). These guidelines apply to intersecting normal faults and intersecting
reverse faults. Dickinson (1954) also discusses intersections of normal faults with reverse faults.
A structure map with intersecting faults showing the intersected (older) fault traces with no
offset must meet one of the two special cases, or it is constructed incorrectly. Usually, this type
422 Chap. 8 / Structure Maps
of error occurs because (1) the interpreter does not understand the relationships of intersecting
fault construction, or (2) no fault surface maps were prepared and used to integrate with the struc-
ture map to arrive at an accurate solution for the intersecting faults. Figure 8-61 is a portion of a
completed structure map with an intersecting fault pattern. A review of the map shows that only
one fault trace is offset for each pair of crossing faults, so the map is immediately suspect.
Consider the two exceptions to the rule that both traces should be offset. The map shows that the
beds are dipping, so exception (1), based on horizontal beds, does not apply. However, exception
(2) might be valid if the movement on Fault B was in exactly the same direction as the trace of
Fault A. However, we cannot determine that. The only way to determine the accuracy of the fault
traces is to have a fault surface map and integrate it with the horizon map. However, for the map
in Fig. 8-61, the fault traces were prepared showing no offset for the assumed older Fault A.
If a pair of intersecting fault traces is shown on a completed structure map with neither trace
offset and the traces resembling a cross (Fig. 8-62), the construction is incorrect. Such errors as
the ones shown in Figs. 8-61 and 8-62 can place suspicion on the rest of the structural inter-
pretation.
(a)
Figure 8-58 (a) Fault surface map for an intersecting fault system including Faults A and B. Fault B does not ter-
minate at the line of intersection. (b) Cross section A-A′ illustrates the shortcut method of constructing the fault
map as if the fault surfaces are not offset. (c) Integrated structure map on the 6000-ft Horizon. (d) Completed
structure map on the 6600-ft Horizon. Observe that all fault traces are offset at the intersection. (e) Fault surface
map and the structure map on the 6600-ft Horizon superimposed to show the accuracy of fault trace construc-
tion. (f) Completed structure map on the 7000-ft Horizon. (g) Overlay of the fault surface map and the 7000-ft
structure map. Notice that all fault trace intersections fall directly on the line of intersection.
Mapping Techniques for Various Tectonic Habitats 423
(b)
(c)
Figure 8-58 (continued)
424 Chap. 8 / Structure Maps
(d)
(e)
(f)
(g)
Figure 8-59 Structure contour map over a deep-seated salt dome in Texas illustrating sev-
eral intersecting faults. (From Dickinson 1954; AAPG©1954, reprinted by permission of the
APPG whose permission is required for further use.)
If the area you are working is complexly faulted with intersecting faults and extreme accu-
racy is required, it is advisable to become familiar with the Dickinson method for resolving the
geometry of fault intersections.
Figure 8-60 Block model and map view of intersecting normal faults in horizontal strata.
Fault dip is 60 deg and the movement is all dip-slip. This result is only applicable in this spe-
cial case of right-angle faults and flat beds. (Modified from Dickinson 1954; AAPG©1954,
reprinted by permission of the APPG whose permission is required for further use.)
Growth Faults. A growth fault is a special type of normal fault. A growth fault is a normal fault
that is contemporaneous with deposition, and it is often referred to as a syndepositional fault.
There are several primary characteristics of growth faults.
1. They are commonly arcuate in lateral extent and concave toward the basin.
2. Fault dip typically decreases with depth, commonly becoming a bedding plane fault at
depth (Chapter 11). We refer to this shape as listric, based on the Greek word listron, or
shovel, due to its curved shape in cross section. However, fault dip can locally increase
with depth (see the section Compaction Effects on Growth Normal Faults, in Chapter 11).
3. The vertical separation for the fault normally increases with depth. Displacements of
several thousand feet are not uncommon. Growth faults can also include displacements that
decrease with depth as well (see Chapter 11).
428 Chap. 8 / Structure Maps
Figure 8-61 Portion of a structure map showing incorrectly interpreted and constructed
intersecting faults.
4. The time-stratigraphic intervals in the hanging wall blocks are thicker than the equivalent
intervals in the footwall blocks (see the section Estimating the Vertical Separation for a
Growth Fault, in Chapter 7).
5. A rollover anticline commonly develops in the hanging wall fault block of a growth fault
as a result of collapse of the hanging wall block.
6. Secondary faulting within the hanging wall block is normally associated with growth
faults. These faults can be synthetic (dipping parallel to the master fault) or antithetic
(dipping toward the master fault).
A majority of growth faults dip toward the current basin or the paleo-basin and strike paral-
lel or semiparallel to the coastline. However, growth faults can dip in any direction, including
toward the margin of the basin. Along strike, most growth faults are generally concave basinward.
Faults are not perfect planar surfaces of slip; they generally have undulations of some type.
As two fault blocks slip past one another, there must be deformation in at least one fault block,
because rocks are not strong enough to support large voids. This is the main reason why many
major rollover folds exist within the hanging wall fault blocks, formed by bending of the fault
blocks as they slip over nonplanar (listric) fault surfaces. This mechanism of folding is called col-
lapse folding (Hamblin 1965). The rollover anticline, associated with listric growth faults, is one
Mapping Techniques for Various Tectonic Habitats 429
Figure 8-62 Completed structure map showing several intersecting faults with each of the
trace intersections forming a “cross.” (The final construction is incorrect. No fault traces are
offset.)
of the most widely recognized petroleum-producing folds, as seen in such areas as the Gulf of
Mexico, Brunei, and the Niger Delta.
For contouring purposes, the hanging wall and footwall may or may not have structural com-
patibility across a growth fault. The primary factors are the vertical separation of the fault and the
amount of deposition in the downthrown block. Usually, the larger the fault, the less the compat-
ibility. If structural compatibility does exist, the vertical separation can be contoured across the
fault. If compatibility is absent, then each fault block must be contoured separately. At times, only
the upthrown or downthrown block of a growth fault is mapped, usually because (1) good con-
trol or correlation is lacking across the fault, or (2) only one block is productive and that is the
only block mapped.
Figure 8-63 is a structure map on the G Sand in East Cameron Block 270 Field, Gulf of
Mexico. This field is productive from 19 separate horizons trapped both upthrown and down-
thrown to the west-dipping growth Fault A. Rollover into Fault A has formed a north-south trend-
430 Chap. 8 / Structure Maps
Figure 8-63 Structure map on the G Sand Unit, East Cameron Block 270 Field, Gulf of
Mexico. A portion of the fault surface map for Fault A is superimposed on the structure map.
The primary hydrocarbon trap is a faulted rollover anticline formed in response to Growth
Fault A. (Published by permission of the New Orleans Geological Society.)
ing anticline, which is the primary hydrocarbon trap. This field has produced more than 38 mil-
lion barrels of oil and condensate and nearly 1 trillion cu ft of gas.
Superimposed on the southern half of the structure map is a portion of the fault surface map
for Fault A from –8000 ft to –9000 ft. The structure contours drawn upthrown and downthrown
to the fault, in this area, indicate reasonable compatibility in structure across this growth fault,
which has a vertical separation of about 800 ft at this level. The existence of this structural com-
patibility does not, however, ensure such compatibility along fault strike nor with depth (recall
the cross section in Fig. 8-51). Notice that the rollover anticline is complexly faulted with sec-
ondary synthetic and antithetic faults. Although the fault maps for these faults are not published,
it is easy to recognize that there is good structural compatibility across all the secondary faults;
therefore, the vertical separation must have been used to contour across these faults.
One final note on growth faults. It is commonly observed that the most productive interval
in a growth fault complex is the stratigraphic section deposited during the most active fault
movement. Therefore, when prospecting around growth faults, it is important to understand the
history of the fault movement. The technique of plotting a growth fault’s history, the Multiple
Bischke Plot Analysis (MBPA) discussed in Chapter 13, is most helpful in evaluating the fault
history and its potential for hydrocarbon accumulations.
Mapping Techniques for Various Tectonic Habitats 431
Hydrocarbon Traps. Salt structures have an extremely varied and complex geometry resulting
in numerous types of hydrocarbon traps. Any single salt structure may have more than one type
of trap, depending upon the history of the salt structure and surrounding sediments. Figure 8-68
432 Chap. 8 / Structure Maps
Figure 8-64 Cross section showing the needle-like Rabbit Island Salt Spine, which is less
than 2000 ft across to a depth of about 7000 ft. (Published by permission of the New
Orleans Geological Society.)
(Halbouty 1979) shows an idealized section through a salt dome illustrating the more common
types of hydrocarbon traps, including (1) a simple domal anticline, (2) graben fault traps over the
dome, (3) porous caprock (limestone or dolomite), (4) up-dip pinchouts of permeable units, (5)
traps beneath an overhang, (6) traps against the salt itself, (7) unconformities, (8) traps along
faults downthrown away from the dome, and (9) traps along faults downthrown toward the dome.
Fault traps like 8 and 9 can also be considered first-fault out-of-the-basin type traps. In addition
to the traps shown in Fig. 8-68, radial and peripheral faults also serve as excellent hydrocarbon
traps.
Considering their complexity, salt structures require very precise interpretation and detailed
mapping in order to exploit all the hydrocarbon potential. When working with salt-related struc-
tures, maps are needed for the salt itself and for salt/sediment, salt/fault, and fault/sediment inter-
faces, in addition to associated unconformities and top of diapiric shale, if present.
Contouring the Salt Surface. No special techniques are required to contour the salt surface.
The key to making a good salt map is having sufficient data. Typically, only a limited number of
Mapping Techniques for Various Tectonic Habitats 433
Figure 8-65 Structure map on the top of the Marchard-Timbalier-Calliou Island Salt
Massif. (From Atwater and Forman 1959; AAPG©1959, reprinted by permission of the
APPG whose permission is required for further use.)
wells penetrate salt, and commonly the quality of seismic data near salt is poor. Figure 8-69 is
the salt surface contour map for Main Pass Block 299 Field, which is of special interest because
the salt dome has an overhang on the south and southwest flanks. Notice how the contours under
the overhang are dashed, making the map easier to use when preparing structure contour maps,
as well as making the salt map easier to read. Some of the deviated wells were drilled to traps
beneath the overhang.
Salt–Fault Intersection. Figure 8-70a illustrates the method for contouring the intersection of
a salt mass and a fault. The technique is basically the same as that presented earlier for contour-
ing the upthrown or downthrown traces of a fault. The salt–fault intersection occurs where the
structure contours on the salt intersect the fault contours of the same elevation. Like all inter-
secting surfaces presented in map view, the salt–fault intersection should be delineated as shown
in the figure. This line of intersection can serve as a directional guide for planning a deviated well
that tests potential reservoir units in the structural trap between the surfaces.
Figure 8-67 Map of salt pillows and piercement salt bodies in the Southern Permian Basin,
North Sea Area. (From Remmelts 1995; AAPG©1995, published by permission of the
AAPG whose permission is required for further use.)
Completed Structural Picture. The mapping of all intersections and the integration of the
fault, salt, and structure maps results in a completed structural interpretation, such as that shown
in Fig. 8-70c for the 8000-ft Horizon. This example is relatively simple, but it does illustrate the
use of the mapping techniques. Salt features are normally associated with highly faulted, highly
complex structures such as the one shown in Fig. 8-71, which is a structure map on the Grand
Isle Ash at Grand Isle Block 16 Field, northern Gulf of Mexico. This structure map is unique in
that it contains contours both on the salt and on the structural horizon. It is the intersection of
these salt and structure contours that delineates the salt/sediment boundary.
As mentioned in Chapter 6, steeply dipping structures, such as those related to salt tecton-
ics, require the layout of a number of cross sections to help develop an accurate geologic inter-
pretation and aid in the structure map construction. Typical cross sections for steeply dipping
structures such as a salt dome are designed to incorporate both straight and deviated wells.
Initially, problem-solving type cross sections are laid out to help resolve correlation problems and
436 Chap. 8 / Structure Maps
Figure 8-68 An idealized diapiric salt structure showing common types of hydrocarbon
traps. (From Salt Domes, by Michel T. Halbouty. Copyright 1979 by Gulf Publishing
Company, Houston, Texas. Used with permission. All rights reserved.)
aid in developing the structural interpretation. Later, during the advanced stages of mapping,
these sections can be converted to finished illustration sections for display and presentation. For
further information on the use of cross sections as an aid to structure interpretation and mapping
of salt structures, refer to Chapter 6.
A salt diapir commonly forms the core of a large structural high. In response to the draping
of strata around the diapir as the structure develops, the strike of the strata is typically parallel or
subparallel to the face of the salt. So, the structural contours on a given horizon tend to be paral-
lel or subparallel to contours on the salt surface. This is illustrated in Fig. 8-71 for Grand Isle
Block 16 Field.
Around certain salt structures, subsurface data show structure contours intersecting salt at a
sharp angle. Where this happens, it may be an indication of several possible situations, including
(1) the possibility of a salt overhang, such as the one shown in Fig. 8-72, or (2) an unrecognized
peripheral fault sliding down or located near the face of the salt, causing the strike direction of
the contours to appear to turn into the salt. Actually, the contours are striking into the down-
thrown side of the peripheral fault. Figure 8-73 is an example of such a situation. Notice on the
Mapping Techniques for Various Tectonic Habitats 437
southwestern portion of the salt dome that the contours appear to strike directly into the salt. Due
to the complex nature of the faulting pattern, Faults C, D, and E, which appear as radial faults
shallow in the section, become peripheral faults with depth, paralleling the southwest flank of this
salt structure. As a result, the structure contours tend to strike at various angles into the fault. In
contrast, observe the general conformity of the structure contours with the salt on the western and
eastern flanks of the structure not affected by peripheral faulting.
Figure 8-69 Surface of salt, Main Pass Block 299 Field, Offshore, Gulf of Mexico. Observe
the dashed contours to clearly illustrate the salt overhang. Some deviated wells extend
beneath the overhang. (Published by permission of the New Orleans Geological Society.)
438 Chap. 8 / Structure Maps
(a)
(b)
(c)
Figure 8-70 (a) Integration of the contour maps on the salt surface and on Fault A delineates the intersection of the fault
with salt. (b) Integration of the contour maps on the salt surface and on the structure contours for the 8000-ft Horizon. The
salt–sediment boundary is located where the salt contours intersect the fault contours of the same elevation.
(c) Completely integrated structure map on the 8000-ft Horizon, Reservoir A, in the southeast portion of this piercement
salt structure. This oil reservoir is bounded to the north and west by salt, to the east by Fault A, and to the south by an
oil/water contact at a depth of –8605 ft.
Mapping Techniques for Various Tectonic Habitats 439
Figure 8-71 Structure contour map on the Grand Isle Ash and on the salt at Grand Isle Block 16 Field, north-
ern Gulf of Mexico. The outline of the salt at this map level is defined by the intersection of the salt and struc-
ture contours at the same elevation. (Published by permission of the Lafayette Geological Society.)
440 Chap. 8 / Structure Maps
Figure 8-72 Generalized cross section of Bethel Dome, Anderson County, Texas, show-
ing a hydrocarbon accumulation below the salt overhang. The significant amount of over-
hang might indicate that the dome is detached from the salt source bed at depth. (From
Salt Domes, by Michel T. Halbouty. Copyright 1979 by Gulf Publishing Company, Houston,
Texas. Used with permission. All rights reserved.)
Figure 8-73 Structure contours striking into the downthrown side of Faults E and D give
the appearance that the contours are striking into the salt.
Mapping Techniques for Various Tectonic Habitats 441
(a)
Figure 8-74 (a) Conceptual model of simple left strike-slip fault: Top – block diagram; middle – plan
view; bottom – cross-sectional views. (From Stone 1969. Reprinted by permission of the Rocky
Mountain Association of Geologists.) (b) Block diagram illustrating different deformational patterns on
opposite sides of a finite strike-slip fault. (From Bell 1956; AAPG©1956, reprinted by permission of the
AAPG whose permission is required for further use.) (c) The change in direction of asymmetry or fold
frequency across the fault is due to a different response to the compressional forces. (From Brown 1982;
AAPG©1982, reprinted by permission of the AAPG whose permission is required for further use.)
442 Chap. 8 / Structure Maps
(b)
(c)
The primary hydrocarbon traps associated with strike-slip faults are anticlines that straddle
the strike-slip system. These anticlines, which may be faulted by either normal or reverse faults,
are good traps because they form early and commonly develop large closures sufficient to trap
economic quantities of hydrocarbons. In this chapter, our interest is in the mapping techniques
that are applicable to these strike-slip fault systems. Figure 8-75 shows an example of a strike-
slip fault system with offset faulted anticlines on each side of the fault. In many cases, the faults
cutting the anticlines are small and simply offset the structures with little, if any, change in struc-
tural attitude across the fault.
Two different types of mapping techniques are required to map these strike-slip faulted
structures. Normally, the structures on either side of the main strike-slip fault must be mapped
independently (Fig. 8-75) because structures tend to terminate against the strike-slip fault (Bischke,
Suppe, and del Pilar 1990). As for the individual anticlines that form on either side of the main
strike-slip fault, the vertical separation technique is commonly applicable for mapping across the
faults within these anticlines. In other words, the faulted anticlines commonly exhibit contour
Mapping Techniques for Various Tectonic Habitats 443
(structural) compatibility across the small normal or reverse faults that cut the anticlines, unless
some type of rotation occurs along the fault.
Figure 8-76 is a structure map from the Rosecrans Oil Field, California, USA. Although
there is currently some debate as to whether the Inglewood fault system is a true strike-slip sys-
tem or a transpressional system, it can be used to illustrate the mapping techniques applicable for
a strike-slip fault system. Notice that the anticlines adjacent to the Inglewood Fault are cut by
small reverse faults. The cross-section insert and the structure map show that there appears to be
structural compatibility across most of these small faults. With the use of well control and seis-
mic data, the vertical separation for each fault can be determined and used to contour across the
faults.
Notice the two faults labeled C and D on the structure map. Based on this structure map and
available cross sections, some fault block rotation, in addition to dip-slip motion, occurred on
these faults and resulted in a loss of structural compatibility across the faults. Therefore, it may
be difficult, if not impossible, to contour across these faults using the vertical separation. The
structures on either side of Faults C (Compton Thrust Fault) and D may require independent con-
touring. However, north of Fault D the structure again has good contour compatibility across the
northern two faults shown on the map. Another possible interpretation for Faults C and D is that
these faults do not exist, and instead the apparent faults are actually axial surfaces that bisect
kinks in the strata (see Chapter 10).
Figure 8-77 is a structure map on the top of the Ranger Zone in the Wilmington Field in the
Los Angeles Basin, California. The anticlinal structure that forms the Wilmington Field is 11 mi
long and 3 mi wide. This giant field has produced over 1.2 billion barrels of oil. During the
Middle Miocene, compressive stresses formed a north-south couple, folding the strata and estab-
lishing the present northwest-southeast Wilmington structure. During its complex structural histo-
ry, the anticline developed a series of normal faults as a result of tensional forces acting along the
structural axis. These normal faults are small, ranging from 100 ft to 400 ft of vertical separation
and dipping at angles between 45 deg and 65 deg. Many of the faults are sealing and therefore
444 Chap. 8 / Structure Maps
Figure 8-76 Rosecrans field structure (After California Div. Oil & Gas, 1961) shows a dis-
tinct pattern of reverse-faulted anticlinal folds oriented obliquely to the Inglewood Fault.
(Modified from Harding 1973; AAPG©1973, reprinted by permission of the APPG whose
permission is required for further use.)
Mapping Techniques for Various Tectonic Habitats 445
WILMINGTON FIELD
TORRANCE FIELD
City of Long Beach
-400
0' N
-3000' -350
0'
-3
-300
00
0'
0'
-3
50 -250
0' 0'
-2
50
0'
-3
-400
50
0'
0'
-250
0'
-350
0'
-300
0'
-400
0'
-300
0'
-350
0'
Figure 8-77 Structure contour map on the top of the Ranger Zone, Wilmington Oil Field, southern California.
The normal faults in the field exhibit several patterns: (1) single, (2) compensating, (3) bifurcating, and (4) inter-
secting. Length of the mapped area is about 10 miles. (After Mayuga 1970; AAPG©1970, reprinted by permis-
sion of the AAPG whose permission is required for further use.)
divide the field into at least seven major production blocks (Mayuga 1970).
A review of the structure map in Fig. 8-77a and the cross section in Fig. 8-77b suggests very
good structural (contour) compatibility across all the faults in the field. Therefore, all structure
and fault map integration in this field would use the vertical separation technique for contouring
across the faults. The Wilmington Field is another example of a strike-slip fault-associated struc-
ture that exhibits good internal structural compatibility across the faulted anticline.
Compressional Tectonics
Our discussion of compressional tectonic settings includes both high-angle reverse and thrust
faults since they commonly occur together in this setting. Also, this section centers on fold-and-
thrust belts, which include forearc, backarc, and collisional belts, as they make up the most pro-
lific compressional habitat. The most common hydrocarbon trap is the hanging wall anticline,
which includes such structures as fault propagation folds (snakeheads), fault bend folds, and
duplex structures (see Chapter 10). For example, in the Wyoming-Utah backarc fold-and-thrust
belt fields in the Rocky Mountains, USA (Fig. 8-78), nearly all the hydrocarbons are trapped in
the hanging wall of the Absaroka Thrust. They include such fields as Painter Reservoir, Whitney
Canyon, Ryckman Creek, and Anschutz Ranch Fields. Nearly all these fields are found in asym-
metric anticlinal folds with the steep limb to the east. Collisional zones, such as the Zagros col-
lisional belt of Iran, are some of the most prolific of all fold-and-thrust belt types. At one time,
the Zagros belt accounted for 75 percent of the world’s fold-and-thrust belt production.
446 Chap. 8 / Structure Maps
Figure 8-78 Maps showing the position of the Fossil Basin in the hanging wall of the
Absaroka Thrust Plate. (From Lamerson 1982. Reprinted by permission of the Rocky
Mountain Association of Geologists.)
An extensive literature search yields very few examples of contoured fault surface maps and
integrated structure maps for compressional structures. There are volumes of published fault and
integrated structure maps for extensional petroleum areas, but they appear to be scarce for com-
pressional areas. The inverse is true for balanced cross sections. A significant number of balanced
cross sections have been published for compressional structures, but very few are published in
extensional areas.
Fault surface mapping is not commonly done in compressional areas, although there is no
reason why it should not be; indeed, the construction of fault surface maps would aid in the
understanding of the geology. One possible reason for not making fault surface maps and inte-
grating them with structure maps is that, in many cases, hydrocarbons are trapped in the hanging
wall anticline, only slightly disrupted or controlled by reverse or thrust faulting. Therefore, fault
surface maps are not prepared. Another reason is that balanced sections are often constructed;
therefore, the construction of fault surface maps is assumed unnecessary. We believe, however,
that the mapping of all related faults is an essential part of any structural interpretation. In previ-
ous sections of this text, we showed that cross sections and seismic sections can misrepresent the
actual structure because of the orientation of the line of section. We also mentioned in Chapter 7,
with regard to the tying of seismic lines, that any nonvertical event can tie. The test of a valid fault
interpretation is the preparation of a fault surface map that is geologically reasonable based on
the available data.
We have shown that fault trace construction for any mapped horizon, particularly in areas of
steeply dipping beds, cannot be done intuitively. The accurate delineation of a fault trace for any
mapped horizon requires the integration of a fault map with the structure map for that horizon.
Remember, a fault trace on a specific horizon may be misleading as to the trend and shape of the
fault surface itself. The construction of a fault trace on a structure map based on a few well picks
and seismic lines is guesswork and is often incorrect. We therefore encourage the construction of
fault surface maps in compressional areas.
Mapping Techniques for Various Tectonic Habitats 447
Reverse Faults. Earlier in this chapter, we showed the techniques for construction of a reverse-
fault surface map and the integration of this map with a structure map (Figs. 8-25 and 8-26). In
this section, we look at intersecting reverse faults. Unlike the case involving normal faults, where
the fault appears as a horizon discontinuity on the map (a gap), no plan view horizon disconti-
nuity is present on a structure map with a reverse fault. The contours on the upthrown block
(hanging wall) overlie the contours on the downthrown block (footwall).
The example in Fig. 8-79 is drawn in a form similar to the normal fault examples except that
fault transport is reversed to show what happens due to compression. The pattern of intersecting
faults illustrates the mapping techniques, but the geometry may not be an accurate predictive
model. In other words, reverse faults on a plunging anticlinal structure may result from a volume
problem created as planar beds are forced into the compound curvature of the nose. Such faults
may not necessarily intersect, because the compression may be accommodated on different strati-
graphic levels by separate faults. Intersecting faults do occur in compressional areas, however,
and therefore understanding the mapping techniques for these faults is important.
(a)
Figure 8-79 (a) Fault surface map for reverse Faults A and B. (b) Integrated structure map on the 6000-ft
Horizon. Contours are dashed on the footwall block in the area of fault overlap for clarity. (c) Fault map super-
imposed onto the structure map to show the accuracy that is achieved by the integration of the two maps
regarding: (1) fault trace construction, (2) position of faults, (3) fault intersections, and (4) proper bed contour
construction across each fault. (d) Completed structure map on the 7500-ft Horizon. Contours dashed on foot-
wall in area of fault overlap.
448 Chap. 8 / Structure Maps
(b)
(c)
(d)
Figure 8-79 (continued)
Figure 8-79a shows the fault contour map for reverse Faults A and B. Since we are dealing
with intersecting faults, the line of intersection must be shown on the fault surface map. Also, the
additive property of faults must again be considered; throughout the area where two faults are
present (west of the intersection line in our example), Fault A has 125 ft of vertical separation
(repeated section in wells) and Fault B has 150 ft of vertical separation. Only Fault A with a ver-
tical separation of 275 ft (150 ft + 125 ft = 275 ft) is present east of the line of intersection.
Therefore, the vertical separation is conserved across the line of intersection.
Figure 8-79b is a completed structure map on the 6000-ft Horizon faulted by Faults A and
B. Figure 8-79c shows the fault map superimposed on the completed structure map. This figure
demonstrates the construction of the upthrown trace (trace on the hanging wall) and downthrown
trace (trace on the footwall) for each fault, the construction of the intersection of the traces of the
two faults, and the proper bed contour construction across each fault using the vertical separa-
tion. We have discussed in detail this type of construction with a number of other examples, so
we do not detail the construction here. Instead, we recommend that you pick at least two contours
and review the construction techniques used to complete this structure map. Notice that the fault
intersections fall on the line of intersection. Figure 8-79d is the completed structure map on the
deeper 7500-ft Horizon; take a moment and review its construction. The fault overlaps are wider
on this horizon because the faults and the horizon are dipping in the same general direction.
As with normal faults, the construction technique of integrating a fault surface map with a
structure map removes all the guesswork with regard to (1) the location of the fault on the mapped
450 Chap. 8 / Structure Maps
horizon, (2) fault trace construction, (3) the location of all fault intersections, and (4) the width
of the fault overlaps. The integration of the two maps constructs these features automatically
without the need for guesses or estimates.
Thrust Faults. Most thrust faults have displacements large enough to offset the two fault blocks
so that there is no continuity of structure (contour continuity) across the fault. In such cases, each
fault block must be contoured independently and then integrated with the fault map or maps.
Figure 8-80a is a diagrammatic cross section of the Mount Tobin Thrust Fault, Nevada.
Notice that there has been significant movement of the upper plate over the thrust fault.
Therefore, the upper and lower plates of the Ph Bed require independent contouring and fault
integration. Figure 8-80b shows the Mount Tobin Thrust Fault map superimposed on the struc-
ture map of the top of the Ph Bed in the upper plate. Observe in the upper portion of the figure
that the shape of the fault trace on the top of the Ph Bed is shown as a closed oval (in plan view).
This type of configuration is not arrived at intuitively; it requires the integration of the fault and
structure maps for accurate delineation.
Figure 8-80c shows the integration of the fault map with the lower plate (the trace on the
upper plate is also shown on the map). This example demonstrates the detail required to contour
the Ph Bed in the upper and lower plates and integrate these plates with the fault map to accu-
rately determine the position of all the fault traces. Take a few minutes and study the map to be
sure you understand how the two maps were integrated to arrive at the finished structure map.
Figure 8-81 is an integrated structure contour map, in two-way time, on the top of the
Mirador Formation within the Medina Anticline in the Eastern Cordillera, Colombia (Rowen and
Linares 2000). The Medina Anticline was interpreted and mapped based on surface outcrop, well
log, and seismic data.
The Medina Anticline is an excellent example of the integration of all data sources to
(1) generate a fault surface interpretation and map (Fig. 7-34), and (2) integrate the fault surface
with the structural interpretation. The Medina Anticline is a fault bend fold in the hanging wall
(a)
Figure 8-80 (a) Diagrammatic cross section of thrust fault (see Mount Tobin, Nevada,
USGS map series GQ7). (Modified from Bishop 1960. Published by permission of author.)
(b) Part of upper plate and fault surface map. Delineates the intersection of the fault with
the upper plate. (c) Overlay of fault map and lower plate to show the intersection of the fault
with the lower plate. (From Bishop 1960. Published by permission of author.)
Mapping Techniques for Various Tectonic Habitats 451
(c)
Figure 8-81 Structure contour map of the top Mirador Formation in the Medina Anticline, Eastern
Cordillera, Colombia. Thick black lines are axial surface traces interpreted on seismic dip profiles, with
arrows pointing in the dip direction. Thick, dashed black line represents erosional truncation at the
ground surface. Contour interval is 400 msec, relative to an arbitrary datum near the ground surface.
(From Rowan and Linares 2000; AAPG©2000, reprinted by permission of the AAPG whose permis-
sion is required for further use.)
Requirements for a Reasonable Structural Interpretation and Completed Maps 453
of the Aguaclara Fault in the southeastern thrust belt of the Eastern Cordillera in Colombia.
Active exploration in this area of Colombia has resulted in the discovery of several large hydro-
carbon accumulations such as Cusiana and Cupiagna. Although the Medina Anticline has well-
defined axial surfaces and nearly planar limbs, the fold appears more rounded than kink in shape.
In order to complete the interpretation and mapping for prospect generation, the fault surface and
structure map should be converted from time to depth.
Finally, remember we discussed the technique of preparing a pull-apart map for compres-
sional areas, or the use of a flap map, when mapping by hand. Such a map allows the contouring
of a given horizon all the way to its footwall cutoff, which may be a considerable distance from
the leading edge of the hanging wall block. With this technique, the mylar or paper used for map-
ping is cut along the leading edge of the hanging wall block, and new material is spliced in under-
neath to allow mapping of the subthrust surface to its fault cutoff. In complex areas with sub-
stantial thrust transport, this eliminates the confusion of dashed versus nondashed contour lines.
The maps can be separated to show the whole surface of the mapped horizon in both the upper
block and subthrust. They are also very helpful during isochore mapping.
If all these requirements are met, there should be a high degree of confidence that the inter-
pretation and completed maps are reasonable and accurate with respect to the available data.
vertical distribution of pay horizons, additional integrated structure maps at various depths may
be required.
Once the structural framework (fault and structure pattern) is established by constructing
several fieldwide or regional structure maps at various depths, the mapping of all pay horizons
between these initially mapped horizons becomes a much easier task.
Discontinuity of Structure with Depth. If the structural geology of an area maintains conti-
nuity with depth, the mapping of at least three horizons at various depths establishes and supports
the structural interpretation. However, in many instances the structural continuity is interrupted
with depth by one or more of three primary geologic events: (1) unconformities, (2) thrust faults,
or (3) listric growth faults. In these cases, the structural geology above the interrupting event most
likely will not conform to the structure below. Therefore, separate interpretations are required
above and below the interrupting event, in addition to multiple framework horizon mapping.
Figure 8-82 is a cross section through the Painter Reservoir Field in Wyoming. Observe that
the structural continuity is interrupted by two different events, first by the Sub-Evanston
Unconformity at about 3500 ft, and second at the Absaroka Thrust Fault. Notice how the struc-
ture of the area changes above and below each interrupting event. Such major events must be rec-
ognized to develop a good structural interpretation and are important in the search for hydrocar-
bons.
CHAPTER 9
INTERPRETATION OF
THREE-DIMENSIONAL
SEISMIC DATA
1. Developing the most reasonable interpretations for the area being studied. Faster data
access on the workstation allows multiple ideas to be tested.
2. Generating more accurate and reliable exploration and exploitation prospects. Three-
dimensional seismic data have created opportunities to visualize and analyze prospects
from multiple points of view to better understand risks. However, remember the cliché
“garbage in, garbage out” still holds true.
3. Correctly integrating geological, geophysical, and engineering data to establish the best
development plan for a field discovery. Workstations in use today allow unprecedented
integration of all types of data at a high rate of delivery.
4. Optimizing hydrocarbon recoveries through accurate volumetric estimates. Workstations
can often eliminate the need for hand-drawing maps, digitizing contours manually, and
calculating volumes by hand.
5. Planning a more successful development drilling, recompletion, and workover depletion
456
Introduction and Philosophy 457
plan. Successful companies have recognized the importance of the geoscience and
engineering staffs working together on a daily basis. Workstations enable fast decision-
making by sharing data among different disciplines (the beginning of a Shared Earth
Model).
6. Accurately evaluating and developing any required secondary recovery program. Here it is
imperative that geoscientists and engineers communicate closely. Sometimes a second
seismic survey is acquired over a producing field, allowing the team to detect any
unproduced areas or monitor a water flood program.
The Philosophical Doctrine, as described in detail in Chapter 1, is valid when work is done
with or without a workstation. It should be reviewed and understood before reading this chapter
because it provides the context and background principles to the techniques and workflows
described in the following sections. Even though most of the ten points in the philosophical
doctrine are self-explanatory, point 4 requires a bit of clarification relative to the workstation.
Point 4 states that “all subsurface data must be used to develop a reasonable and accurate sub-
surface interpretation.” In the 3D seismic world, it is usually impractical to interpret by hand
every line, crossline, time slice, and arbitrary line in the 3D data set. Methods are discussed in
this section on how to accomplish point 4 without costing your company valuable time waiting
for an interpretation. Workstations allow us to make interpretations faster than conventional paper
methods, but that doesn’t mean we should cut corners on quality. It means we are able to do more
quickly the quality geoscience or engineering that we need to do.
Project Setup
One of the first things to consider before starting on a workstation interpretation project is the
layout of the office or workroom where the work is to be accomplished. The workstation should
be located in a low traffic/low noise area. Lighting should be arranged in a way that does not
cause screen glare but allows other team members to work nearby with ease. Large worktables
should be close, allowing layout of maps, logs, or sections while the seismic data are being
worked or viewed. Frequent breaks from sitting at the workstation are required to reduce fatigue.
Chairs should be comfortable and adjustable. Monitors should be raised only enough so that the
head is tilted forward slightly in a relaxed position. A well-organized work area is an important
first step to an effective interpretation environment.
Figure 9-1 High contrast colors in outer midrange. Polarity preserved. (Published
by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-2 High contrast colors at extremes. Polarity preserved. (Published by per-
mission of Subsurface Consultants & Associates, LLC.)
460 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-4 High amplitudes emphasized. Low amplitudes suppressed. Polarity preserved
at extremes. (Published by permission of Subsurface Consultants & Associates, LLC.)
Introduction and Philosophy 461
Figure 9-6 Variable density display. (Pub- Figure 9-7 Wiggle trace display. (Published
lished by permission of Subsurface Consul- by permission of Subsurface Consultants &
tants & Associates, LLC.) Associates, LLC.)
462 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
horizons, and viewing the results in map view. For example, cascading menus, which are very
common, take longer than using icons or “hot keys.” The icon location on screen should be flex-
ible to minimize the distance the cursor travels to reach it. Anything that takes the interpreter’s
attention away from analyzing, interpreting, and mapping of the seismic data is a drag on work
efficiency and should be avoided. Small incremental time savers will build up large savings over
a project of long duration.
one or both horizons. Fault patterns from multiple horizons can be overlain on each other to test
the vertical compatibility of the fault interpretation and horizon mapping. If fault trends are not
compatible, this must be explained or corrected before proceeding with detailed target mapping.
If not done early, there can be much more later work that is necessary to clean up an invalid inter-
pretation. Once framework horizons are interpreted, mapped, and checked for vertical compati-
bility, it becomes much faster and easier to work additional target horizons internal to the exist-
ing framework.
every day during the project. With a good plan, the team stays on course to meet their objectives.
When roadblocks are encountered, they communicate and work together toward a solution. The
plan is not cast in stone but is adjusted periodically to reflect progress or changes in scope of the
work.
Project plans are a simple but effective tool to track progress and identify critical interde-
pendent tasks that need to be completed prior to others. This is especially helpful when team
members are relying on each other for completed maps, production data, or log correlations, for
example. A good plan acts like a road map to success for interdisciplinary work teams. Project
plans generally consist of a list of interdependent and independent tasks to be performed by each
team member. Tasks that are related or dependent on each other are connected in a way to show
the relative timing. Critical tasks are those that, if delayed, will delay the entire project. Plans can
be relatively simple, with a lot of flexibility, or very complex, depending on the scope of the
work. Whatever the case may be, you have heard it said: “Teams don’t plan to fail, they fail to
plan” (Tearpock 1997).
Documenting Work
One way to keep everything organized is to document your work in a notebook or in digital form
using an electronic journal of some sort. This will benefit you and those who follow. Even if an
area has a history, such as a previous interpretation, most people still want to do their own inter-
pretation. This is understandable. However, there is usually a time saving benefit from under-
standing what previous interpreters have done. This is more effective if the previous person has
taken good notes and is willing to share knowledge with incoming team members. Avoiding mis-
takes made by predecessors is a great time saving tool, and building on the good they have
accomplished increases efficiency during the project.
Documenting your work (Philosophical Doctrine point 10) seems like the “interpreter over-
head” we mentioned previously to avoid. However, this is not the case. Taking a few seconds now
and then to jot down a key file name or the steps taken to generate a particular display will help
jog your memory in the days to come. Keep track of parameters when the software won’t do it
for you. It helps with consistency and avoiding a “reinvention of the wheel” every time a process
is tried. Clearly, the work you perform for your company has value. Interpretation procedures and
methods are assets just like seismic data, well data, and unproduced reserves. They should be pre-
served and protected like any other asset.
FAULT INTERPRETATION
Introduction
The principles discussed in the following sections apply to any workstation software and are rel-
evant to extensional or compressional tectonic environments. The importance and art of making
fault surface maps is discussed in detail in Chapter 7. This section is intended to provide tech-
niques of fault interpretation and mapping for the workstation. A three-dimensionally correct
fault interpretation provides the solid foundation upon which all subsequent horizon interpreta-
tion and mapping are built. It is not enough to just pick fault segments on a series of parallel ver-
tical and/or horizontal seismic profiles. The resulting fault surface must make sense geologically
466 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
and geometrically, and it must be valid in 3D. Therefore, faults must not only be interpreted but
also contoured as a fault surface map and checked for validity.
Fault surfaces tend to be smooth, nearly planar, or arcuate surfaces. They typically do not
change radically in strike or dip unless the fault surface has been deformed. Vertical separation
varies in a systematic manner along a fault surface and can increase or decrease with depth.
Vertical separation and fault length are related. Obviously, large faults extend to greater lateral
distances than small faults, which are more local in character. The distinction between a long fault
and a series of small en echelon faults should be made, as in the following example. On widely
spaced parallel lines, we commonly interpret a fault segment on one profile that appears to belong
to the same fault as segments picked on other profiles. However, as we interpret the intervening
areas in detail, the faults are found to be either a series of disconnected en echelon faults or a
series of faults that extend laterally and actually coalesce to form a fault system. So what may be
interpreted to be a single fault may in fact be several separate faults or faults connected together
end-to-end. Figures 9-9 through 9-13 demonstrate how this would look on 3D seismic data.
Figures 9-9 through 9-11 are parallel dip lines from a 3D data set. Each fault is color-coded to
identify it in subsequent figures. There are similarities in the faulting on each dip line, but the true
relationship between the faults begins to emerge when viewed on a strike line (Fig. 9-12). Note
how each fault has a unique concave shape and how several faults are connected end-to-end,
forming a single fault system. Figure 9-13 shows the same relationship in horizontal slice view.
Notice in the upper right portion of Fig. 9-13 the small en echelon faults that do not connect to
form a single fault system.
Finally, it should be stated that without interpreting faults as surfaces, horizon mapping and
prospect generation become more challenging and risky. Fault surface maps and their integration
Figure 9-9 Dip line A-A′. Parallel to lines in Figs. 9-10 and 9-11. (Published by
permission of Subsurface Consultants & Associates, LLC.)
Fault Interpretation 467
with the structural horizons provide the best means for the location and orientation of the
upthrown and downthrown traces of a mapped horizon and the delineation of a prospect or reser-
voir.
Figure 9-12 Strike line D-D' along fault system interpreted in Figs. 9-9 through 9-11. (Published by
permission of Subsurface Consultants & Associates, LLC.)
Figure 9-13 Time slice showing fault system formed by faults F1, F2, and F3.
(Published by permission of Subsurface Consultants & Associates, LLC.)
Fault Interpretation 469
Reconnaissance
There is a tendency among some geoscientists to just dive right into a new data set and start work-
ing the data before determining exactly what needs to be done and at what level of detail. We rec-
ommend that at least some time be set aside at the beginning of an interpretation project to scroll
through the data in vertical and horizontal orientations in order to get a sense of which direction
the structures are trending and where future interpretation problem areas may exist. Seismic
attributes such as coherence are also a valuable reconnaissance tool. Figures 9-14 and 9-15
demonstrate how difficult it can be to interpret faults on amplitude time slices through a typical
3D data volume. Figures 9-16 and 9-17 show the same time slices, but they are now displayed
using a coherence attribute that enhances discontinuities. Fault trends and relationships are more
easily determined using this type of attribute display.
Creative color bars (palettes) used on variable density displays can enhance discontinuities
in the data and allow better fault identification. Some ideas for color schemes were shown previ-
ously, but geoscientists are not limited to those few. The goal is to identify faults as quickly as
possible, so use your imagination when it comes to use of color and various display scales. From
this reconnaissance, develop a plan to approach the fault interpretation. Take into account well
locations, any missing or repeated section-correlated in the well logs, and how much detail is nec-
essary to correctly map the faults in your 3D data set.
1. Display two or three seismic sections that pass through a well location but without the well
information displayed on the seismic data. Pick orientations that are based on your best
guess for the dip direction of the fault. These should form an X or asterisk (*) pattern on
the map (Fig. 9-18).
2. Look for event terminations that indicate the presence of a fault, and interpret a fault
segment that would represent a most likely location for the fault. If no fault is obvious, pick
several possible segments or pick another line orientation.
3. Now redisplay the seismic section with the well data and your picked segment. How close
do the fault segment and well-correlation fault data agree? The seismic fault segment
should tie the subsurface data. If it does not, consider the following five possible reasons
for the mis-tie:
(1) the time–depth relationship between the well and seismic is incorrect;
(2) the seismic interpretation is incorrect or needs adjustment;
(3) the fault pick in the well is incorrect or needs adjustment;
(4) there is an unresolved lateral or vertical velocity gradient between velocity control
points; or
(5) the well location is inaccurate (incorrect surface location, or erroneous directional
survey).
Any one of these explanations is equally likely. Begin by reevaluating the seismic interpre-
tation and attempt to adjust the interpreted fault segment. If the time–depth relationship appears
reasonable, and the fault interpretation cannot be reasonably altered to fit the well control, review
the well log correlations. Attempt to adjust the fault cut in the well to match the seismic inter-
470 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-15 Typical 3D time slice with faults interpreted. (Published by permission of
Subsurface Consultants & Associates, LLC.)
Fault Interpretation 471
Figure 9-16 Same time slice as Fig. 9-13, but displaying a coherence attribute. (Pub-
lished by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-17 Coherence attribute with faults shown. (Published by permission of Subsur-
face Consultants & Associates, LLC.)
472 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-18 Example of lines selected to begin reconnaissance for fault interpre-
tation. (Published by permission of Subsurface Consultants & Associates, LLC.)
pretation. If the correlation is reasonably certain, look for evidence of a velocity gradient. A
velocity gradient should affect all the correlation markers between two or more wells. If all else
fails, assume for a deviated well that the directional survey is in error, or in the case of a vertical
hole that it is not really vertical, or that the surface location of the well may be mis-spotted. Well
location problems are more common than you may suspect. Because fault surfaces typically dip
at much higher angles than bedding surfaces, errors in well location generate much larger dis-
crepancies in fault ties than they do in horizon ties. The key point is that all discrepancies
between the seismic data and the subsurface control must be resolved. Also, make sure the ver-
tical separation interpreted on the 3D seismic sections agrees with the vertical separation as deter-
mined from log correlation. After tying the well control, continue working this fault with one of
the following strategies.
Each strategy has its advantages, but the first two are probably used 80 percent of the time.
The multiple fault method is normally used after the single fault method in cases where the
remaining faults are difficult to sort out. Three-dimensional visualization is best used in complex
areas where fault conditions and trends change rapidly. Visualization software should always be
used as a quality-check tool for the faults and to refine a preliminary fault interpretation. As visu-
alization capabilities expand, more and more geoscientists are using this tool on a daily basis.
Figures 9-19 through 9-26 demonstrate the interpretation workflow for the single fault meth-
ods (strategies 1 and 2). Interpret a fault segment on the first vertical line (Fig. 9-19). Choose two
arbitrary lines that tie the first line (Fig. 9-20) and interpret the fault based on the tie with the
previous line. Two fault segments are interpreted on the two tied seismic profiles shown in
Fig. 9-21. Figure 9-22 shows the resulting fault surface based on the three interpreted lines.
Notice the fault surface passes very close to Well No. D1 in the south-central part of the map. Is
there a correlated fault in Well No. D1? Yes, Figure 9-23 shows the well tie and fault interpreta-
tion on a fourth line. This is all it takes to make a preliminary fault interpretation using strategy
1. Adding a time slice (Fig. 9-24) to the flow (strategy 2) is highly recommended to achieve a
better representation of the fault surface. The surface should be checked in strike view and addi-
tional interpretation added as needed (Fig. 9-25). The fault surface generated from the five inter-
preted lines and Well No. D1 is reasonable for a first pass (Fig. 9-26).
Using a series of intersecting seismic displays and interpreting one fault at a time, such as in
strategies 1 and 2, is generally the most efficient and effective methodology. It is important to
note that these are preliminary fault surfaces that need to be refined as necessary during horizon
interpretation. The result should be a good representation of the geometry of the fault, how far it
extends, and in what direction. It should be noted that as faults are interpreted on seismic data,
the map view of the resulting contoured fault surface should be updated. This allows quality con-
trol “on the fly” to catch any big errors such as abnormal changes in fault dip, strike, or miscor-
relation before getting too deep into the interpretation. Do not pick more segments than are need-
ed to define the fault surface. The key points are to determine where the major faults occur and
how they connect in the subsurface, if indeed they connect.
In complex areas, where there are several closely parallel faults or multiple fault intersections
in a small area, it may be necessary to interpret several faults at one time in order to sort them out
using strategy 3, the multiple fault method. Start by interpreting several unnamed fault segments
in vertical seismic views. These can be parallel or intersecting views (Figs. 9-27 and 9-28 are par-
allel profiles). Then view the interpreted faults in a time slice (amplitude or coherence attribute)
474 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-19 First fault segment interpreted on vertical seismic sections. (Published
by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-20 Map showing location of fault segment (magenta) and seismic lines selected
next (white). (Published by permission of Subsurface Consultants & Associates, LLC.)
Fault Interpretation 475
Figure 9-21 Second and third interpreted fault segments based on tie to original line.
(Published by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-22 Resulting fault surface based on three fault segments. (Published by permis-
sion of Subsurface Consultants & Associates, LLC.)
476 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-23 Fault segment picked at tie with Well No. D1 on profile parallel to dip
of fault. (Published by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-24 Fault surface projection onto time slice at 1580 ms. The quality of interpreta-
tion looks good. (Published by permission of Subsurface Consultants & Associates, LLC.)
Fault Interpretation 477
Figure 9-25 Strike view through Well No. D1 with interpreted fault segment. (Published by
permission of Subsurface Consultants & Associates, LLC.)
Figure 9-26 Resulting fault surface after interpretation of strike profile. Note the quality of
contouring with only five segments interpreted and a tie with one well. (Published by per-
mission of Subsurface Consultants & Associates, LLC.)
478 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-27 Seismic Line 1, using the multiple fault strategy. Two fault segments are in-
terpreted, but neither is assigned a name. (Published by permission of Subsurface Con-
sultants & Associates, LLC.)
Figure 9-28 Line 2 parallel to line 1. Fault segments are not assigned fault names.
(Published by permission of Subsurface Consultants & Associates, LLC.)
Fault Interpretation 479
view (Fig. 9-29). Look for trends where some of the segments that intersect the slice view line
up along seismic event terminations and appear to form a reasonable fault trace (Fig. 9-30). Next,
assign each fault a name and view the resulting contoured fault surfaces in map view (Fig. 9-31).
Add control from tie-lines to support the interpretation (Fig. 9-32). Choose one surface to com-
plete first (Fig. 9-33). If the fault surface appears valid, then proceed to the next fault.
In areas where it is important to sort out the finest details of a fault interpretation and the
data quality is sufficient, geoscientists should use some kind of cube or 3D visualization display.
This allows the data to be viewed in rapid succession in any orientation. Putting the data in
motion, so to speak, allows you to see subtle changes occurring over a small area. This process
can be time-consuming and at times unproductive, but it is well worth the effort when a prospect
is drilled and completed successfully. Many cube visualization tools are available. One example
is shown in Fig. 9-34.
Figure 9-29 Time slice showing intersections of fault segments interpreted on parallel east-west
vertical sections. Dip symbols show direction of fault dip. (Published by permission of Subsurface
Consultants & Associates, LLC.)
Figure 9-30 Segments assigned to a named fault. Note arrow pointing to possible miscorrelated
segment. It is unclear to which fault the segment belongs: the green fault or the, as yet unassigned,
segments to the north. (Published by permission of Subsurface Consultants & Associates, LLC.)
HORIZON INTERPRETATION
Selecting the Framework Horizons
The choice of framework horizons is important to the success of an interpretation project, so they
must be carefully selected. We recommend a minimum of three framework horizons: one shal-
low, one intermediate, and one deep. The intermediate horizon should be near the primary objec-
Horizon Interpretation 481
Figure 9-31 Map of the two contoured fault surfaces, which dip toward each other. Color
scheme from shallow to deep is orange to green to blue. (Published by permission of Sub-
surface Consultants & Associates, LLC.)
Figure 9-32 Interpret tie lines for 3D validity. Color scheme from shallow to deep is orange
to green to blue. (Published by permission of Subsurface Consultants & Associates, LLC.)
482 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-33 Choose one of the two fault surfaces to complete first. Final fault surface is
shown. Color scheme from shallow to deep is orange to green to blue. (Published by per-
mission of Subsurface Consultants & Associates, LLC.)
Figure 9-35 An example of a possible result when too many fault segments are
interpreted before the resulting surface is checked. This part of the fault surface is
not geologically reasonable. The entire interpreted fault surface is shown in Fig.
9-36. (Published by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-36 A fault surface that is not geologically reasonable and should not be
used in generating a structure map. More than 60 fault segments were used to
define this surface. (Published by permission of Subsurface Consultants &
Associates, LLC.)
484 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-39 Down-dip strike view. Segments picked on parallel lines only and not
loop-tied. (Published by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-40 Up-dip strike view. Note apparent abrupt change in strike, which could
indicate two faults or a deformed fault surface. (Published by permission of Subsur-
face Consultants & Associates, LLC.)
Figure 9-41 Map view of the fault surface as represented in Figs. 9-39 and
9-40. The change in strike direction can indicate the presence of two faults or
a deformed fault surface. (Published by permission of Subsurface Consultants
& Associates, LLC.)
486 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-42 Down-dip view with interpretation as two separate faults. Original fault also shown.
(Published by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-43 Up-dip view with interpretation as two separate faults. Original fault also shown. (Pub-
lished by permission of Subsurface Consultants & Associates, LLC.)
tive(s). The shallow horizon should be at or above the shallowest objective, and the deep horizon
should be at or below the deepest objective. If the structure is relatively simple and does not
change appreciatively through the objective zone, two framework horizons may be adequate. If,
however, there are stratigraphic and/or structural complexities, or the target horizons cover a large
depth interval, four or more horizons may be required to define the framework.
Framework horizons, if possible, should correspond to higher amplitude seismic events that
extend laterally over the entire study area. Objective reservoirs are often not good framework
horizons because they may change in reflection character over the area of interest. Clastic reser-
Horizon Interpretation 487
Figure 9-44 Map showing the final interpretation of the two fault surfaces. (Pub-
lished by permission of Subsurface Consultants & Associates, LLC.)
voirs, such as sandstone, often exhibit porosity variations and/or fluid variations, making these
reservoirs difficult to correlate over large areas on seismic data (unless you are working around
a mega-giant accumulation). Reservoirs with variable acoustic characteristics often yield struc-
ture maps that are not always representative of the true structure. You should be able to follow the
framework horizons around with confidence and relative ease over the study area. Also, the seis-
mic horizon should correspond to a reliable subsurface marker found in a majority of nearby
wells, which is usually a regional shale marker or a sand/shale sequence.
Unconformities must often be mapped as one of the framework horizons because the struc-
ture and rock properties are likely to be different above and below the unconformity. Angular
unconformities, onlaps, and downlaps sometimes generate complex reflections that are easily
identified but difficult to interpret as a horizon. Such an interface may need to be interpreted
entirely by hand, which could be a time-consuming process. The techniques described in the fol-
lowing section are intended to make the process of horizon interpretation more effective and effi-
cient.
coefficients in time (or depth), which have been computed from well log data (measured in
depth). The velocity function used to convert depth to time can be obtained from integration of
the sonic log and/or from a checkshot (or VSP) survey, from time–velocity pairs used in data pro-
cessing, or simply from time–depth pairs determined by the geoscientist through observation. The
primary purpose is to match the character of synthetic seismic traces derived from well logs with
the corresponding traces from the 3D seismic volume. Most of the time, sonic and density logs
are used to compute the acoustic impedance and reflection series. In cases where either, or both,
the sonic or density logs are unavailable, pseudosonic or pseudodensity logs can be generated
from resistivity (or other) logs. The quality of the tie between a synthetic seismogram and seis-
mic data varies from very good to no tie at all. However, the details of techniques for improving
the tie are beyond the scope of this book. One additional benefit of generating a synthetic seis-
mogram is that it indicates the reflection character of critical seismic reflections such as those that
occur at shale markers used as framework horizons and at tops and bases of reservoir sands.
Figure 9-45 shows an example of a synthetic seismogram.
Interpretation Strategy
Horizon interpretation is a rather straightforward process, but there are things that can be done to
make a more accurate interpretation in a reasonable amount of time. One important time saver
has already been mentioned: complete the preliminary fault interpretation prior to starting the
framework horizon interpretation. This is very important because it significantly reduces the pos-
sibility that work will have to be redone if something doesn’t fit. Also, we must remember that
most hanging wall structures are formed by large faults. As discussed in Chapters 10 and 11, a
geometric relationship commonly exists between fault shape and fold shape. Therefore, the more
the geoscientist understands the faults, the more likely the interpretation of the horizons will be
more accurate, geologically valid, easier to generate, and result in viable prospects.
Based on the fault interpretation strategies described previously, it is likely that framework
horizons will be interpreted, at least at first, on a different set of lines than the faults. Ultimately,
it may be desirable to have fault segments and horizons on the same lines and crosslines, but the
procedure to get to that point will be discussed later. In other words, preferred line orientations
for faults may not be preferred for horizons. With the speed of workstations and data servers ever
increasing, it is not necessary to compromise an accurate interpretation by selecting only one set
of lines.
Our basic philosophy can be summed up this way. Don’t interpret horizons in a way that will
take a lot of time to revise, and don’t be reluctant to revise things that do not fit or make sense
geologically. Once a lot of time has been invested in an interpretation, some geoscientists are
reluctant to back up and resolve problems that show up in mapping. Having a 3D survey on a
workstation does not change the basic need to tie lines to crosslines and resolve mis-ties, such as
would be done with a big eraser on paper copies of 2D data.
It is best to begin at a well (or wells) and work outward from there. When working areas
between wells, either of two approaches will work. The first is to use a combination of lines and
crosslines to tie data between the wells. The second is to use an arbitrary line directly in line with
the wells in question. In either case, there will be a cross-posted interpretation to use as a guide
while interpreting other lines and crosslines. When using arbitrary lines to initially “seed” an area
with interpretation, it is best to keep that horizon separate from the line/crossline horizon that will
be used for mapping. The reason for this is that on most 3D workstations it is difficult to edit hori-
zons on arbitrary lines once the focus has moved to other lines. Keeping them separate allows for
fast editing and clean up later. They can always be merged later if necessary for mapping. Figures
9-46 and 9-47 show examples of seismic lines that intersect wells in a 3D data set. Note that cor-
relation marker A in the wells corresponds to the blue seismic event (interpreted in green) that
will be the upper framework horizon. Figure 9-48 shows a nearly complete well-tie interpretation
done on lines and crosslines only. The same wells can be tied on arbitrary lines, as shown in Fig.
9-49. There are advantages to either method, and both can be done quickly. These ties should then
be used to work outward from the well control to complete the interpretation. Complete the ties
to the other framework horizons that are seen in the same wells. In the case of highly direction-
al wells, be sure to locate the seismic tie at the horizon penetration point in the wellbore.
How dense should the interpretation be when handpicking seismic horizons for structure
mapping? The answer is similar to the fault interpretation philosophy: only as dense as needed
to accurately define the surface. In our Philosophical Doctrine, point 4 states, “All subsurface
data must be used to develop a reasonable and accurate subsurface interpretation.” The way this
applies to 3D seismic data is probably not obvious. It is not necessary to interpret every line,
crossline, time slice, and arbitrary line. All orientations should be used as needed to ensure the
interpretation is valid in 3D. Well data must be tied to the seismic data. Engineering data, if avail-
able, should be considered as well. The handpicked interpretation performed on selected sections
should be preserved and not altered by automated processes such as autotracking or interpolation.
The latter should be written out to a separate horizon so results can be compared to your hand-
picked interpretation. If editing becomes necessary, it is much easier to work on a relatively loose
grid of data than to fix every line or crossline after infilling.
So, what is the optimum density of handpicked interpretation? That has already been
answered to some extent, but for example, if a 20-line by 20-crossline mesh of interpretation
yields an accurately contoured structure map on a simple unfaulted structure, then it is not nec-
490 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-46 Ties of seismic event with mapping horizon picks in wells. (Published by permission of
Subsurface Consultants & Associates, LLC.)
Figure 9-47 Another example of seismic-to-well ties for a horizon. (Published by permission of Sub-
surface Consultants & Associates, LLC.)
Horizon Interpretation 491
Figure 9-48 Map view of well ties on lines and crosslines. (Published by per-
mission of Subsurface Consultants & Associates, LLC.)
Figure 9-49 Map view of well ties on arbitrary lines. (Published by permission
of Subsurface Consultants & Associates, LLC.)
based on a 5 x 5-interpretation grid. Based on the 10 x 10 density in Fig. 11-50, the fault pattern
and the horizon dip between the faults could be considered reasonable. Figure 9-51 shows a
5 x 5 interpretation spacing. With the higher interpretation density, the faults and the horizon
shape are more accurately defined, and the difference in the fault interpretation is significant. The
three faults in the central area of Fig. 9-50 are reinterpreted on the 5 x 5 grid (Fig. 9-51) to be
three pairs of en echelon faults. The clues that the 10 x 10 interpretation is incorrect are the bends
in the fault polygons. A fault surface map of each fault would also show a bend in the contours,
which should alert you to the possible error. This is critical if fault-trap prospects are generated
along the faults in Fig. 9-50. The 5 x 5 interpretation takes more work than the 10 x 10, but it is
definitely worth the effort.
If prospects are located within narrow fault blocks and additional definition is needed, you
may need to work the even-numbered or odd-numbered lines and crosslines (a 2 x 2 spacing)
within the fault blocks, as in Fig. 9-52. There is no specific formula regarding how densely to
interpret 3D seismic data. It is up to you to decide, based on data, time constraints, and the use
for which the final maps are intended. Ultimately, the maps and the 3D seismic interpretation
should agree completely. It should be noted that the process described here involves some pre-
liminary structure contour mapping. Be aware that mis-ties in the interpretation or improper grid-
ding and contouring parameters will also affect the accuracy of the contoured map. The good
news is that if mis-ties are spotted early, they can be fixed with minimal effort. Gridding and con-
touring will be discussed later in this chapter.
Given what we have said about the density of interpretation, if a framework horizon is a
strong, laterally continuous seismic event, it may require only one or two interpreted lines in an
area or a fault block to allow computer autopicking, or autotracking, to infill the rest. If this is the
case, then go for it! If for some reason the autopick wanders off the event, it is better to spend
your time adding handpicked interpretation to the input and rerunning the infill rather than try-
ing to clean up after autopicking. However, autopicked horizons are notorious for being noisy and
generally do not contour smoothly.
Infill Strategies
There are many ways to infill a horizon accurately. The acceptability of the result depends on
data quality, structural complexity, desired objectives, and which software parameters are being
used. Four methods are discussed here. They are autopicking, interpolation, gridding, and hand-
picking. Before looking at the methods, these questions should be asked: “Why infill a horizon
at all? Can an accurate structure map be constructed without an infilled horizon?” Yes, if that is
the only objective, then that can be done without additional interpretation. There are many rea-
sons, however, why an infilled horizon may be necessary. It can be used as a reference for seis-
mic attribute extraction such as amplitude. It can be used for detailed structure mapping where
there are subtle complexities. It can be used as a reference for stratigraphic waveform classifica-
tion. Finally, it is more effective when performing layer computations such as isochores,
isopachs, isochrons, or flattening on seismic. The objective for using infilled horizons will deter-
mine which method is best suited to the task.
Handpicking each line is an option that should be used only when working (1) complexly
deformed or difficult areas such as very small fault blocks, (2) near fault intersections, (3) around
faults of limited extent that could affect viability of a prospect, or (4) within intersecting fault pat-
terns. These types of problems usually drive autopickers crazy, not to mention the difficulties they
create in map gridding and contouring. Detailed handpicking to fill in small areas is one of those
tedious tasks that should be used as a last resort but is necessary from time to time. Figure 9-52
is an example of detailed handpicking within narrow fault blocks.
Horizon Interpretation 493
Interpolation, particularly linear interpolation, is faster than handpicking every line. Simply
stated, with linear interpolation the computer considers two interpreted lines and mathematically
interpolates horizon values between them. This is usually done in a line or crossline direction.
The prerequisite for this to work is that interpretation must be done on parallel or subparallel lines
close enough to represent the characteristics of the seismic event. Interpolation can be used to
cross gaps in data or interpretation where no changes are expected to occur. It can be used as a
“quick and dirty” infill strategy when little time is available for more sophisticated techniques.
One drawback of using linear interpolation is the jagged-edge effect where features such as faults
cut diagonally across the 3D area (Fig. 9-53).
The next technique is based on gridding a horizon. Gridding is a process usually performed
when mapping (contouring) a horizon. A grid is computed from raw input data, which may be
irregularly distributed throughout the data set. Regularly spaced grid points are derived from sam-
ples of surrounding raw data points and represent the horizon surface topography. Grid parame-
ters can be selected to show every detail of an interpreted horizon or to act as a smoothing filter
to remove irregularities in the interpretation. Whatever your needs are for accuracy versus a visu-
ally pleasing map, find a set of parameters that works for you. Using a gridded horizon will usu-
ally yield the best looking map, but remember to always quality-check it against the seismic to
make sure it agrees with the data (Fig. 9-54).
Autopicking can be a very effective way to infill a horizon. Quality control is the key. Good
results depend on data quality, number of seed points, parameters selected, and software used.
Seed points are the hand-interpreted areas that autopicking software uses as a starting point for
infilling. Success is defined by how well the autopicked horizon tracks the seismic event. In high
continuity data areas, only a few hand-interpreted lines are necessary to seed a good autopicked
horizon. In complex or noisy data areas, more hand interpretation will be a necessary preliminary
for a horizon to be effectively infilled. In certain complex structural areas or in areas with much
incoherent data, autopicking cannot be used. To be effective as a timesaving tool, autopicking
should work quickly and accurately with minimal editing. Early autopicking software and hard-
ware were slow enough that interpreters had to start a job before lunch and return an hour later
to check results. Thankfully, the time involved has been significantly reduced. The most frequent
problem with autopicking is finding a balance between parameters that allows large areas to be
covered without the horizon wandering up or down a leg in the data. If tracking parameters are
too restrictive to limit leg jumping, only the most continuous events will be tracked, and some
areas will not fill in. Sometimes the compromise solution is to autopick only in areas of the data
set where results are good. Then, follow up with interpolation and/or handpicking to complete the
infill process. Autopicked horizons of high quality can be used for any purpose from structure
mapping to seismic attribute extraction to surface attributes or any layer-related function, such as
flattening or depth conversion (Fig. 9-55).
Accurate interpretation of framework and prospect horizons is the most important element
(after accurate fault interpretation) when making accurate structure maps. Efficient interpretation
techniques allow large data sets with many framework and prospective horizons to be evaluated
in a reasonable amount of time. By working in increasing levels of detail, interpreters can have
preliminary maps ready at an early stage for quick evaluation and then continue to work areas of
higher interest in more detail. In the next section, we discuss techniques to generate preliminary
structure maps that are based on the current horizon and fault interpretation. These “work in
progress” maps can be generated at any time to evaluate the status and quality of a current inter-
pretation.
Figure 9-54 Infill using a gridded data set. Note problem areas. (Published
by permission of Subsurface Consultants & Associates, LLC.)
interpretation process. They can be made either in time or in depth. Prior to making the first struc-
ture maps, geoscientists should have preliminary fault surfaces interpreted and mapped, and at
least part of one framework horizon interpreted. Preliminary structure maps are useful to quali-
ty-check an interpretation, such as for interpretation mis-ties, dip direction, fault location, and
prospect potential. They are also useful to test the 3D validity of the initial horizon and fault inter-
pretation. Other uses for preliminary structure maps are to check how accurately the wells are tied
to the seismic interpretation, how good the time–depth conversion is, and how much more detail
is needed to accurately define a horizon or fault surface.
Figure 9-56 Interpreted horizon cut by several faults, forming individual fault blocks.
(Published by permission of Subsurface Consultants & Associates, LLC.)
498 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
drawn polygons, but the computer does not always capture the details or a more geologically rea-
sonable depiction, which can be drawn by hand editing, as shown in Fig. 9-58. Fault gap poly-
gons are useful for a variety of reasons. Obviously, they represent an important component of any
structure map depicting the location of each fault trace and the width of the gaps. In the work-
station environment, they are also useful to control horizon infilling (discussed previously) and
computer contouring.
Table 9-1 Comparison of mapping parameters for Figs. 9-59 and 9-60.
The challenge to the geoscientist is to use available software tools to display several types
of map data in the same view. Some software makes this easier than others, and some tools allow
the process to be done in 3D. Geoscientists should be able to display a contoured fault surface
map on-screen with a contoured horizon surface map at the same contour interval. Also needed
are the fault polygons that define the fault gaps for that horizon.
Contours of equal value should be annotated or highlighted using color to view the inter-
section of the two surfaces. The point of intersection should correspond to a point on the fault
polygon.
The technique works best if one fault surface is integrated at a time to minimize clutter and
confusion. If only a slight adjustment is necessary, the fault polygon (trace) should be edited to
match the intersection of the horizon and fault surfaces. If a large discrepancy is discovered, more
interpretation work may be needed to determine the cause. Common causes for mislocated fault
polygons include sloppy or incorrect interpretation of the fault and/or horizon, incorrect correla-
tion of the fault gaps, and inconsistent depth conversion of the fault surface and the horizon sur-
502 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
face. The integration technique works on normal or reverse faults with large or small displace-
ment. Large faults may require that the upthrown and downthrown cutoff traces be edited indi-
vidually due to the large range of depth involved and practical limits of contour intervals. The fol-
lowing is a step-by-step description of the process.
Step 1. Display a fault contour (2700 ms), a horizon contour of the same value, and fault
polygons in the same map view (Fig. 9-61). The fault contours, the horizon contours,
and the fault polygon line should all intersect at the same point. If only a slight
correction is needed, edit (move) the polygon lines until they do intersect.
Step 2. Display a second contour (2710 ms) for the fault and horizon. Notice in Fig. 9-62
that both the upthrown and downthrown cutoffs represented by the polygons need
adjustment.
Step 3. Continue with as many contours and adjustments as possible on the same fault, then
repeat the procedure for other faults critical to the prospect or reservoir (Fig. 9-63).
Figure 9-64 shows the horizon and fault contours with the adjusted fault polygon.
Figure 9-65 compares the original polygon with the adjusted polygon. Note that in this
example the adjustments are very small (less than 150 feet). There are two points to be made here.
The first is whether or not to bother to adjust a fault polygon 150 feet. Assuming the fault and
structural interpretation to be accurate, the answer depends on how critical the fault trace posi-
tions are to well planning and prospect economics. We have seen a number of cases where a
directional well is planned to penetrate multiple objectives within 50 feet of a fault surface to
obtain the best structural position in a reservoir. Tens of millions of barrels of additional oil have
been recovered by accurately mapping faults and fault traces as shown here. Wells can then be
designed based on the fault interpretation to accurately hit the targets.
We have also seen wells miss a hanging wall objective and penetrate the footwall instead,
missing the target because the fault interpretation and maps(s) were off by as little as 100 feet or
less. So accuracy is important.
The bottom line is this: The process of integrating a horizon and a fault can be done quick-
ly and should be done on all faults critical to a prospect. The second point is that in areas of steep
dip, such as around a salt dome, this process becomes more critical because adjustments can be
much greater than in low dip areas. Seismic data can be of poor quality and, in that case, proper
integration becomes critical.
Finally, the ability to post certain well data is also very useful if the software allows. This
includes correlation depths (including restored tops) and fault picks with vertical separation
annotated. Most software permits display of only some of this information, but a “workaround”
is usually available. The more information you can display, without making the map unreadable,
the better.
Horizon and Fault Integration on a Workstation 503
Figure 9-61 Step 1. Horizon and fault integration, showing a fault contour of 2700 ms (red), a horizon con-
tour of the same value (yellow), and fault polygons (green) in the same map view. The red, yellow, and green
lines should intersect at the same point. (Published by permission of Subsurface Consultants & Associates,
LLC.)
Figure 9-62 Step 2. Horizon and fault integration. A second contour is displayed (in this case 2710 ms) for
the fault (red) and horizon (blue). Both the upthrown and downthrown cutoffs represented by the polygon
(green) need adjustment. (Published by permission of Subsurface Consultants & Associates, LLC.)
504 Chap. 9 / Interpretation of Three-Dimensional Seismic Data
Figure 9-63 Step 3. Horizon and fault integration. A third contour at 2720 ms (purple) is then integrated with
the polygon. (Published by permission of Subsurface Consultants & Associates, LLC.)
Figure 9-64 The adjusted polygon is shown in magenta. Contours of equal value intersect at the polygon
line. (Published by permission of Subsurface Consultants & Associates, LLC.)
Conclusion 505
Figure 9-65 Comparison of original polygon (green) versus adjusted polygon (magenta) with horizon con-
tours for reference. (Published by permission of Subsurface Consultants & Associates, LLC.)
CONCLUSION
We have discussed ways to increase efficiency by planning and organizing your work so that
important data are not lost and priorities can be clearly defined. We have also shown how to use
workflows to provide fast but accurate and consistent methodology to interpret and map data on
a workstation. Workstation technology provides unprecedented integration of geophysical, geo-
logical, and reservoir engineering data, leading to more efficient and effective prospect genera-
tion, well design, field development, and reservoir volumetric determinations. The philosophy
and techniques described in this chapter are meant to inspire geoscientists to make correct inter-
pretations in a timely manner that lead to accurate prospect maps and, ultimately, economically
successful wells.
CHAPTER 10
COMPRESSIONAL STRUCTURES:
BALANCING
AND
INTERPRETATION
INTRODUCTION
The first edition of ASGM contained one chapter on structural geologic methods. Since knowledge
of structural geology plays a key role in interpretation and mapping, as discussed in number 2 of
the Philosophical Doctrine, we believed the chapter was important to the overall content of the
textbook. Because of advances in structural geology and balancing during the past decade, in this
second edition we have expanded the one chapter into four separate chapters covering compres-
sional, extensional, strike-slip and growth structures. Knowledge of structural methods in these
various tectonic settings will improve your ability to generate viable, three-dimensionally valid
interpretations, maps and prospects, as well as improve your ability to develop field discoveries.
We begin the structural geology section of the textbook with a review of compressional tech-
niques and methods. Much of modern structural geologic analysis began with the study of com-
pressional tectonics, and therefore it is appropriate to start here. These four structural chapters
center around specific structural methods and techniques. A basic understanding of rock mechan-
ics, structural geology, and balancing, presented in such textbooks as Billings (1972), Suppe
(1985), Woodward, Boyer, and Suppe (1985), Marshak and Mitra (1988), is a prerequisite to
understanding and applying the techniques presented in this chapter.
506
Structural Geology and Balancing 507
of the data is the interpreter’s knowledge of compressional structural geology and the application
of techniques that lead to geologically reasonable interpretations and accurate maps.
One of the most important of the compressional structural geologic techniques is structural
balancing. The ultimate goals of balancing are to restore complexly deformed rock to its initial
state or to its correct palinspastic restoration and to determine the geologic sequence of events.
Such information can be very useful to the geologist or geophysicist. Not only is the geometry of
the structure better understood, resulting in better and more accurate prospect and reservoir maps,
but geologic trends such as sand patterns can be more accurately located. An understanding of
the timing of the structural events should aid in oil migration studies and define how and where
fluids may have entered the structure. If the geometry of the structure is understood, then this
knowledge can be used to more accurately process seismic data, which in turn results in an even
better understanding of the geometries. Balancing can also be effectively used to check assump-
tions and interpretations (Tearpock et al. 1994). Lastly, balancing tends to keep the interpreter
more focused. If the section does not balance, then perhaps it is time to reconsider the interpre-
tation. Why drill a well to determine that the interpretation does not balance when restoration can
determine a misinterpretation prior to the drilling? Our experience with balancing, as well as that
of our colleagues, indicates that balanced, geologically possible interpretations can discover sig-
nificant additional reserves. In short, balancing works.
Structural balancing is based on the intuitively satisfying concept that the interpreter must
neither create nor destroy volume during the interpretation process (Goguel 1962). Interpreters
may inadvertently introduce a volume imbalance anytime a fault is mis-picked or a horizon is
miscorrelated. Fortunately, balancing can detect volume problems prior to the drill bit. Thus, it
follows that an interpreted map or cross section, whether it be a geologic or seismic section,
should volumetrically restore without overlaps or voids in the stratigraphic section. Faulted and
folded beds should be restorable to their initial subhorizontal state (Tearpock et al. 1994).
Thereby, a structural interpretation may be tested for admissibility. An analogy might be a child
who removes a new block puzzle from a box and places it on the floor. Once all the pieces of the
puzzle have been removed from its container, the puzzle can be restored to its initial state by plac-
ing each block back into its proper position. The first attempt by the child at restoring the puzzle
may result in most of the pieces being placed into the box, with one or two pieces remaining on
the floor. A second attempt could result in all of the pieces being placed in the box, but with some
of the pieces being tilted at various angles or forced to fit.
The geoscientist experiences similar problems when attempting to retrodeform (restore) geo-
logic and/or geophysical data. Of course, the correct solution to a puzzle is one that has been per-
fectly restored to its initial position. There are two types of interpretations: interpretations that are
admissible, or geologically possible, and interpretations that are inadmissible, or not geological-
ly possible (Elliot 1983). A balanced interpretation is an admissible interpretation in which the
horizons can be restored to their initial subhorizontal position by unfolding the horizons and
rotating the beds back to a subhorizontal position along the interpreted faults.
The benefits of balancing are fundamental to correct geologic interpretations. The earth’s
subsurface contains no voids or mass overlaps; thus, a section that does not balance cannot be
geologically reasonable on simple geometric grounds. Unfortunately, a balanced section,
although physically reasonable, need not necessarily result in the correct geologic interpretation.
Balancing is not unique, and two geoscientists can produce two balanced sections that are not
alike. Obviously, the more complete the data set and the better the interpretive techniques, the
more likely that the balanced section will reflect reality.
508 Chap. 10 / Compressional Structures: Balancing and Interpretation
Balancing is still a developing science, and new techniques and interpretations are progres-
sively being introduced. Nevertheless, an interpretation tempered by a concept of mass conser-
vation is the key to admissible geologic interpretations and constructions. If the structural inter-
pretation is correct, then balancing techniques can be used to quantify the interpretation.
Balancing can be subdivided into two disciplines: classical balancing, which was primarily
developed by Goguel (1962), Bally et al. (1966), and Dahlstrom (1969) and his coworkers, and
nonclassical balancing, which was primarily developed by Suppe (1983, 1985) and his students
and coworkers. Most of the concepts presented in this introduction can be attributed to Goguel
and Dahlstrom.
MECHANICAL STRATIGRAPHY
For many years, structural geologists have argued about the mechanical properties of the upper
crust. Does it exhibit elastic and/or frictional behavior as indicated by earthquakes, or is it vis-
coelastic or viscoplastic, as indicated by the bent strata in the hinge zone of folds? Could time be
a factor? Do the sedimentary strata buckle out (Biot 1961), or do the strata follow faults within
the sedimentary section (Rich 1934)? Although all of these mechanisms are possible, the evi-
dence now strongly suggests that the deformation that occurs in petroleum basins is primarily
controlled by brittle (low temperature) deformation processes, and that the viscous deformation
expressed by fold trains (Fig. 10-1) is confined to metamorphic belts (Tearpock and Bischke
1980). The fold style depicted in Fig. 10-1, with its near constant wavelength, is not commonly
observed in petroleum basins, and thus another deformation mechanism is required to explain the
folds that trap hydrocarbons. This mechanism appears to be frictional deformation. Davis et al.
(1983), Dahlen et al. (1984), and Dahlen and Suppe (1988) formulate a frictional, or brittle, the-
ory of crystal deformation that applies to both compressional and extensional regimes. The the-
ory resolves the overthrust paradox (Smoluchowski 1909; Hubbert and Rubey 1959) and is con-
sistent with the geologic and seismic information collected from petroleum basins. Our intention
here is to apply this theory and its observations to our areas of interest. Those readers who main-
tain an interest in mechanics can consult the references listed at the end of this textbook.
The frictional theory of crystal deformation states that when folds form, the maximum prin-
cipal stress (σ1) is inclined slightly to the bedding surfaces (Fig. 10-2). The rock will then frac-
ture along angles that are dependent on the pore pressure and the intrinsic strength of the rock.
The weaker the rock, the lower the angle between σ1 and the fracture.
For example, consider an alternating sequence of limestone and shale layers (Fig. 10-2).
Intuitively, shale layers seem to be weaker than better consolidated limestone layers, and it is well
known that shales can contain abnormally high fluid pressures that drastically weaken these
rocks. The theory states that because shales are weaker than limestone, the angle (α1) between
σ1 and the fractures in shales must be smaller than the angle (α2) between σ1 and the fractures
Mechanical Stratigraphy 509
in limestones (Fig. 10-2). As σ1 is slightly inclined to the bedding, the fractures in the shales are
more subhorizontal than the fractures in the limestones. This leads to the primary conclusion of
this section: In more competent or stronger rocks, the fractures will form at a high angle to bed-
ding, and in the incompetent rocks, such as overpressured shales, the fractures tend to form par-
allel or subparallel to bedding.
Figure 10-2 Cross section of ramp geometry. For explanation, see text. (Modified after Rich
1934; AAPG©1934, reprinted by permission of the AAPG whose permission is required for fur-
ther use.)
If motion along these fractures causes them to coalesce, then a decollement, or zone of
detachment, will form along the flat-lying bedding and may follow incompetent (shale or evap-
orite) horizons for tens or even hundreds of kilometers (Davis and Engelder 1985). In areas where
the weaker layers gain strength or are pinched or faulted out, the decollement may ramp to a high-
er structural level (Fig. 10-2). As these ramps must pass through rocks that are stronger and have
lower pore pressures than shales, the angle (α2) between σ1 and the fractures will be larger. Thus,
ramps have higher angles with respect to the bedding than do the flatter portions of thrust faults
(Fig. 10-2).
Where the ramp connects to a weaker layer on a higher structural level, the ramp transforms
into a flat once again. Once a network of ramps and flats form and a large force is applied to the
back of the wedge-shaped region in Fig. 10-2, the strata above the flats and ramps will begin to
move along the fault. Material will begin to slide along the flats and up the ramps, forming a fold
in the hanging wall block. Eventually, large folds will form in a manner that was initially
described by Rich (1934), but this process is the subject of a later section.
The angle at which the ramp steps up from the bedding is called the cutoff, or step-up, angle
(θ in Fig. 10-2). This angle is often characteristic, or fundamental, to a particular fold-thrust belt
and depends on both the pore pressure in the rock and the rock type. Similar relationships may
exist in extensional terrains. The characteristic cutoff angle in certain fold-thrust belts is general-
ly less than 20 deg and tends to vary within several degrees of its mean value. For example, in
510 Chap. 10 / Compressional Structures: Balancing and Interpretation
Taiwan the characteristic step-up angle is 13.3 deg +/– 2.4 deg (Suppe and Namson 1979; Dahlen
et al. 1984). An attempt must be made to determine this angle prior to a balancing study. This
step-up angle will be used to balance your structures. Note that the step-up angle is measured rel-
ative to regional dip rather than to the horizontal.
There appear to be at least three methods which give insight into estimating the characteris-
tic step-up angle. Field studies or a literature search can be conducted in the area of interest. As
the step-up angle is the angle between the flat and the ramp, field measurements or a description
of this relationship will provide the required answer. A second, less direct measurement technique
is to observe a well-imaged ramp and flat on a seismic section that is perpendicular to the strike
of the fault surface. The section must first be depth-converted to make this measurement. The
strata above the ramp will parallel the ramp, and thereby the step-up angle can be determined rel-
ative to regional dip (Fig. 10-33). Therefore, a study of the dips across an area may give insight
into the characteristic step-up angle. For this method, it is first necessary to know the regional or
undeformed dip of the area. For example, suppose that an area has no regional dip. It therefore
follows that the nontilted beds will have zero dip. Strata that have moved up ramps, and are
deformed, may dip at 12 deg. The characteristic step-up angle is therefore 12 deg. We might,
however, be faced with a situation in which 20 percent of the dips are near zero, 30 percent are
3 deg, and 50 percent are about 9 deg or greater. The problem here is attempting to decide
whether the regional dip is zero or 3 deg and whether the step-up angle is 9 deg or 12 deg or
greater. This matter is often resolved by finding that one of these choices simply works better than
the other during the restoration, or balancing, process.
widespread volume removal, regional balancing and structural analysis may indicate that anoth-
er process is occurring and to what extent. We normally find, however, that these volume reduc-
tion processes are not a major concern and that the interpreter normally can think in terms of vol-
ume conservation while being prepared for alternatives.
The economic issue that needs to be addressed here is much more practical and much more
likely to confront the interpreter on a daily basis than is pressure solution. Interpreters often
unknowingly have a tendency to introduce mass overlaps and gaps into their interpretations
(Tearpock et al. 1994). Often, these gaps or overlaps are confined to a particular region of their
cross sections or to a particular structure. For example, a given seismic-based cross section and
prospect, upon retrodeformation, has twice as much volume between sp (shotpoint) 320 and sp
420 (at about 1.5 sec to 2.2 sec) and no volume between sp 285 and sp 400 (at about 2.8 sec to
3.1 sec). An obvious question thus arises: Does this volume incompatibility affect the viability of
the prospect, and would a better interpretation enhance or detract from the prospectivity of the
area? Therefore, balancing literally attempts to take the “holes” out of our interpretations, as is
shown in the retrodeformation section in this chapter.
Area Accountability
In the Mechanical Stratigraphy section, we describe the petroleum basin as a low temperature
regime subject to brittle (i.e., frictional) deformation. In such an environment, flow, elongation,
and flattening are not of primary importance, and thus the 3D volume problem can be reduced to
2D. In other words, we shall assume that material is not entering or leaving the plane of the geo-
logic cross section, and therefore the problem can be reduced to 2D (Goguel 1962). Notable
exceptions to this rule would be shale and salt diapirs, which are typically 3D phenomena. These
salt structures, which are associated with withdrawal and rim synclines surrounding the diapir
(Trusheim 1960), contain a wealth of information that defines salt flowage and can be used to
balance salt diapirs in 3D. Another exception is the bifurcating normal fault structure, which
moves material out of the plane of cross section. Techniques for studying this type of deforma-
tion are briefly addressed in Chapter 11. In the meantime, however, and as long as the deforma-
tion is brittle and the transport direction is subperpendicular to fault strike, the 3D problem can
be reduced to a 2D cross section that is subperpendicular to the strike of the fault.
Bed Length Consistency
If we accept the premise that petroleum-bearing rocks are brittle and deform at temperatures
within the hydrocarbons window, then the 2D problem can be linearized (Goguel 1962). In other
words, if there is no large-scale material flow within or across the plane of the 2D cross section,
then the seismic reflection or bed length before deformation will remain the same after deforma-
tion (Fig. 10-3). This logic will also hold true for the thickness of each bed involved in the defor-
mation, which means that the folding will be of the parallel type. Thus, bed length can be utilized
to balance cross sections. If a sedimentary sequence is 2 km long before deformation, it must
remain 2 km long after the deformation. The bed may be bent and it may be broken, but it should
still be 2 km long.
Although the logic inherent in the above statement may seem self-evident, it appears to be
one of the primary causes of the so-called “balanced” cross section, which is prevalent through-
out the literature. The above logic implies that if one measures the bed lengths across a cross sec-
tion, and the bed lengths are equal on all levels, then the cross section will balance. In practice,
512 Chap. 10 / Compressional Structures: Balancing and Interpretation
however, small changes in the lengths of lines can result in significant volume changes that result
from inaccuracies in, or a lack of, subsurface dip data. This follows from the trigonometric rela-
tionship that at low angles the length of the adjacent line is about equal to the hypotenuse (Fig.
10-4). Consequently, we can see that the line segment AB is about equal to AC, even though the
thickness AX is not equal to the thickness CZ. Therefore, we can often check existing cross sec-
tions by simply observing whether beds or formations are subject to unexplained or nonuniform
thickness variations. If these thickness variations are not due to logical variations in stratigraph-
ic thickness, then the interpretation should be subjected to further analysis.
Figure 10-3 Deformation map of petroleum basins versus metamorphic belts. Low tempera-
tures tend to preserve cross-sectional volume, whereas in metamorphic belts, material will flow
in and out of the plane of cross section.
Figure 10-4 Noticeable changes in bed thickness result in small changes in bed length.
Classical Balancing Techniques 513
Pin Lines
A significant development was made by Dahlstrom (1969), who realized that you can check the
validity of any cross section by measuring bed lengths, while keeping an eye out for variations in
the thickness of units. This is accomplished through the use of pin lines (Dahlstrom 1969). In this
procedure, one attempts to locate regions that are not subject to deformation (such as shear or
bedding plane slip) and then affix these regions to the basement by driving an artificial pin ver-
tically through the cross section. Pins are used as a basis for measurement, and bed length con-
sistency is then measured relative to these pin lines (Fig. 10-5). Dahlstrom realized that bed
length consistency must be preserved on all structural levels in both 2D and 3D, and that if the
bed length consistency does not hold from one section to another, then the interpretation is like-
ly to be in error. Figure 10-5 is modified from Dahlstrom (1969) with Fig. 10-5a signifying the
undeformed pin state. If the unit is concentrically folded and displaced a distance S, then the bed
length (lo) within the concentric fold after deformation should be the same length (lo) as it was
before deformation (Fig. 10-5a and b).
In Fig. 10-5b, the bed length (lo) within the folded unit is not the same as the pin length (l).
This follows, as the folded unit has been shortened a distance S (compare Fig. 10-5b and c). In
Fig. 10-5c dipping beds overlie flat beds, which is the classic indication of a geometric disconti-
nuity or decollement (thrust fault). We call this method for picking thrust faults Dahlstrom’s Rule,
and the thrust fault exists between the steeper dipping and the flatter dipping beds (Fig. 10-5c).
Thus, when picking thrust faults on seismic data, simply look for steeply dipping beds over more
gently dipping beds. These steeply dipping beds must be structurally deformed and typically are
inclined at more than 5 deg to regional dip.
Line Length Exercise
Line-length balancing can be a powerful quick-look tool (Tearpock et al. 1994). We present an
example of how line-length balancing may find additional oil in producing fields. Figure 10-6a
represents two dip profiles that are similar to those in a large producing trend in South America.
The two profiles are from the same field, traverse the same anticlinal structure, and are a short
distance from each other. Good to fair quality seismic data from the field image the top of the
structure, but do not clearly image Faults A, B, and C. Well No. M-5 on profile A and other wells
in the field cut Fault A, but Fault B is inferred from the relatively dense well control (Fig. 10-6a).
Notice that Wells No. M-1 and M-3, which penetrate the front of the structure on profiles A and
B, encounter the reservoir section at a greater depth than do the structurally higher Wells No.
M-2 and M-4. Seismic data from an adjoining field on the same structure are of good to excel-
lent quality and clearly image Fault C, which is a bedding plane thrust. Fault C was mapped into
the area of profiles A and B from the adjoining field.
The interpretation shown on profile A contains three imbricate blocks formed by Faults A,
B, and C. Faults A and B link to large Fault C. The footwall reservoir section has the same bed
length along profiles A and B, so pin the structure at the hanging wall cutoff position located in
the structurally lowest imbricate blocks (left-hand pin). The pin on the right penetrates the syn-
cline in an off-structural position. In the hanging wall portions of the fold, use a balancing pro-
gram or a ruler to measure the bed lengths of the reservoir bed along its top. The beds are cut by
the faults, so the top of the bed in each imbricate block terminates at the faults. Therefore, do not
include as bed length the distance along a fault. On profile A, the hanging wall bed lengths in the
three imbricates are about 11.8 km total.
514 Chap. 10 / Compressional Structures: Balancing and Interpretation
t BED
X X X X X X X X X X X X X X X X X X
(a)
BED GTH
o LEN
t
(b)
o
t
X X X X X X X X X X X X X X X X X X
(c)
Figure 10-5 Pin lines and bed length consistency. (a) Undeformed bed state. (b) and (c)
Deformed bed. (Modified after Dahlstrom 1969. Published by permission of the National
Research Council of Canada.)
Classical Balancing Techniques 515
Repeating the bed length measurements on profile B, located a short distance from profile
A, results in a hanging wall bed length, at the top of the reservoir horizon, of about 10.8 km.
Thus, between the two profiles there is a line length imbalance along the top of the reservoir
bed, and profile A contains 1 km more bed length than profile B. Perhaps the faults are dying
out, but the profiles are near the center of the trend, which is over 100 km long. Over short dis-
tances, the slip along faults is not likely to change significantly along strike (Dahlstrom 1969;
Elliot 1976; see Bow and Arrow Rule in the Cross Section Consistency section of this chapter).
How may we reconcile this line length imbalance between the two profiles, and what are the
implications?
Notice on profile A that the reservoir horizon is repeated in Well No. M-5. Abundant well
log data from the field demonstrates that Fault A dies out before reaching profile B. In fold-thrust
belts and over short distances, the slip along thrust faults is about constant along strike (see
Cross Section Consistency). Thus, it is unlikely that Faults A and B would both grow smaller over
such a short distance. Alternatively, slip transfer between faults is common in fold and thrust belts
(Dahlstrom 1969) (Fig. 10-14). The slip on Fault A may transfer to Fault B. In other words, as
Fault A dies out, the slip on Fault B increases at the expense of fault A.
What are the consequences of a 1 km slip transfer between the two fault surfaces, and how
could this slip transfer affect reserves? If Fault B is larger than shown in profile B, then Fault B
may overthrust a larger portion of the lower imbricate block penetrated by Well No. M-3. We pro-
ceed to line-length balance the data, and present an alternative interpretation of the data shown in
profile C in Fig. 10-6b. Profile C contains an additional 1 km of bed length relative to profile B,
so that the bed lengths on profiles A and C are both about 11.8 km. The interpretation shown on
profile C uses the concept of a ramp-flat fold geometry that is common to fold-thrust belts (Bally
et al. 1966), rather than the upward-listric reverse fault shown on profiles A and B. Upward-listric
fault surfaces are common to extensional terranes (Chapter 11). As line-length balancing con-
cepts suggest that the bed length should be about 11.8 km on the two profiles, and as we must
honor the existing well control, we consider the solution shown in profile C. Profile C contains
an additional 1 km of slip on imbricate Fault B. This increase in slip creates more repeated sec-
tion in the lowermost imbricate block beneath Well No. M-4. This interpretation of the data is
exciting, as the new interpretation extends the reservoir horizon in the lower block between Faults
B and C by about 1 km to the right, introducing upside potential. This potential exists up-dip of
the producing Well No. M-3. The solution shown in profile C may require a reinterpretation of
profile A. This example shows how line-length balancing may find new oil in old fields.
Balancing sections using the structural workstation (see the following section Computer-
Aided Structural Modeling and Balancing) is an alternative to manual line-length balancing pro-
cedures. Profile D in Fig. 10-7, generated on a structural workstation, uses area-balancing con-
cepts. Profile D not only maintains line-length balance, but also cross-sectional area balance (see
the section Area Balancing in this chapter). Therefore, profile D is more geometrically accurate
than profile C in Fig. 10-6b. However, the two profiles are similar.
Computer-Aided Structural Modeling and Balancing
Structural analysis, interpretation, and modeling rely heavily on the graphical representation of
structural horizon and fault surface geometry. Using structural workstation software, the end
product of this graphical representation results in the construction of cross sections. Graphical
methods of structural analysis can be applied to geologic data to determine the viability of cross
sections. Historically, structural modeling relied heavily on manual drafting to create cross
516 Chap. 10 / Compressional Structures: Balancing and Interpretation
Profile A
SEAL
FAULT B
FAULT A
M-3 M-4
Profile B
SEAL
FAULT B
FAULT C
0 km 1 km 2 km 3 km 4 km 5 km
(a)
Figure 10-6 (a) Profiles A and B constructed across an anticline that forms a producing field.
The slip imbalance between the two profiles creates a line-length imbalance, as described in
text. (b) Profile C represents a reinterpretation of profile B using line-length balancing concepts.
Profile C, which uses a ramp-flat thrust fault geometry common to fold-thrust belts, introduces
additional potential in the reservoir in the lowermost imbricate block. (Published by permission
of R. Bischke.)
Classical Balancing Techniques 517
M-3 M-4
Profile C
SEAL
FAULT B
FAULT C
0 km 1 km 2 km 3 km 4 km 5 km
(b)
Figure 10-7 Profile D is a reinterpretation of profile C in Fig. 10-6b, using structural workstation
methods based on balancing concepts. Profile D is similar to the line-length balanced profile C.
(Published by permission of R. Bischke.)
518 Chap. 10 / Compressional Structures: Balancing and Interpretation
sections. The emergence and enhancement of computer workstations during the 1990s provided
a powerful tool for 2D and 3D structural evaluation. The workstation facilitates the visualization
and modeling of structural data and allows interpreters to attack more complicated structural
problems (Fig. 10-8). Utilizing workstation software, it is possible to move quickly from the time
domain of seismic data into the depth domain of structural visualization. Depth visualization by
geoscientists enables the creation of a more complete and accurate depiction of the subsurface
structural geology. The technical and economic benefits of computer-aided structural analysis are
important, if not key, to the success of petroleum exploration and production in structurally com-
plex areas.
After reviewing the different structural styles presented in Chapters 10 through 12 and their
associated algorithms, one may ask, “What is the best and most effective method of applying
structural information?” One important approach is the proper use of structural workstation
software.
Seismic data are the primary subsurface information; therefore, it is critical to translate seis-
mic time models into seismic depth models. Once depth intervals are selected and assigned
respective velocities, the structural workstation software should provide a means to readily move
between the time sections and the related depth domains. Data quality and knowledge of related
acoustic interval velocities determines the accuracy of the time–depth transition. Again, the work-
station is an excellent tool for testing different time–depth pairs. Iteration of structural models
utilizing an array of alternative concepts helps to refine and perfect the interpretation, which is
another strong justification for the implementation of computer-aided structural analysis.
Animated models of fault bend folds, fault propagation folds, and so on are possible on the
workstation. These animated models are helpful when visualizing and constructing forward mod-
els of simple structures and illustrating the origin of structures. The identification and accurate
depiction of fault surfaces from seismic data sets are one of the most important steps of seismic
structural interpretation. There is a direct relationship between the geometry of the fault surfaces
and the geometry of structure horizons related to the fault surfaces. The relationship between fold
shape and fault shape is often overlooked by many geoscientists during the seismic interpreta-
tion phase of a project. We believe this is often due to a limited structural background by geo-
scientists, which restricts their understanding of fault-fold relationships. Interpretation errors
related to the geometry of faults and horizons become obvious when viewed in the form of a bal-
anced cross section. Risk can be reduced significantly by using comprehensive, balanced 3D
structural models.
A validated 3D structural model is not only kinematically correct, but also helps to eliminate
any errors of interpreted displacement along selected fault surfaces. The elimination of displace-
ments that are kinematically incorrect creates higher quality interpretations. Whereas a balanced
3D structural interpretation may not be unique, it does add substantially to the validity of any
interpretation. From structural workstation analysis, it can be readily seen that the term balanc-
ing encompasses validation, retrodeformation, and/or restoration (see the section Retrodeforma-
tion in this chapter). The complexity of retrodeforming a structural cross section manually may
be difficult if not impossible in many cases, yet it can be readily and accurately completed with
a computer.
Two-dimensional and three-dimensional structural workstation software can significantly
expand the interpretive capacity and accuracy of the geoscientist. Software links provide direct
communication between structural applications and other geophysical and geological software
programs. Accuracy, efficiency, and completeness are improved by the sharing of data in a work-
station environment.
Classical Balancing Techniques
Figure 10-8 Structural cross section across Savanna Creek Duplex
and Canadian Rockies. (Published by permission of D. Roeder.)
519
520 Chap. 10 / Compressional Structures: Balancing and Interpretation
A comprehensive structural model incorporates all the available geologic and geophysical
data for a given area. In most cases, structural analysis forces the geoscientist to “fill in the
blanks” beyond the limited available information. The good data areas can be readily projected
into the poor data areas. Workstations can access and store volumes of data beyond the reason-
able capacity or efficiency of manual manipulation.
Accurate dip analysis, sonic logs, lithology logs, deviation surveys, and all other well data
are incorporated into an accurate structural interpretation. Detailed surface geology maps, includ-
ing topography, provide a wealth of information for land-based study areas. All stratigraphic data
are an integral part of a comprehensive structural interpretation. Computer-aided structural analy-
sis enables you to analyze all your data accurately and completely. The accuracy and reliability
of subsurface maps are enhanced and perfected with a detailed, computer-generated structural
model.
Structural modeling is the keystone to subsurface modeling and visualization. Therefore,
from an industrial point of view, technical and economic success ties directly to the accuracy and
effectiveness of the subsurface structural interpretation. The structural analysis not only provides
the framework for detailed production activities, but it also drives frontier exploration. Pre-seis-
mic structural models are cost-effective prospecting tools during the initial phases of a study.
Structural models can help in planning and guiding a seismic acquisition program and can aid in
improving the quality of acquired seismic data. Digital cross sections and assigned interval veloc-
ities lend themselves to ray-tracing programs and resultant models to help facilitate the planning,
acquisition, and interpretation of seismic data. The economic success of a new discovery or the
cost of a dry hole dwarfs the cost of a proper structural evaluation. The process of structural mod-
eling and restoration forces the geoscientist to critically think about the interpretation, to ques-
tion the data, and to understand the hydrocarbon potential of the region. The computer-based
structural interpretation allows geoscientists to quickly and accurately converge on viable geo-
logic solutions to complex structure problems.
Retrodeformation
In a previous section on classical balancing techniques, we introduce a number of powerful rules
and constraints to check interpretations. These rules concerning preservation of line length and
bed thickness can be quickly applied to cross sections to insure cross-section viability. We now
demonstrate that line length and thickness preservation is an important first step in a two-step
operation of retrodeformation.
In the introduction to this chapter, we emphasize that, with time, structures move, and that
structural interpretations should be restorable. The process is called retrodeformation, or
palinspastic reconstruction. Any interpretation of subsurface data should be restorable to an ini-
tial undeformed state because the stratigraphic units were deposited parallel to regional dip.
Faults induced by compressional forces may cut the strata, causing the hanging wall beds to move
over footwall beds. The structure is thrust forward and into its present position. Let us assume
that this structure is presently imaged on seismic profiles. The retrodeformation process is the
reverse of the forward-thrusting process. Any interpretation of the faults contained in this seismic
data set should be compatible with the hanging wall beds moving back along the fault surface
into their undeformed state. The pieces of the seismic puzzle should be restorable without mass
overlaps or voids. These principles apply to every tectonic regime, but they are most easily
applied to compressional and extensional regimes. However, the retrodeformation principle is an
excellent consistency check on interpretations of compressional, extensional, strike-slip, and salt
Classical Balancing Techniques 521
structures. We apply line length and bed thickness preservation concepts to a seismic line to show
how these concepts can improve prospect integrity.
Examine Fig. 10-9, which is taken from Bally’s (1983) classic monograms on seismic inter-
pretation entitled “Seismic Expressions of Structural Styles.” In the forward to his monograms
Dr. Bally states, “As to the interpretations presented, the reader will have frequent occasion to
disagree or to be unconvinced of the interpretation offered. This properly reflects the fact that
seismic reflection profiles are not easily interpreted in a unique way. Because the marked seismic
lines are frequently supporting published papers, less critical readers often feel that such illus-
trations constitute geological proof, while in reality they are much more like drawings on a seis-
mic background that illustrate an author’s concept” (our emphasis). Dr. Bally’s statement has
many important consequences to industry, so let’s examine his statement in more detail.
Dr. Bally makes several important points that management, accountants, economists, and
working teams should remember every time geoscientists propose a multimillion dollar well.
Economics dictates that wells are expensive and that geoscientists are cheap, and not the other
way around. Money should always be available to test the viability of all prospects prior to
drilling (Tearpock et al. 1994).
The other concept inherent in Dr. Bally’s forwarding statement is that there are two sets of
interpretations: those that constitute “geologic proof” and those that constitute “drawings.” We
call the first type of interpretation an admissible interpretation (Elliot 1983). An admissible inter-
pretation maintains 3D structural validity and is a geologically possible interpretation. The sec-
ond type of interpretation is the inadmissible interpretation that does not maintain 3D structural
validity and is therefore impossible on simple geometric grounds. Chapters 10, 11, and 12 con-
centrate on admissible interpretations as applied to prospects and prospect evaluation. With this
in mind, we next test Fig. 10-9 for its admissibility.
Often, during a prospect review and evaluation of compressional structures, we first check
for apparent horizon thickness changes. For example, in nongrowth environments horizons
should not change thickness across fault surfaces. The eye is very sensitive to vertical thickness
changes, and with a little practice can readily detect problems in the time domain. Notice on the
time profile in Fig. 10-9, within the front limb of the structure between sp 125 and sp 175, that
the section between the top of Pierre and the top Permo-Pennsylvanian strata apparently thick-
ens. Could this thickness variation result from higher velocity rocks thrust over lower velocity
rocks or, alternatively, from imbricate thrusting? Time profiles are not geologic profiles and are
subject to geometric distortions. In order to remove the geometric distortions, the time section
needs to be digitized and depth-corrected on a workstation.
Notice on depth-corrected Fig. 10-10a that the thickness variations within and beneath the
fault zone are exaggerated in the depth domain. These thickness changes are more pronounced
within the “fault zone” (refer to Fig. 10-9) that was interpreted in order to retrodeform the struc-
ture. The bed dips in the “fault zone” exceed 40 deg. An interpretation of the depth-corrected
section strongly suggests that the “fault zone” in Fig. 10-9 results from high bed dips that are
common to compressional terranes. In compressional regimes, high bed dips can result in time
sections that dramatically distort structures, and we strongly recommend that all interpreta-
tions be analyzed in the depth domain. The time section in Fig. 10-9 bears little resemblance to
the depth section in Fig. 10-10a.
Figure 10-10b, which represents a restoration of the depth interpretation in Fig. 10-10a,
shows regions of area imbalance and contains voids in the undeformed state. The horizons change
thickness across the restored faults, particularly in the Pierre (Kp) and Niobrara (Kn) units. This
522
Chap. 10 / Compressional Structures: Balancing and Interpretation
Figure 10-9 Time profile of a fold from the Colorado Rocky Mountains. Beneath the “fault zone,” dipping Niobrara reflec-
tions over flatter Dakota sandstone reflections may indicate a detachment near the level of the Dakota. (From Bally 1983;
AAPG©1983, reprinted by permission of the AAPG whose permission is required for further use.)
Classical Balancing Techniques 523
indicates a violation of the bed-thickness conservation rule. On a properly restored thrust fault,
the beds will maintain approximately constant thickness across the restored structure. This fol-
lows because the sedimentary units were deposited parallel to a gentle regional dip. One of the
reasons the structure does not area-balance is that no detachment exists to produce the dipping
beds above the “flat” Dakota and top Permo-Pennsylvanian strata (between sp 125 to 175 on
Fig. 10-9).
How can we improve the interpretation? Refer to Fig. 10-11, a profile from the Canadian
Rocky Mountains (Bally et al. 1966). In the Moose Mountain sheet and in the lower central
portions of the profile beneath Bow Valley is a structure that resembles the one in Figs. 10-9 and
10-10. In Fig. 10-11 the thrust fault is observed to ramp beneath the western limb of the fold and
flatten beneath the structure’s eastern limb. This ramp-flat fault geometry is consistent with high-
quality seismic data and is observed in outcrops (Boyer 1986). We use this geometry to reinter-
pret and balance the structure in Fig. 10-9. As mentioned previously, the structure does not bal-
ance due to the lack of a detachment located between the level of the dipping Niobrara and the
flatter Permo-Pennsylvanian formations. Applying Dahlstrom’s rule for picking thrust faults (dip-
ping beds over flatter beds) to the time or depth section, we proceed to balance the structure. The
results, shown in Fig. 10-10c, require an imbricate fault block, or horse, which is common to
fold-thrust belts (Boyer and Elliot 1982). This solution is interesting in that the structure could
possess additional hydrocarbons on the level of the repeated Dakota sandstone within the horse.
A ramp-flat fault geometry, when applied to Fig. 10-9, results in an admissible interpretation, as
shown in Fig. 10-10c.
Lastly, we convert Fig. 10-10c back to the time domain in Fig. 10-10d. You can now com-
pare Fig. 10-10d to the original time section (Fig. 10-9).
All interpretations of prospects have consequences, which may influence the success of a
project and the interpretation of the petroleum system. In Fig. 10-9 a possible fault trap exists
beneath the fault zone in the upturned beds of the Dakota sandstone. Fig. 10-10c indicates that
the trapping fault may not exist and that the Dakota strata maintain stratigraphic thickness and
may not turn up beneath the proposed fault. This affects prospect risk. The balancing software
also predicts the position and thickness of the horizons that are missing from Fig. 10-9.
Figure 10-10c predicts that the thrust fault beneath the fold continues toward the northeast
to possibly link to other prospects in the petroleum system (Boyer and Elliot 1982). Figure 10-9
suggests that no such link exists in the system, which also affects migration risk.
Picking Thrust Faults
Picking thrust faults on seismic sections is not as straightforward as it may seem. This subject is
complicated because thrust faults are typically “thin skinned” and may follow, or parallel, bed-
ding surfaces over long distances (Rich 1934; Bally et al. 1966).
A major insight into picking thrust faults came from the Canadian Rockies, where petrole-
um structural geologists noticed in outcrops of thrust faults that steeply dipping beds overlie flat-
ter dipping beds (Bally et al. 1966). In the discussion of pin lines, we show this bed dip discor-
dance or discontinuity in Fig. 10-5c, and called this method for picking faults Dahlstrom’s rule
(Dahlstrom 1969). The method works for both dip lines and strike lines.
In Fig. 10-12, we can observe a thrust ramping to the left of the fold hinge. The dashed line
represents an axial surface, which bisects the limbs of the syncline. The outcrop is perpendicular
to the strike of the fault and therefore in the dip direction. To the left of the synclinal axial sur-
face, steeply dipping beds overlie flatter dipping beds, showing a discontinuity and a thrust fault.
524 Chap. 10 / Compressional Structures: Balancing and Interpretation
(a)
T Tertiary
Tl Lower Tertiary
Tl
Tl Kp K. Pierre
Kn K. Niobrara
Kp Kp
Tl Kb K. Benton
Tl Kp Kn Kd K. Dakota SS
Kn
Kn Jm Jur. Morrison
Kp Kn Kd
Kn Kd Pl Perm. Lykins
Kd Pen f Penn. Fountain
Pl
Pl Pz Pz carbonate
Pl
Pen f
Pen f
VOID
(b)
Figure 10-10 (a) Depth-corrected interpretation of time profile shown in Fig. 10-9, generated using structural interpreta-
tion software. The depth-corrected figure suggests a much tighter fold than the horizontally stretched seismic profile (Fig.
10-9). In the depth domain the frontal limb fold geometry contains unusual thickness changes above the Dakota sand-
stone. Fault zone on Fig. 10-9 correlates to region of high bed dips in this figure. (b) Retrodeformed Fig. 10-10a contains
voids and formation thicknesses that do not match or are not uniform across the interpreted faults. This mismatch indi-
cates area and thickness imbalances. (c) Reinterpretation of Fig. 10-9 using workstation software and structural princi-
ples. Unnatural thickness changes shown in Fig. 10-10a indicated an area imbalance that may contain an untested horse
block. This figure area-balances and is restorable. (d) Balanced section Fig. 10-10c converted to the time domain. This
figure can be compared to Fig. 10-9 to check for consistency. (Published by permission of R. Bischke.)
Classical Balancing Techniques 525
(c)
TWT:
seconds
0
Kp
Ti
Kn
Kd Pl
Kn
1.0 1.0
Penn Kd
Pl
THRUST Penn Penn
Penn
2.0 2.0
Pz
(d)
0' S.L.
8,000' 8,000'
16,000' 16,000'
24,000' 24,000'
0 2 4 6 8 10
TERTIARY MESOZOIC CLASTICS LOWER PALEOZOIC CARBONATES 1 MOOSE MOUNTAIN SHEET
MILES
CARDIUM UPPER PALEOZOIC CARBONATES PC SHIELD 2 McCONNELL SHEET
The thrust in the left part of Fig. 10-12 represents the ramp, or the area near axial surface BY in
Fig. 10-33c. Alternatively, dipping beds over flat beds can also be observed at the front of the
fold, or the region between axial surfaces AX and A' X' along the upper flat in Fig. 10-33c. A sim-
ilar relationship exists on seismic lines in the strike direction of the fault.
Figure 10-12 Ramp in a thrust fault from the Canadian Rocky Mountains. Dipping
beds over flatter beds and the synclinal axial surface define the structural ramp.
(From Boyer 1986. Published by permission of the Journal of Structural Geology.)
Figure 10-13 is a spectacular strike-line profile imaging the lateral termination of a fold-
thrust belt in Eastern Venezuela. The profile images several thrust faults that peel back and fold
the younger cover rocks along a back-thrust fault that forms along the top of a triangle zone (see
Triangle Zone and Wedge Structures in this chapter). On the profile between sp A and B the dip-
ping beds image a feature called a lateral ramp. The discontinuity between the dipping and flat
beds defines the main thrust. The largest thrust is interpreted to be in the more poorly imaged
region at sp A and 2.3 seconds.
A question may arise as to how to distinguish stratigraphic dips from structural dips. We first
refer to Rich (1951), who found that clinoforms in the steeply dipping portions of deltas rarely
exceed 5 deg. Second, clinoforms reflect downlap or toplap. Thus, if the seismic reflections are
folded and dip at angles exceeding about 5 deg, then the dips are most likely structural and not
stratigraphic. We recommend that you use a 3D workstation to scan the strike direction looking
for lateral ramps along the flanks of a fold. If such a strike ramp is located in the data, pick the
fault on several strike lines and construct an initial fault surface map. To complete the fault inter-
pretation and fault surface map, tie the fault surface to the dip lines.
1.0
2.0
Figure 10-13 Seismic section oriented parallel to strike in a fold-thrust belt in eastern Venezuela. This time profile images several thrust faults that move
material toward the observer. Our interpretation is that the location of the main fault is positioned at the bed dip discontinuity defined by dipping beds over
flatter beds. (Published by permission of Corpoven.)
Cross Section Construction 529
section to cross section. However, the slip can decrease to zero as the result of deformation in the
cores of folds.
For example, if a cross section of a complex structure exhibits three thrust faults with a total
of 3 mi of slip, then it is very likely that a nearby cross section will also contain three thrust faults
of similar shape and form that also contain about 3 mi of slip. If these three thrust faults radical-
ly change position and/or shape, then some intervening transverse structure must exist to accom-
modate the deformation. Such intervening structures are called transfer structures, and these
structures exist in compressional (Dahlstrom 1969) as well as tensile extensional environments
(Gibbs 1984). Transfer structures often occur as tear faults, or cross faults, which form at high
angles to the major structural trend. Furthermore, these transverse structures are often responsi-
ble for changes in the trends and shapes of structures from cross section to cross section. Figure
10-14 illustrates a transfer by lateral shear from one fault bend fold to another. In Fig. 10-14a, the
displacement on Fault 1 is compensated by displacement on Fault 2 (see left side of diagram Fig.
10-14a). The sum of the displacements on Fault 1 and Fault 2 remain constant; thus, as the slip
on Fault 1 decreases, the slip on Fault 2 increases. The amplitude of the folds above the faults
also change in a like manner. On profile F the slip on Fault 1 is equal to the slip on Fault 2, and
the folds that form above the two faults have the same amplitude. The resulting structures caused
by the lateral shear are shown in map view in Fig. 10-14b. The result is that the fold on Fault 1
plunges to the south and is replaced by the fold on Fault 2, which plunges to the north. This slip
transfer between folds is very common in fold-thrust belts (Fig. 10-15).
Therefore, we see that small changes are permissible from cross section to cross section, but
how much change is possible? Elliott (1976) answers this question with the Bow and Arrow
Rule (Fig. 10-16). This rule states that the amount of displacement can vary along a fault zone,
but at an amount equal to 7 percent to 12 percent of its strike length. For example, suppose you
mapped deformation along a large thrust fault zone that has a total length of 10 mi. From the Bow
and Arrow Rule, one would predict that the maximum dip-slip motion on the fault would be on
the order of 0.7 mi to 1.2 mi. Next, assume that the amount of displacement along another fault
is known to increase to a maximum along a 10-mi portion of the fault zone. We can now predict
not only that the fault is at least 20 mi long, but also that there are at least 1.4 mi to 2.4 mi of dip-
slip motion on this fault. Elliott (1976) developed the Bow and Arrow Rule for thrust faults, but
a similar relationship may exist for normal faults, particularly for faults in excess of 10 mi in
length (Morley 1999). Merret and Almendinger (1991) studied 562 faults from different envi-
ronments and found that
The Bow and Arrow Rule is based on scaling laws, and it follows that laterally restricted
faults have small displacements, whereas only laterally extensive faults have large displacements.
Figure 10-14 Transfer zone from one fault bend fold to another. (Published by permis-
sion of Ted Snedden.)
Cross Section Construction 531
(b)
Figure 10-14 (continued)
532
Chap. 10 / Compressional Structures: Balancing and Interpretation
Figure 10-15 Radar aperture image of Appalachian fold-thrust belt near Harrisburg, PA, showing en echelon arrangement of plunging
anticlines. This displacement transfer is common to fold-thrust belts. Although repeated section exists in the well logs from this area,
notice the near absence of surface faulting. The absence of surface faulting is common to portions of many fold-thrust belts, where the
deep thrusts occur as blind or as bedding plane thrust faults. (Published by permission of the United States Geological Survey.)
Cross Section Construction 533
Figure 10-16 Bow and Arrow Rule. Slip perpendicular to fault strike is approximately 10
percent of the fault length. (Modified after Elliott 1976. Published by permission of the
Royal Society of London.)
1985; Boyer 1986). Both methods assume that the folding is parallel; i.e., stratigraphic unit thick-
ness remains constant (in the absence of more detailed information). The Busk or the kink
method can be used to extrapolate any type of dip data. It is important, however, to be consistent
in the use of the data. For example, the top of a stratigraphic unit is projected to the top of an
adjacent unit only if the units being mapped do not change thickness, which is commonly the case
over short distances. A dipmeter recording within a stratigraphic unit is not projected to a dip-
meter reading in an adjacent well unless these recordings are on the same stratigraphic level. In
other words, it is important to understand that you are projecting time-stratigraphic surfaces
across the structure.
Busk Method Approximation
The Busk method (Busk 1929) assumes that the folds are parallel (constant-thickness of strati-
graphic units) and that they are concentric; i.e., the folds consist of segments of circular arcs.
These arc segments are used to project data to depth. Normally, dip data measured from surface
outcrops, well logs, or seismic sections will not lie along the plane of cross section. Thus, the data
must be projected to the plane using the methods discussed in Chapter 6. Let us assume for sim-
plicity that the data, measured from outcrop, are shown in Fig. 10-17a. The data points are usu-
ally defined on specific stratigraphic unit tops or bases. Normals (lines perpendicular to dip) are
drawn downward from the position of the dip measurement data. These normals intersect at a
point that represents a radius of curvature for an arc (point O in Fig. 10-17b), which is used to
project the stratigraphic data in the area between the two data points A and B. A compass cen-
tered at point O is extended so that it has a radius OA, and then an arc is constructed from point
A to line D (Fig. 10-17c). This procedure is then repeated for point B, using radius OB. The
results of this exercise are two concentric arc segments, AE and FB, which define a curved layer
534 Chap. 10 / Compressional Structures: Balancing and Interpretation
AE-FB, of constant thickness AF, or EB. If another data point G is introduced (Fig. 10-17d), the
normal to this adjoining data point will intersect line segment OB at a different location, point O,'
and now several different radii (O' B, O' G, OI) are used to complete the stratigraphic extrapola-
tion. In Fig. 10-17e, a well with dipmeter data is added and more normals and arcs are drawn to
depict a more complete fold.
The method can be visualized as consisting of several adjoining regions, or domains, in
which the curvature of the beds is constant, and at the intersection of these domains the curva-
ture of the beds changes. The Busk method is therefore a curved dip domain method. It suffers
from an inability to retrodeform easily and to correctly project the front limb of a fold into the
adjoining syncline.
Kink Method Approximation
The next method that has proven extremely useful for extrapolating data to depth or along a cross
section is the kink method, or constant dip domain method (Faill 1969, 1973; Laubscher 1977;
Suppe and Chang 1983). In the Busk method, bed dips that were mutually related were assumed
to represent a common curvature domain. However, we could have just as readily bisected the
angle between the dips from two adjacent dip data points and created two regions of constant dip
related to the two data points. In the limit, or where the data are closely spaced, both methods
would be identical.
As shown in Fig. 10-18a, the first task in the kink method is to project the bed dip data in
cross section. For example, the dip at point B is projected in the direction of bed dip data point
A. Next, place two triangles adjacent to each other so that the upper triangle (X) is parallel to bed
dip A and can be moved over the lower triangle (Y). (If preferred, a parallel glider can be used in
place of two triangles.) Now move the upper triangle upward past the bed dip data point B and
construct a line CD so that point D is approximately halfway between bed dip points A and B
(Fig. 10-18b). When working with real data, point D need not be halfway between points A and
B, and its position will depend on where the beds change dip. This position can often be deter-
mined from outcrop or depth-corrected seismic data. Bisect the angle between lines CD and DB
(a) (b)
Figure 10-17 (a) - (e) Busk Method Approximation. The strata are projected to
depth along segments of circular arcs. (Modified from Marshak and Mitra 1988.)
Cross Section Construction 535
(c) (d)
(e)
with a protractor or compass, and then project the dip data at A to the dip domain boundary line
with the triangle (line AE, Fig. 10-18c). Move the triangles to a new position so that one of them
is parallel to dip data point B, and move this triangle down to continue line AE into the domain
of dip data point B (line EF, Fig. 10-18c). The projection process results in two dip domains with
each domain containing a constant dip and a theoretical interval of constant thickness (DE).
Repeat the process as additional data are introduced (Fig. 10-18d). Notice that in Fig. 10-18d, dip
domain B converges and terminates at point O, which is called a branch point. A branch point
occurs at the intersection of two axial surfaces. A dip domain is eliminated at a branch point, in
this case dip domain B. Only two dip domains exist beneath the branch point, whereas three
domains exist above the branch point. Notice that the axial surfaces bisect the bed dip domains
both above and below the branch points. It is important to remember to bisect the angle between
the fold limbs and not the angle between the axial surfaces.
In many folded areas, extensive regions of relatively constant dip adjoin smaller regions of
rapidly changing dip. This is commonly seen on seismic sections. These relationships suggest
that many folds possess limbs that have a uniform or near-constant dip, but have hinge zones that
are curved. As a result of this uniformity in dip, the kink method is readily adapted to work in
low temperature fold belts.
When applying the constant dip domain method, always remember to bisect the angle
between the bed dips, thereby creating two adjoining and individual dip domains. Usually, the
data are generalized or averaged to eliminate aberrant data points. This can be accomplished by
taking two triangles and aligning them so that the top triangle can be passed across the data. In
(a) (b)
Figure 10-18 (a) - (d) Kink method approximation. The sedimentary beds are projected to
depth along planar surfaces. The method applies to the majority of folds, which possess sub-
planar limbs.
Cross Section Construction 537
(c)
(d)
this manner, the triangle can be used as a filter to generalize or average the data. Areas of differ-
ent generalized dip are defined as individual or separate dip domains, and the dip is then assumed
to be approximately constant within each domain. The method also works very well with depth-
corrected seismic or well data. The bisection procedure is in fact the continuity principle as
applied to balancing (Suppe 1988) (Fig. 10-19):
Notice that if the fold does not change thickness across the axial surface, then t1 = t2 and α1
= α2. If this procedure is judiciously applied, the cross section is more likely to line-length bal-
ance and area-balance.
When mapping using the kink method, you will find that as the stratigraphic intervals change
thickness, the theoretical structural level of the interval as predicted by the kink method will devi-
ate from the observed level. Thus, periodic adjustments in bed thickness must be made, usually
at the position of the axial surface, which is the dip domain boundary line (Fig. 10-18c). Our pref-
erence is to follow the observed stratigraphic unit or sequence boundary in regions of onlap, etc.,
even though this results in a divergence of once-parallel lines. If units above the unconformity do
not change thickness dramatically, little harm is done by accurately representing the strata.
In areas of good data, the bisected dip domain data will ensure proper line length and area
balancing. In regions where the data are poor or nonexistent, the kink method can be used to proj-
ect the units being mapped. Even under these conditions, the uniform thickness assumption can
be a very powerful tool. Assume, for example, that you are mapping units A and B in Fig. 10-20
from the north but that you encounter a region where no data exist. Mapping toward the no-data
area from the south results in a good match on unit A but a poor match on unit B. What would
you conclude in this case? The mismatch could result from either a dramatic change in thickness
or an unrecognized fault in the south-central area that ceased growth prior to the deposition of
unit A.
Figure 10-20 Example utilizing the uniform thickness approximation. Major change in
the thickness of Unit B, but not in Unit A, implies that a structure or a stratigraphic
change is present in the region that lacks data.
SEAL OIL
WATER OIL
SEAL
WATER
CONSTANT
THIN THICKNESS
LIMB
(a) (b)
Figure 10-21 Different fold interpretations can result in different proposed well locations.
(a) Interpretation of a fold based on surface dip and seismic data, but not using the kink
method. An attempt was made to maintain the vertical thickness of the beds within the
steeply dipping front limb of the fold (see Fig. 10-22). (b) Interpretation of a folded structure
based on surface dip and seismic data and using the kink method. An attempt was made to
maintain the stratigraphic thickness of the beds within the steeply dipping front limb of the
fold. (Published by permission of R. Bischke.)
Geoscientists who work fold-thrust belts know a majority of the folds within hydrocarbon-
producing regions approximate parallel folds rather than thin-limb folds (Suppe and Medwedeff
1990; Tearpock et al. 1994). Unless data exists in support of a thin frontal limb, the parallel fold
interpretation is likely to be the better interpretation.
If the fold is a constant-thickness fold, then the likely result after drilling the two wells is
shown in Fig. 10-22. In the figure, the two well positions shown in Fig. 10-21a and b are redrawn
on the constant-thickness fold shown in Fig. 10-21b. The well on the right is positioned using the
thin frontal limb interpretation. This well is likely to encounter steeply dipping beds in the seal
horizon and never test the reservoir. Perhaps the geoscientists who generated the profile shown
in Fig. 10-21a believed that a seismic time profile is a geologic profile. Time profiles distort
geometry and the distortion increases with increasing bed dip (Chapter 5).
Notice on the profile shown in Fig. 10-21b that if the well were drilled deeper, it might have
crossed the axial surface and entered the front limb of the structure. When drilling asymmetric
folds, there is always the risk of crossing the axial surface that separates the gently dipping back
limb from the steeply dipping front limb. If the front limb of the fold is slightly overturned, then
beneath the axial surface, the stratigraphic units penetrated by a well will become younger with
increasing depth (Fig. 10-39). Drilling the syncline in front of the fold is also possible when
attempting to exploit asymmetric folds. The fault propagation fold is the second most common
type of fold in fold-thrust belts, so interpreters should be aware of the pitfalls associated with
asymmetric folding (Tearpock et al. 1994).
The profile shown in Fig. 10-21b illustrates the kink law. The kink law states that if the beds
do not change thickness, then the axial surface bisects the limbs of the fold. In other words, on
constant-thickness folds the angles between the two fold limbs and the axial surface are about
Cross Section Construction 541
CORRECT INCORRECT
WELL POSITION WELL POSITION
=
1 2
SEAL OIL
RESERVOIR CONSTANT
THICKNESS
Figure 10-22 As most folds are constant-thickness folds and obey the kink method, wells spud-
ded on the crests of asymmetric folds will typically intersect steeply dipping beds within the front
limbs of these folds. On asymmetric folds, wells spudded on the back limbs are more likely to
discover hydrocarbons. This cross section was generated using the kink method, and thus the
axial surface bisects the fold limbs. (Modified from Tearpock et al. 1994.)
equal, or γ1 = γ2, as in Fig. 10-22. Most petroleum-related folds come close to obeying a con-
stant-thickness relationship and the kink law (Tearpock et al. 1994). At the correct well position,
shown on Fig. 10-22, the well was positioned so that it did not cross the axial surface at the reser-
voir level. This well intersects the reservoir horizon, whereas the dry hole (Fig. 10-21a) crosses
the axial surface (Fig. 10-22). A well that crosses an axial surface can even penetrate vertically
dipping or overturned beds.
Figure 10-23 is redrawn from Fig. 10-21a to demonstrate that the angles between the two
fold limbs and the axial surface are not equal, or γ1 is not equal to γ2. This is an indication that
the fold was constructed as a thin-limb fold, and therefore the front limb may be incorrectly locat-
ed. On the other hand, if the fold is actually a constant-thickness fold, then the well will cross the
axial surface and penetrate the steeply dipping beds in the front limb of the fold, as shown in Fig.
10-22.
The kink law is a powerful tool when constructing cross sections. Remember to construct
the cross section on a scale of one-to-one. To apply the method, simply bisect the angle between
the fold limbs. These procedures eliminate geometric distortions and provide a clearer picture of
the complex relationships concerning folded structures.
As a final exercise, examine the three profiles of folded structures from three different fold-
thrust belts in Asia, shown in Fig. 10-24. Seismic and surface dip data constrain the profiles.
Using the kink law, can you recognize which one of the three wells was drilled on structure, and
why? Remember to bisect the angle between the fold limbs. The kink method is easy to use and
rapid to apply, and often generates cross sections that accurately predict wellbore results. The
method deteriorates if the projection crosses a large fault. The method assumes that the bed dips
542 Chap. 10 / Compressional Structures: Balancing and Interpretation
INCORRECT
INTERPRETATION
75o
≠
1 2
OIL
SEAL
WATER
RESERVOIR AXIAL
SURFACE THIN
LIMB
Figure 10-23 On a thin-limb fold the axial surface does not bisect the limbs of the structure.
This geometry contrasts with Fig. 10-22 in which the axial surface bisects the fold limbs.
(Modified from Tearpock et al. 1994.)
remain about constant within each dip domain. However, bed dip may change significantly where
crossing large faults and thus violate the constant bed dip assumption. The target depths of the
three wells are between 1500 ft and 3500 ft, and the cross sections are drawn at a scale of one to
one.
Figure 10-24a shows a well spudded into the front limb of a symmetric monoclinal-type
fold. A high-quality seismic line crosses the fold that images a steeply dipping west limb, a flat
crestal area, and a more gently dipping east limb. Surface bed dips exist to aid the interpretation.
Were high or low bed dips encountered in the well? Using the outcrop data, we know that an axial
surface would bisect the angle between the 30-deg to 35-deg front limb dips and the flat crestal
dips. We position the axial surface along the change in bed dips that image on the seismic pro-
file, which results in an axial surface that dips steeply to the east (Fig. 10-25a). If seismic data do
not exist, position the axial surface halfway between the surface data points. Dips exceeding 30
deg exist below and to the west of the axial surface. The kink method predicts that the well should
encounter bed dips in excess of 30 deg at depths exceeding 1000 ft (Fig. 10-25a).
Data from the well, shown in Fig. 10-25a, confirm the accuracy of the kink method. At a
depth of 1200 ft, tadpole dips on the dipmeter log range between 25 deg and 40 deg. The well
encountered a sand horizon below 3000 ft that requires more than a 500-ft hydrocarbon column
for a discovery in this well. Notice that if a well were positioned on the back limb of the struc-
ture, near sp 700, then a smaller hydrocarbon column in that sandstone nevertheless would result
in a discovery. Thus, wells spudded at the back of structures are more likely to encounter hydro-
carbons than wells spudded on the front of structures. Old-timers learned this rule after drilling
numerous wells. Furthermore, wells drilled into the back-limb axial surface are not only more
Cross Section Construction 543
(a)
(b)
(c)
Figure 10-24 (a) - (c) Cross sections of three structures constrained by surface dip and seis-
mic data of varying qualities. Using the kink method, can you predict which one of the three
wells discovered hydrocarbons and which wells encountered steeply dipping beds? (From
Bischke 1994. Published by permission of the Houston Geological Society.)
544 Chap. 10 / Compressional Structures: Balancing and Interpretation
(a)
(b)
(c)
Figure 10-25 (a) The well crosses an axial surface at approximately the 1000-ft level to intersect
25-deg to 40-deg dipping beds. (b) The well is drilled on the crest of an asymmetric fold and
encounters near-vertical beds. (c) The well is drilled into the back limb of a tightly folded structure
and encounters hydrocarbons. (From R. Bischke 1994. Published by permission of the Houston
Geological Society.)
Depth to Detachment Calculations 545
likely to encounter hydrocarbons, but are commonly the most productive wells. We will return to
this empirical observation and suggest a cause for the increased production in this chapter’s sec-
tion on the kinematics of fault bend folds.
Next, examine Fig. 10-24b, which contains a well spudded into an asymmetric fold at the
crest of structure. This fold, constrained by surface bed dips, exhibits a near-vertical front limb.
Seismic data do not image these steeps bed dips. Applying the kink method to the surface bed
dips results in the interpretation shown in Fig. 10-25b. The kink method predicts that the well
should encounter near-vertical bed dips below 1000 ft. Tadpole dips obtained from dipmeter data
below 1000 ft confirm the accuracy of the kink method prediction. The kink method solution sug-
gests that a well spudded between sp 250 and sp 300 would encounter the sand horizon at a depth
of 1200 ft.
Last, examine Fig. 10-24c, constrained by surface bed dips and poor quality seismic data.
The interpreters who drilled this well used the kink method to constrain the interpretation. The
well was spudded off the crest of structure on the more gently dipping back limb. Dipmeter data
confirms the interpreted dip. This well resulted in a hydrocarbon discovery and appears to inter-
sect the subsurface crest of the reservoir horizon (Fig. 10-25c).
We have used the recognition of axial surfaces on seismic data and applied them to fold
interpretation. When interpreting seismic data, you must always be aware of the probable exis-
tence of axial surfaces. Seismic interpreters commonly mistake axial surfaces for faults, due to
the abrupt changes in dip. You can avoid that mistake if you understand the geometry of the pos-
sible structures in the area and follow some of the common-sense methodology discussed.
In conclusion, the kink method is relatively easy to use and makes accurate predictions when
applied to depth-corrected seismic data and outcrop bed dip data. If the kink projection method
does not cross a large fault, then the method typically generates accurate cross sections of sub-
surface geometry. Bed dips can change across large faults, causing the solution to deteriorate.
Remember that surface bed dip data are some of the cheapest data available to interpreters. When
employing 2D seismic data, collect the bed dip data along and adjacent to the seismic survey
lines. On well-constrained structures, the method typically generates accurate results (Suppe and
Medwedeff 1990).
Structures typically contain steeper dipping frontal limbs relative to gentler dipping back
limbs. Wells positioned on the back limb of symmetric and asymmetric structures increase the
odds of encountering hydrocarbons. On the other hand, wells spudded near the steeply dipping
frontal limbs of structures often encounter steep bed dips. If the well penetrates the overturned
limb of an asymmetric fold, then the beds will become younger as the well deepens, potentially
missing the prospective horizons.
Tearpock et al. (1994) discuss additional pitfalls concerning fault propagation folds and
other complex structure styles. A strong structural geologic background is key to exploring in
these areas. The understanding of compressional structural styles, including the types of faults
and folds and their inseparable relationship, is paramount when exploring in fold-thrust belts
(Bischke 1994).
marker or reference bed, the present pin length (l), and the average amount that the marker bed
has been uplifted ( U ) above the undeformed level of the bed, as shown in Fig. 10-26. The aver-
age uplift ( U ) is calculated using the same methods engineers use to calculate reserves (Fig.
10-26). The amount of shortening (S) that the unit has experienced is defined as
S = lo – l
The average amount of uplift times the present length ( U) equals the average area of uplift,
which is then equated to the amount of material that enters the structure from the sides (S x d),
where d is the depth to detachment (Fig. 10-27).
Figure 10-26 The average amount that a marker bed has been uplifted can be determined by
measuring equally spaced line segments that are drawn between a base level and the marker
bed, and then averaging the line lengths.
d = l x U/S
Alternatively, if the depth to detachment is known, then the method can be used to check fold
shape.
A closely related method employed by Laubscher (1961) and described by Goguel (1962)
has been used in the petroleum industry. This method also assumes that no material is entering
the structure from below, as in a duplex (see the section on duplex structures), and that all the
material in the core of the structure is derived from the sides of the structure. Mitra and Namson
(1989) point out that these assumptions are invalid if there is interbed shear (i.e., distortion of the
vertical pin line) or if material is transferred out of the area of the cross section, as occurs in fault
bend folds.
If the material enters the structure from the sides, then the area within the core of a structure
(Au) at a given reference level is measured, as are the final pin length (l) and the initial length (lo)
Depth to Detachment Calculations 547
of a reference or marker bed (Fig. 10-27). As before, the shortening at the reference level is
S = lo – l
The area (Au) within the core of the structure beneath the reference bed is assumed to be
equal to an equivalent volume that comes in from the side (As) (Fig. 10-27). The area can be
obtained by planimetry.
Figure 10-27 Depth to detachment calculation. The amount of material entering the cross section
from the sides is equal to the material that has been uplifted above base level. (Modified after
Laubscher 1961; Suppe 1985. Published by permission of the Swiss Geological Society.)
Therefore,
As = (S)(d)
Au = As
it follows that
Au = (lo – l) d
and
d = Au /(lo – l)
548 Chap. 10 / Compressional Structures: Balancing and Interpretation
NONCLASSICAL METHODS
Introduction
Newer methods of structural interpretation are more precise and more robust than the classical
balancing techniques, which have distinct limitations. For example, Dahlstrom (1969) empha-
sized that within a given area, only a limited number of geologic structures are likely to exist. He
also realized that these structures must area-balance and line-length balance, but exactly how
does the interpreter accomplish these tasks? One obvious method is to measure formation bed
lengths to check for balance, but this can only be accomplished after the interpretation is finished.
In addition, two geoscientists given the same data set are very likely to place lines of equal length
at different positions within a cross section, although both products may be line-length balanced
sections! How are we to evaluate which of the two sections is “correct,” and how could the inter-
pretations be improved? One problem is that line-length balancing has no rules associated with
the method, other than that the bed lengths must be consistent and that the structural styles are
limited.
This problem becomes particularly acute when the data in an area are underconstrained, as
is often the case, and leads to what John Suppe has referred to as the “blank paper” problem
(Woodward et al. 1985). For example, you are studying an area in which the only data available
are at shallow depths, and these data strongly suggest that the structures continue with depth.
Classical balancing lacks constraints, so any attempt to continue the interpretation to depth is
likely to result in as many interpretations as there are interpreters. Furthermore, line-length bal-
ance can be conducted only after an interpretation is finished. Do methods exist to more directly
balance a section, either by hand or with the use of a workstation? More direct methods would
certainly be friendlier to the working environment.
Those of you who have worked with various tectonic settings know that several relationships
are recurrent from area to area. In extensional terranes, the faults are commonly listric and
rollover structures are present, which typically contain antithetic and synthetic minor faults and
perhaps a keystone structure. This suggests that some fundamental process controls the develop-
ment and formation of normal faults and associated structures. In the compressional regime, folds
are either symmetric, as described by Gwinn (1964) in his work on the Appalachians, or they are
asymmetric, such as many of the folds in the Rocky Mountains (Link 1949). Geologists have
noticed that where folds are present, faults also seem to exist in association with the folding
(Bally et al. 1966; Jones 1971; Woodward et al. 1985). Many different regions around the world
possess thrust belts that contain within them symmetric and asymmetric folds, so fundamental
processes seem to control the formation of orogenic belts. If we could develop realistic models
of these fold-and-thrust belt folds, then the petroleum industry would have powerful tools in
which to aid interpretation.
Hence, we enter the world of nonclassical methods that utilize mathematical formulas,
graphs, and models. Perhaps a word of caution is required at this time for the new geology stu-
dent. Although models can be very powerful tools (e.g., the plate tectonic model), the improper
application of a correct model to the wrong situation will only result in error. To make matters
worse, model balancing is nonunique. Different geoscientists, applying the same model to a given
structure, are likely to generate similar results, as we shall see. However, the skeptic may point
out that this is merely an artifact of being schooled in the same interpretation techniques.
Before we enter the exciting world of kinematic processes, we restate that this book is
designed primarily to present subsurface mapping techniques and is not a complete reference on
interpretation per se. In the balancing sections of this book, the mapping techniques are commonly
Nonclassical Methods 549
difficult to separate from the interpretation because you must choose which technique to apply to
a given structure, and this choice involves interpretation. Let us caution you that other interpre-
tation techniques exist that do not involve any particular mapping technique, such as growth sed-
imentary patterns and structures (Medwedeff and Suppe 1986). These growth patterns are often
extremely helpful in determining which model or technique to apply to the structure, so we rec-
ommend that the serious geoscientist gain familiarity with all the approaches to structural inter-
pretation.
Suppe’s Assumptions and Dahlstrom’s Rules
When presented with the problem of a poor or nonexistent data set, several approaches are open
to the geoscientist. Solutions to this problem seem to involve the following.
1. Collect more and/or better data. This subject is left to the data contractors.
2. Make more assumptions in order to solve the structural problem. If data are lacking or are
unobtainable, it is still possible to solve the structural problem, providing you can
extrapolate known data, using known geologic principles, into the area of interest. For
example, if we assume that the kink method is appropriate, we can extrapolate units
within the limb of a fold to depths beyond the control data. In this sense, assumptions can
substitute for data.
We recommend that you employ the following assumptions and rules credited to Suppe
(1988) and Dahlstrom (1970).
Suppe’s Assumptions
(1) Thrust faults step up abruptly from a decollement and (unless deformed) do not have
continuously curved listric shapes.
(2) All thrust faults (that produce a given structural style) in a given area step up at
approximately the same angle.
(3) Layer-parallel slip in a thrust sheet is limited to that caused by changes in dip. This is
another way of stating that the kink method applies at all times.
Dahlstrom’s Rules
(1) Dipping beds over flatter beds define decollements or thrust faults (Fig. 10-5c).
(2) Thrust faults cut up, and not down, stratigraphic section.
(3) Invent more powerful interpretation methods and techniques so that you can extrapolate
existing data into the no-data areas. This is the subject of the remaining sections in this
chapter.
Fault Bend Folds
Our examination of seismic sections from various areas of the world (e.g., Australia and through
the Pacific rim to Alaska, western and eastern United States, western Europe, Argentina,
Venezuela, etc.) indicate that there are two commonly recurring fold styles within the low tem-
perature portions of thrust belts: the symmetric, or fault bend fold type (Figs. 10-28 and 10-29)
(Rich 1934; Suppe 1983) and the asymmetric, or fault propagation fold type (Fig. 10-38) (Link
550
Chap. 10 / Compressional Structures: Balancing and Interpretation
Figure 10-28 Migrated seismic line of fault bend fold from the Taranaki Basin,
New Zealand. A symmetric fold is imaged in the vicinity of the well and sp 100,
between the two-way travel times at 1.5 sec to 1.9 sec. In the vicinity of sp 75
and sp 100, dipping beds overlie flat beds, indicating a decollement. (From
Seismic Atlas of Australian and New Zealand Basins, Skilbeck and Lennox
1984. Published by permission of Earth Resources Foundation, University of
Sydney.)
Nonclassical Methods 551
Figure 10-29 Frontal limb of a fault bend fold in Hudson Valley, New York, USA, located on Route
23 about 300 meters west of New York Thruway. (Compliments of Jon Mosar.)
Figure 10-30 Examples of fault-related fold types. (Published by permission of John Suppe.)
552 Chap. 10 / Compressional Structures: Balancing and Interpretation
1949; Suppe 1985). We stress here that complications in these structures, such as multiple and
back thrusts, often exist and that other thrust-related geometries are present (Fig. 10-30). We wish
to emphasize, however, that these two structural styles are the simplest types of compressional
folds that are commonly present in petroleum basins. Fault bend folds appear to be the most com-
mon of the two structural styles.
Fault bend folds were described by Rich (1934) in the Pine Mountain thrust region of
the Appalachians, where he recognized that this fold style consisted of symmetric anticlines (Fig.
10-31). Rich also recognized that these folds were associated with thrust faults, and he postulat-
ed that the folds were the result of “thin skinned” deformation. Notice that if motion were to
occur along the decollement in Fig. 10-2, hanging wall material would ride up the ramp and onto
the flat. Rich recognized that if this occurs, anticlines and synclines would form (Fig. 10-32).
This example was eventually modeled utilizing a volume conservation concept (Suppe and
Namson 1979; Suppe 1980; and Suppe 1983).
Figure 10-31 Fault bend fold forming over a step-up on a thrust fault. (From Rich 1934;
AAPG©1934, reprinted by permission of the AAPG whose permission is required for further use.)
Figure 10-32 Model of fault bend fold constructed from paper sheets. (From Rich 1934;
AAPG©1934, reprinted by permission of the AAPG whose permission is required for further use.)
The kinematics of the process are as follows. Folds form along nonplanar thrust faults where
a decollement on a lower structural level (Y level, Fig. 10-33a) ramps to a higher stratigraphic
level (X level) (Rich 1934; Bally et al. 1966). Motion along the fault and the conservation of vol-
ume principle cause the beds to ride up the ramp and roll through axial surface BY, forming the
back limb of the anticline. This causes the back dip panel (or flap) BYY' B' to form (Fig. 10-33a).
The two axial surfaces (BY and B' Y' ) terminate at the fault surface, because they are produced
by the bend in the decollement as the beds move up the ramp. Axial surface B' Y' , which is pinned
to the bend in the fault, is actively deforming the hanging wall beds. The bend in the fault caus-
es the deformation. Axial surface, which was initially at the BY position, passively moves up the
ramp as material moves through the bend in the fault. Similarly, the beds moving up the ramp
and onto the flat must roll through axial surface AX, forming the frontal dip panel AXX' A' (Fig.
10-33a). Axial surface AX, which is also pinned to a bend in the fault, is a locus of active defor-
mation and rotation of the hanging wall beds. The beds roll down at axial surface AX and form
the front limb of the anticline. As the beds roll through axial surface AX, they experience bed-
ding plane slip. This slip produces shear in the frontal limb of the fold (Fig. 10-33b) and causes
the frontal limb to dip at a higher angle than the back dip panel. This point is emphasized here
Nonclassical Methods 553
Figure 10-33 Fault bend fold kinematics illustrating the progressive development of beds riding up a
thrust ramp. The beds are deformed by the active axial surfaces. (Modified after Suppe 1983, 1985.
Published by permission of the American Journal of Science.)
554 Chap. 10 / Compressional Structures: Balancing and Interpretation
because it will be applied to the solution of more complicated problems in the section on duplex-
es. As the beds roll through the active axial surfaces BY and AX, a fracture porosity is likely to
form in the deformed beds. Some of the best producing wells drilled on anticlines produce from
rocks close to the active axial surface B′ X′, shown in Fig. 10-33c. Bending of the strata along
this active surface apparently imparts an excellent permeability in some folds.
As fault slip increases and the fold grows, the dip panels extend in width, and point Y′
migrates toward point X (Fig. 10-33b) until the fold attains its maximum amplitude. When this
occurs, axial surface B′Y′ has migrated to the top of the ramp and point Y′ has reached the upper
footwall cutoff (point X in Fig. 10-33b). With additional deformation, the fold now extends by
the lateral motion of axial surface AY′ away from axial surface B′X (Fig. 10-33c). The fold has
reached its maximum amplitude and now is only widening, so no material is currently rolling
through the AY′ axial surface. This surface has become inactive. However, material continues to
roll through the B′ X and BY surfaces, probably further fracturing the rock.
The resulting idealized fold shape, caused by simple step-up of the hanging wall material
along a ramp and onto a flat, has a frontal dip panel that contains a slightly higher dip than the
back dip panel (β is usually slightly greater than θ, as in Fig. 10-33c). Thus, the fold geometry is
roughly symmetric, particularly at cutoff angles (θ) of less than about 20 deg.
The mathematics of this volume-balanced model can be summarized in the form of a graph
(Fig. 10-34). This model vigorously utilizes the kink method, and as this method conserves vol-
ume, line length, and bed thickness, it is not necessary to retrodeform a solution that is derived
from Fig. 10-34. If the data conform to the angles presented in Fig. 10-34, the interpretation will
automatically retrodeform. Thus, the graphical methods presented in this section are useful in
both exploration and exploitation activities.
Let us apply Fig. 10-34 to the case of a fault that steps off a decollement at a 20 deg (initial)
cutoff angle and ramps to an upper flat that parallels the lower decollement (Fig. 10-35a). This
means that φ = θ (Fig. 10-34, inset on left). Also, notice that when φ = θ, θ cannot exceed 30 deg
(see Fig. 10-34). The other assumption that we shall make for purposes of demonstration is that
the amount of slip on the lower decollement is equal to the ramp length. This means that the axial
surface (B′Y′ in Fig. 10-33b) has moved up to the top of the ramp. The initial cutoff angle of 20
deg can now be read off the left part of the abscissa and projected vertically on Fig. 10-34 until
this line intersects the θ = φ line. Next, the dip of the frontal flap (β) can be read off the more
steeply dipping lines on Fig. 10-34, which in this case is about 23 deg (also see Table 10-1). The
axial surface angle (γ) can be read off the ordinate, which in this case is 78.5 deg. The final solu-
tion, shown in Fig. 10-35b, will automatically area-balance and line-length balance, but there is
a final check that should be made.
The amount of slip on the upper flat is less than that on the lower flat. Previously, we stated
that as the beds rolled through axial surface AX (Fig. 10-33a), the deformation was accommo-
dated by bedding plane slip within the frontal dip panel. This is required to conserve both vol-
ume and bed thickness, and it causes angle β to be larger than angle θ. Thus, some of the fault
slip is consumed within the beds of the frontal dip panel, and that causes the amount of slip along
the upper flat to be less than the amount of slip along the lower flat.
The amount of slip to be expected along the upper flat can be determined by using Fig.
10-36. Again, a vertical line is projected from the 20-deg cutoff angle on the left part of the
abscissa vertically upward to the θ = φ line that we have assumed for this example. The ratio of
the slip on the upper flat relative to the slip on the lower flat can now be read from the R lines on
the diagram, which in this case is about 0.87. Therefore, the slip along the upper flat must be 0.87
Nonclassical Methods
Figure 10-34 Fault bend fold graph showing angular relationships between the ini-
tial cutoff angle (θ), the frontal dip panel (β), and the axial surface angle (γ). (From
Suppe 1983. Published by permission of the American Journal of Science.)
555
556 Chap. 10 / Compressional Structures: Balancing and Interpretation
(a)
(b)
Figure 10-35 Fault bend fold exercise for beds ramping up a fault with a 20 deg cutoff
angle and with slip on the lower decollement equal to the ramp length.
of the slip along the lower flat. Field geologists have often observed that the slip on faults dies or
decreases within the cores of folds, and this is one reason why fold belts die toward the foreland.
This exercise for checking slip is most useful when experimenting with structures that exhibit
unusual geometries or complicated shapes. Another check on the solution would be to measure
bed lengths on more than one structural level.
We can now present a major conclusion concerning fold geometry. Perhaps you have noticed
the relationship between fault shape and fold shape. The shape of the fault is related to the shape
Nonclassical Methods
Figure 10-36 Fault bend fold graph showing the amount of slip to be expected
along different portions of a fault surface. The R lines indicate the ratio of slip
along the upper flat relative to slip along the lower flat. (From Suppe 1983.
Published by permission of the American Journal of Science.)
557
558 Chap. 10 / Compressional Structures: Balancing and Interpretation
of the fold (Rich 1934; Dahlstrom 1969) and indeed, fault shape determines fold shape (Fig.
10-32). Thus, if you know fault shape, you can predict fold shape, and conversely, if you know
fold shape, you can infer something about fault shape. For example, on Fig. 10-33b and c, notice
how the thrust ramps up where the fault flat intersects the synclinal axial surface BY. So, the fault
ramp parallels the back limb of the upturned hanging wall beds. Where the thrust intersects the
active axial surfaces AX or B' X, the thrust forms the upper flat.
Fault Propagation Folds
Fault propagation folds are a common fold type observed in outcrop and on seismic data (Figs.
10-37 and 10-38), and like fault bend folds, they are known to be good hydrocarbon producers.
Fault propagation folds possess the particular characteristic that as the fold grows, the deforma-
tion advances at the tip of a propagating thrust fault (Fig. 10-38), hence the name “fault propa-
gation fold” (Suppe 1985). As long as the structure has not been faulted through (i.e., been sub-
ject to breakthrough), the slip is consumed by bedding plane slip within the frontal limb of the
fold (Fig. 10-39).
Fault propagation folds typically have higher cutoff angles than fault bend folds, in the range
of about 20 deg to 40 deg, which causes these fold types to possess steeply dipping to overturned
frontal limbs that commonly do not image on seismic sections, along with a characteristic asym-
metry (Fig. 10-40). This striking asymmetry, where imaged on seismic sections across folds with
less dip, has the appearance of a striking snake, giving rise to the expression snakehead structure
(Fig. 10-37).
The kinematics of fault propagation folds are as follows. A fault, propagating upward from
a decollement, causes beds at the front of the propagating fault tip to bend forward as material
moves up the ramp (Fig. 10-39a). As in fault bend folding, the beds will also bend up the ramp
created by the propagating thrust fault as they move through axial surface B, creating the back
dip panel outlined by axial surfaces B and B′ (Fig. 10-39a).
In this style of folding, an increase in the amount of deformation within the core of the fold
accommodates the slip on the fault. Therefore, the beds near the tip of the thrust fault bend for-
ward, commonly at steep angles (Fig. 10-39a). The rotation of these steeply dipping beds, and
bedding plane slip between the beds, consumes the slip along the thrust fault. Thus, the slip on
the fault dies out within the core of the fold. This type of fault is referred to as a blind thrust. The
more steeply dipping beds between the front and the top of the structure form two axial surfaces,
A and A′, as shown in Fig. 10-39a.
During fault propagation folding, all the axial surfaces are active and, with the exception of
axial surface B, move through the material as the beds deform (Fig. 10-39). As the propagating
fault extends and the fold grows in amplitude, it incorporates more material into the frontal limb
of the structure. Consequently, as the fault propagates forward and as axial surface A′moves
away from axial surface A, point 2 of Fig. 10-39a and b rolls through A′ into the steeply dipping
frontal limb. With increasing deformation, the dip panels, as defined by axial surfaces A and A′
and by B and B′, broaden (Fig. 10-39c). Axial surface B′ is an active surface, as beds roll through
it from the crest of the fold into the back limb. Axial surfaces A and B′ form a branch point at the
same stratigraphic horizon as the fault tip. As the fault propagates, the loci of deformation, and
thus the axial surfaces and the branch point, move forward and upward. The structurally lower
beds fold more tightly and the back limb of the fold widens. As the fold grows, the deformation
also fractures the rock, which can affect the porosity and permeability.
Fault propagation folding can exhibit a variety of structural styles, depending upon the cut-
Nonclassical Methods
Figure 10-37 Seismic section imaging asymmetric fault-propagation folds,
Southern Appalachians, Alabama. (Interpretation by R. Bischke. After Sachnik and
More in Bally (1988); AAPG©1988, reprinted by permission of the APPG whose
permission is required for further use.)
559
560 Chap. 10 / Compressional Structures: Balancing and Interpretation
Figure 10-38 Fault propagation fold, Appalachians, Tennessee. The frontal limb dips more
steeply than the back limb. Thrust fault dies near synclinal axial surface at the front of the fold.
(From Suppe 1985.)
off angle (Fig. 10-40) and the amount of slip. As the cutoff angle increases, and for the same
amount of slip, the folding will appear to be more symmetric on seismic sections even though the
amount of slip remains unchanged. If the fold forms according to the processes described in Fig.
10-39, the cutoff angle can be determined directly from the dip of the beds within the back dip
panel as these beds parallel the ramp.
Given additional amounts of slip, the fault propagation may find a weak or incompetent
horizon that parallels bedding and becomes a hybrid fault bend fold (Fig. 10-41c). Alternatively,
the structure can break through the anticlinal, the synclinal, or the overturned limb portions of the
fold, creating more complex geometries (Fig. 10-41a, b, and d).
As with fault bend folds, fault propagation folds can be balanced using formulas or graphs,
as in Fig. 10-42 (Suppe 1988; Suppe and Medwedeff 1990). In outcrop or on seismic sections,
fault propagation folds can be balanced by observing either the ramp angle (θ) or the back limb
dip of the fold. When using seismic data, remember to depth-correct the seismic or choose sec-
tions that are roughly on a scale of one to one. This can be readily accomplished on the work-
station over a given interval by using checkshot data or velocity information. We have found this
procedure to be adequate for most cases.
We study a simple case for balancing fault propagation folds, using Fig. 10-43. For exam-
ple, you may observe beds on a depth-converted seismic section dipping and overlying horizon-
tal beds. The back limb beds are determined to dip at 30 deg, so θ = 30 deg. The corresponding
axial surface angles γp and γp* can be read off Fig. 10-42 (γp = 53 deg and γ*p = 38 deg). The kink
method can now be employed by using Fig. 10-43. (Note how angles are measured in the figure.)
Nonclassical Methods 561
(a)
(b)
(c)
Figure 10-39 Fault propagation fold kinematics, illustrating the progressive development of
beds deforming at the tip of a propagating thrust fault. (Modified after Suppe 1985.)
First, construct a 30-deg dipping ramp, with the tip of the thrust fault as best determined from
seismic data. Construct the structurally lower γp axial surface, which dips at 53 deg.
The tip of the thrust fault and the front limb dip are used in determining the position of the
branch point defined by the upper γp, B′, and γ*p axial surfaces (Fig. 10-43). The inclination of the
front limb (β) is defined by
β = θ + 2 (γp*)
562 Chap. 10 / Compressional Structures: Balancing and Interpretation
so the frontal limb inclination in this case is 106 deg (Fig. 10-43b). Draft a flat horizon to the left
of the fault tip and then, using β = 106 deg, project the horizon upward from the fault tip. The
position of the branch point is then located by projecting that same horizon, which is above the
fault flat and on the level of the fault tip, across the fold’s back limb. This horizon is bent upward
at the active axial surface at the base of the ramp (surface B, Fig. 10-39). Bisect the angle between
the ramp and the flat to determine the position and inclination of the axial surface (105 deg in
Fig. 10-43b). Then project the horizon parallel to the fault ramp, upward to the intersection with
the front limb dip panel that was projected upward from the fault tip. The projected horizons
intersect at the branch point (Fig. 10-43b). Next, the structurally higher γp axial surface (53 deg
Nonclassical Methods 563
Figure 10-42 Fault propagation fold graph for a simple step-up from a decolle-
ment surface. (Modified from Suppe 1985.)
564 Chap. 10 / Compressional Structures: Balancing and Interpretation
dip) can be drawn from the branch point. The γ*p axial surface (38 deg dip) can be drawn from the
branch point into the core of the fold and to the fault (Fig.10-43b). Then draft the axial surface
at the top of the back limb upward from the branch point, parallel to the axial surface at the base
of the back limb. The elements of the structure are now complete, and additional layers can be
projected throughout the structure. The line lengths then can be measured for area conservation
between pinpoints (Fig. 10-43c).
γp
β
180o 0o
Fault tip
θ = 30o
90o
(a)
γp
53o θ
β
105o
γp*
β = 106o
38o
(b) (c)
Imbricate Structures
As the thrust belt moves progressively over the foreland, there is a tendency for new thrust faults
to form near the toe (front) of the thrust belt and for these thrusts to seek a lower structural level.
Where a thrust fault forms below a pre-existing fault(s), motion along the deeper fault will cause
the shallow fault(s) and its overlying structure(s) to fold. The deformation can produce some
rather interesting and complex geometries of stacked folds (Fig. 10-44). Thrust faults that form
at a higher structural level and above the newly formed imbricate thrust faults have been, perhaps
inappropriately, called out-of-sequence thrusts.
This complex process is best described through example. We shall first assume that a fault
bend fold formed near the front of a thrust belt, as shown in Fig. 10-45a. In this example, we have
assumed that faulting formed ramp AB and that the cutoff angle is 20 deg. We can now determine
the frontal dip panel angles by using the methods developed in the section on fault bend folds.
We also assume for purposes of demonstration that, at a particular time, the fault breaks through
at the lower level and another ramp forms in front of the ramp that formed the fault bend fold.
The new ramp along the lower decollement is ramp CD in Fig. 10-45b.
Nonclassical Methods 565
(a) (b)
Figure 10-44 Diagram showing different types of duplexes. (a) Stacked fault propagation
folds (b) Foreland and anticlinal stacked duplexes. (From Mitra 1986; AAPG©1986, reprint-
ed by permission of the AAPG whose permission is required for further use.)
angle of ramp AB is 20 deg, the upper segment of ramp AB can be projected downward at a 20-
deg angle in Fig. 10-45d (i.e., line segment FB). Similarly, the lower portion of ramp AB (locat-
ed near point A of Fig. 10-45b) slid along the lower flat without being deformed, and thus the
lower part of the ramp AB can be projected upward at a 20-deg angle (Fig. 10-45c). The central
part of the horse, however, has been subject to deformation as the wedge-shaped horse moved up
ramp CD and through axial surface EC. Axial surface EC is pinned to the lower footwall cutoff
at point C and is an active axial surface. Ramp CD has a 20-deg initial cutoff angle, so the bisect-
ing axial surface EC dips at 80 deg (Fig. 10-45d). It also defines the extent of line AE.
(a)
(b)
78.5o
23o
D'
?
D
A
C
(c)
B G
F D'
E 38o D
A
C
(d)
+I 52o
-I
-II +II
0
0
-I +I
(e)
Figure 10-45 (continued)
568 Chap. 10 / Compressional Structures: Balancing and Interpretation
We have now determined that as fault ramp AB moves through axial surface EC, it must
deform (bend upward). Ramp AB initially dipped at 20 deg before deformation, so it must dip at
an even higher angle after deformation. If a 20-deg dipping line (line AB) moved up a 20-deg
dipping ramp (line CD of Fig. 10-45d), one might incorrectly conclude that the central deformed
portion of fault ramp AB (line segment EF) would dip at 40 deg. We use Table 10-1 to provide
the correct dip angle for line EF.
The following is an example of the rationale for Table 10-1. In the section on fault bend fold-
ing, we learned that in order to maintain line lengths, the angle β, the dip of the frontal limb, must
be greater than the (initial) cutoff angle θ. This relationship must be maintained on every struc-
tural level within imbricate structures. Thus, as the frontal part of the horse rides over the top of
the ramp, it rotates forward to a dip angle of 23 deg (Fig. 10-34, or Table 10-1). This rotation
causes the overlying beds to dip forward at a higher angle (dip panel +II in Fig. 10-45e). The dips
at the higher structural level experience a quantum increase in dip (Suppe 1980, 1983). In other
words, the insertion of the horse onto the upper flat will cause the beds above it (panel +II) to dip
at an angle that is greater than twice 23 deg, or in this case 52 deg, as determined from Table
10-1. In Table 10-1, the central column contains values for the cutoff angle θ and the other
columns provide the calculated dips within each panel of the fold. Thus, for a 20-deg cutoff
angle, the front limb dips at 23 deg and at 52 deg, in panels +I and +II respectively. In order to
maintain line length and formation thickness, the strata above the horse will shear in such a man-
ner that an increase in dip in the frontal panel is accommodated by a decrease in the expected dip
in the back panel. Thus, the compensating dip of line segment EF (panel -II) in Fig. 10-45d and
e is 38 deg, rather than 40 deg (Table 10-1). The dip panels III through VII in Table 10-1 relate
to higher order duplex structures.
The length of line AE was defined previously by constructing the axial surface at point C in
Fig. 10-45d. Point F is determined from the amount of slip on the lower fault. Measure that slip
up the ramp and draw the inactive axial surface parallel to the one at point C; then draw line EF
TABLE 10-1
Nonclassical Methods 569
at a dip of 38 deg (Table 10-1) upward to the axial surface. Complete the deformed fault surface
by drawing line FB (point B was determined above). The inactive axial surface beneath F will be
the line at which you project horizons within the horse downward and parallel to the lower ramp,
as in Fig. 10-45e.
The strata above deformed ramp AB will parallel that surface and bend at axial surfaces
located at points E and F. Bisect the angles along the deformed upper ramp (angles AEF and EFB
in Fig. 10-45d) and complete the dip domains toward the surface (Fig 10-45e). The crest and the
front limb (with a 23 deg dip) of the original fold were above the upper fault flat, to the right of
point B in Fig. 10-45a. As described previously, part of this frontal dip panel of the original fault
bend fold was subsequently deformed by the frontal portions of the horse, and it dips at 52 deg
(Fig. 10-45e). However, the part of the horse between points F and G (Fig. 10-45d) is unde-
formed, so the dips of the overlying strata and the axial surfaces at B are maintained from the
original fold (dip panels –I, 0 and +I in Fig. 10-45e). A small dip panel (+I) exists to the right of
panel +II and has a dip equal to that of the original front panel (23 deg), as it is a part of the frontal
limb of the original fold that was not deformed (Fig. 10-45b). The axial surface on the right of
small dip panel +I has an inclination of 78.5 deg, maintained from the original fold. The axial
surfaces bounding panel +II are placed at bends in the upper fault surface and drafted at inclina-
tions that bisect the adjacent dips. Complete the cross section by using dip data in Table 10-1 to
draw the horizons in panels –I, +I, and 0 (flat). The finished, balanced cross section in Fig. 10-45e
depicts a duplex structure.
Although this exercise may at first seem to be an unnecessary complication, we shall use
these small changes of dip to our advantage in what is called dip spectral analysis (Suppe 1980,
1983). Dip spectral analysis can be used to interpret poorly imaged subthrust plays.
Consider the geometry present in Fig. 10-45e, which has several implications concerning
petroleum exploration. Assume that the thin horizon above the lower flat is a productive reservoir
horizon. Notice that this reservoir can be intersected on two structural levels, resulting in two
potential plays. The first play is the closure associated with the original fault bend fold. The sec-
ond play is a partial closure located within the horse, and thus its prospectivity would depend on
the trapping mechanism or the permeability of the beds above the thin reservoir horizon.
Dip Spectral Analysis. This section discusses the method to locate potential subthrust plays in
practice. We learned from the previous imbricate structure exercise that for uniform step-up
angles, and as one moves up the structural pile, the dips at the front of the imbricate structures
increase at an increasing rate. Furthermore, the back dips, whereas exhibiting a corresponding
increase, do so at a decreasing rate. Therefore, the frontal dips exhibit a unique quantum increase
in dip (in our case 52 deg is greater than twice 23 deg), whereas the back dips exhibit a unique
quantum rate decrease in dip (Table 10-1). These unique changes in dip allow us to estimate the
number of subthrusts and their approximate position. For example, in Fig. 10-44b the structure
in the lower figure has four different back dips relative to the flat regional dip. This means that
the duplex has four or more thrusts and four or more potential repeated sections.
Notice that in Fig. 10-45e the final structure exhibits three frontal dip panels or domains
(labeled +I, +II, and +I) and three back dip panels (labeled –I, –II, and –I) that are separated by
a region of initial dip (panel 0). In nature, detailed surface mapping across the structure shown in
Fig. 10-45e could result in the topographic section shown in Fig. 10-46a. In this figure, the fol-
lowing dips occur from left to right: 0, –38, –20, 0, 23, 52, 0, –20, –20, –20, and 0 deg (desig-
nating leftward dips as negative). If a regional dip of 5 deg to the left were to exist, then the cor-
responding dips would be –5, –43, –25, –5, 18, 47, –5, –25, –25, –25, and –5 deg. Regional dip
570 Chap. 10 / Compressional Structures: Balancing and Interpretation
should be removed prior to dip spectral analysis. All angles are determined relative to regional
dip, but by removing regional dip we can conveniently use angles measured from the horizontal,
as given in Table 10-1. The resultant dips corrected for 5 deg regional dip are 0, –38, –20, 0, 23,
52, 0, –20, –20, –20, and 0 deg.
We have shown that higher dips exist at the front of the structure, whereas lower dips occur
at the back (Fig. 10-45e). Thus, the 52-deg dip and its associated 23-deg dip are forward dips,
(a)
(b)
(c)
Figure 10-46 (a) - (d) Duplex exercise, inverse model.
Nonclassical Methods 571
(d)
whereas the –38-deg dip and its associated –20-deg dip are backward dips (see Table 10-1). As
there are two forward dips, the 52-deg dip represents a +II domain, and the 23-deg dip represents
a +I dip domain. These numbers, each of which represents an individual dip domain, can be com-
pared to Table 10-1 (line 13) to indicate that there are two thrust faults ramping at 20 deg, caus-
ing second-order frontal and back dips of 52 deg and –38 deg respectively, and first-order frontal
and back dips of 23 deg and –20 deg respectively, in the upper structure. If three faults are pres-
ent, then we would expect an additional forward dip of about 98 deg and a back dip of about 53
deg (Table 10-1). As only two forward dips exist in our example, Table 10-1 suggests that only
two thrust faults are present. Therefore, surface and subsurface data can be compared to Table
10-1 in order to determine the number of imbrications that may exist in the area of study. The
data could be surface bed dips, subsurface dipmeters, and/or depth-corrected seismic data.
Let us proceed to solve the structure presented in Fig. 10-46a, which is based on Fig 10-45e,
by using only outcrop data. After analysis of the surface dip data, the first task is to determine the
dips of the axial surfaces, which separate the observed dip domains. Once the related dips with-
in each of the dip domains have been averaged, as was described in the section on the kink
method, the dip of the axial surface between given adjacent dip domains can be determined from
the following formula:
For this equation only, structural dip is taken to be negative if it is to the right. It would fall
within the 90-deg to 180-deg quadrant, as defined in Fig. 10-46a. If dip is to the left, it falls with-
in the 0-deg to 90-deg quadrant and is assigned a positive value in the equation. The dips of the
axial surfaces for the appropriate dip domains presented in Figs. 10-46a and 10-45e were calcu-
lated and are presented in Table 10-2. For example, at the top of the hill on the left, a 0-deg dip
exists adjacent to a 23-deg dip to the right (Fig. 10-46a). These two dips represent a 0 dip domain
572 Chap. 10 / Compressional Structures: Balancing and Interpretation
and a +I dip domain (Fig. 10-45e and Table 10-1). They are separated by a dip domain boundary,
which is the 0 +I axial surface. Thus, the 0 +I dip domain boundary dips at (0 – 23 + 180)/2 =
78.5 deg (Table 10-2).
TABLE 10-2
These axial surface dip calculations are best applied to the solution of problems using a
method that was suggested by John Suppe and is shown in the lower portion of Fig. 10-46a. First,
calculate the dips of all axial surfaces for the adjacent pairs of dip domains. These are shown in
Table 10-2. Then create a reference set of axial surfaces, as shown in Fig. 10-46a. The dip of each
axial surface is projected downward from a central point and labeled. The dip of each axial sur-
face is measured off in a counterclockwise direction. Two triangles can now be aligned with the
axial surface dip data and slid into any position on the cross section that the interpreter desires.
This procedure will make the interpretation process more rapid.
Possessing the dip data set presented along the topographic profile in Fig. 10-46a and the
knowledge of the solution to the problem presented in Fig. 10-45e will not allow us to arrive at
a unique solution to our problem. Trial and error and some guessing will be required to solve the
problem presented in Fig. 10-46a. The advantage of solving this problem is obvious for areas
where seismic data fails to image imbricate structures. It may result in the identification of duplex
structures and generation of additional prospects and perhaps additional oil and gas discoveries.
Before proceeding further, notice that Fig. 10-46a does not include the –20-deg dip panel
above the lower flat, which is present in Fig. 10-45e and which is to be expected from Table
10-2. This should make our problem more interesting and realistic. The 0 –II data are included in
Table 10-2 for reasons which will soon become apparent.
The first step in the solution of our problem is to examine the data from a geometric point
of view. Two observations are critical to an accurate interpretation. First, the observed dips tend
to follow the topographic slope, which is often the case in nature. Second, the beds above the
thrust fault at the outcrop dip at about the same angle as the thrust fault, suggesting that a ramp
is responsible for the 20-deg tilt to these beds. Therefore, the thrust fault can be projected from
the outcrop downward to where it intersects the adjacent (0 deg) dip domain. But where is this
point? It exists where the 0 –I axial surface intersects the projected thrust fault. The change in
topography is used to position and project the 0 –I axial surface with a 100-deg dip (from Table
10-2) downward to the point where it intersects the thrust fault, thus determining the base of the
upper ramp (Fig. 10-46b, point A). In addition, we have predicted the structural level of the upper
flat or decollement (Fig. 10-46b).
Looking at the dip data further, we see the 0-deg dip at the hilltop on the left and the 23-deg
and 52-deg dips to the right. Using Table 10-1, we can infer that two 20-deg ramping thrusts cre-
ate the observed 23-deg and 52-deg forward dips, and thus these two thrusts must be imbricated
Nonclassical Methods 573
(i.e., stacked) in order to produce the observed quantum increase in dip. The data suggest that a
52-deg dip domain adjoins a 0-deg dip domain and, therefore, a 64-deg axial surface (between
the 0 and +II domains of Table 10-2) is positioned at the appropriate change in topographic slope
(Fig. 10-46b). We know that two things exist to the left of the intersection of the 0 +II axial sur-
face with the upper flat at point B. First, the structurally higher beds dip at 52 deg to the right,
and second, a 23-deg deformed thrust must exist beneath the 52-deg dipping beds. This follows
from a direct application of Table 10-1 and the theory that we presented earlier.
Consequently, at point B we interpret and construct a 52-deg dipping bed and the deformed
23-deg dipping fault (Fig. 10-46c). The 52-deg dipping bed and the deformed 23-deg fault are
then projected up to the +I +II axial surface, which has a dip of 53 deg (Table 10-2) and is posi-
tioned at the break in topographic slope. To the left of this axial surface, the fault will be flat as
that in an undeformed part of the older fault. Now the kink method is applied to the deformed
horse block in order to map the forward dip within it, beneath and parallel to the deformed fault.
An axial surface must extend downward from the kink in the fault. Given the adjacent dips of 0
deg and –23 deg, the dip of the axial surface calculates to be 78.5 deg (Table 10-2). A (0 +I) axial
surface is drawn downward from the kink in the deformed fault to where it intersects the flat, thus
locating point C in Fig. 10-46c. The frontal portions of the deformed horse block have now been
properly defined (i.e., we have separated the undeformed and the deformed regions of the horse
block). Point C not only marks the position where the horse block has ridden onto the upper flat,
but it also determines where the structurally lower ramp can be projected downward at a 20-deg
angle (Line EC, Fig. 10-46c).
Proceeding with the construction, the 52-deg dipping beds in panel +II above the horse can
be projected from the +I +II axial surface at a 23-deg angle up to the 0 +I axial surface (point F
in Fig. 10-46c). The axial surface was positioned from topographic data. The flat of the upper
thrust is projected to the 0 +I axial surface to establish point G, which defines the edge of the
horse block, or the top of the ramp on the original fault (Fig. 10-45a, point B). From point G, the
upper fault can be projected downward at a 20-deg angle (defined by the dip data in panel –I) to
the –I –II axial surface (point H, Fig. 10-46c). The axial surface was positioned at a minor break
in the topographic slope. The structural and fault dip must change at this axial surface, so the
deformed fault is projected downward at a 38-deg angle, as defined by the dip data in panel –II,
until the upper fault intersects the 0 –II axial surface (point I, Fig. 10-32c). This axial surface,
with an inclination of 109 deg (Table 10-2), was located at a break in the topographic slope.
Using the 0 –II axial surface as a bisector, project the upper fault to level out on the I-J structur-
al level parallel to the overlying flat structural dip. The lower flat on the structurally lower fault
still must be interpreted. An axial surface must exist between the –20-deg ramp and the flat. It
calculates to have a 100-deg dip and it must extend downward from point I. Thus, point E is locat-
ed as the base of the ramp. The fault flat is then drawn to the left of point E.
Finally, the axial surface between points H and D is constructed parallel to active axial sur-
face IE (Fig. 10-46c). It is the inactive axial surface within the horse. The length of the line ED
should be the amount of slip on the younger fault, according to this interpretation.
What is the final result of our interpretation process? We have followed the method with a
result that is somewhat complex and confusing! Notice that the upper and lower faults do not
merge on the same structural level (compare I to E, Fig. 10-46c). How can we improve the solu-
tion, and what features should we look for when attempting to arrive at a more satisfactory solu-
tion? First, compare the slip on the upper flat (CB) of the younger fault to the slip up the ramp
(ED). They are incompatible. Line CB should be 0.87 of ED, using Fig. 10-36 to determine the
slip correction factor for a step-up angle of 20 deg. However, it is less than 0.87. Thus, an error
574 Chap. 10 / Compressional Structures: Balancing and Interpretation
was made in this portion of our analysis that represents not only a clue to the proper solution, but
also the approximate position of our difficulties. Second, returning to Table 10-1, we reexamine
the possible dip domains that are associated with two 20-deg ramping thrusts. As there are obvi-
ous problems at the back of the structure, so perhaps the problem lies in this region. The data in
Table 10-1 suggest that a –I dip domain can exist between a 0 dip domain and a –II back dip
domain of 38 deg. Labeling the dip panels on our solution reveals that a –I deg dip domain was
not included in the back area part of our solution.
We now proceed to backtrack, modifying the first solution by inserting a –I back dip domain
between the –II and the back dip panels. Using Fig. 10-36, we know that the distance ED in Fig.
10-46d should be equal to CB/0.87, and this distance is entered on the figure by measuring down
the ramp from point D. From the calculated position of point E, we draw the fault flat. Then, from
point E, we project a 0 –I axial surface upward at 100 deg (Table 10-2) to the deformed upper
thrust, which dips at 38 deg from point H. From this point K, we project a –I –II axial surface
upward at 119 deg (Table 10-2). We can also conclude that to the left of point K exists a 20-deg
–I back dip panel and a 0 –I axial surface that intersects the decollement at point L, which is
established by projecting the upper fault at a 20-deg dip to the lower fault flat. This solution,
although only slightly different from Fig. 10-46c, creates a deformed horse block and is a more
reasonable interpretation.
From this complex exercise, we conclude that (1) duplexes produce more rounded struc-
tures; (2) structural balancing can be nonunique, even under ideal situations; and (3) the inter-
pretation process can be rigorous, but in this age of global energy shortfalls the rewards could be
substantial. The better we understand the detailed geometry of structures, the more likely we are
to find additional reserves of oil and gas.
Box and Lift-Off Structures
Box and lift-off structures represent a particular class of folds, which when viewed relative to
the regional dip are roughly symmetric but angular structures that contain steeply dipping limbs
(Figs. 10-47 and 10-48). Both structural types form along a zone of weak detachment located at
depth and possess the characteristic that the decollement is isoclinally folded into the hanging
wall (Laubscher 1961; Namson 1981). In the Jura Mountains, this zone of weakness consists of
evaporites (i.e., gypsum), although over-pressured shales are likely to produce a similar defor-
mational style. Box and lift-off structures differ from diapiric structures in that there is less mass
transport or flow into the cores of these folds. This causes the box and lift-off structures to have
almost vertically-dipping limbs at the lower structural level. In addition, diapiric structures typi-
cally result from a gravity instability, whereas box and lift-off folds result from compression.
The box fold structural style was once thought to be a relatively rare structural style. Today
the structures have been observed in many compressional environments and can be productive of
hydrocarbons. Box folds can be distinguished from other structural styles in that the width of the
box fold, across the crest of the fold, maps as a region of constant width (Fig. 10-47a). No other
fold style exhibits this geometry in map view.
Box and lift-off structures can be recognized in outcrop from their bilateral symmetry and
also from their angular geometry. If broad zones of vertically dipping beds are encountered in
outcrop (e.g., 70 deg to 80 deg), then these structural styles are suspect (Fig. 10-48).
Box folds have nearly flat tops, vertically dipping limbs, and axial surfaces that dip at about
45 deg (Fig. 10-47a). If in a region of vertically dipping beds two axial surfaces intersect at near-
ly right angles, then you should consider the possibility that box folds are present. On seismic
sections, vertically dipping beds do not image, and thus a pattern of gently dipping reflectors
Nonclassical Methods 575
separated by two zones of noncoherent reflectors, representing the almost vertically dipping beds,
may be an indication of this style of deformation. However, zones of noncoherent reflectors on
seismic sections can result from other causes, such as strike-slip faulting, rock type, or data acqui-
sition problems. Seismic reflection analysis (Payton 1977; Sheriff 1980) could resolve the corre-
lation problem because the sedimentary sequences on the flanks of box and lift-off folds are ele-
vated within the cores of these structures.
Lift-off folds differ from box folds in that the shallow limbs of the lift-off fold style tend to
dip in the 45 deg to 60 deg range relative to the regional dip. At depth, the shallow, steeply dip-
ping limbs merge into a zone of nearly vertical-dipping limbs (Fig. 10-47). If you observe near-
ly vertical beds just above a decollement, it may be impossible to determine if you are observing
a lift-off or box fold, as both structural types possess 45-deg dipping axial surfaces on this struc-
tural level. In practice, however, this difference may be academic.
If the lift-off fold is not subject to bedding plane shear, then the limbs of the fold at a high-
er structural level dip at about 53 deg (Fig. 10-47b). This dip angle changes with increasing bed-
ding plane shear (Namson 1981), and the amount of shear can be calculated from the dip of the
fold limbs (Mitra and Namson 1989). If there is bedding plane shear within the structure, then
Mitra and Namson (1989) show that this shear affects the depth to detachment (as presented in
the section Depth to Detachment Calculations in this chapter). However, the difference is not
major for small amounts of shear. Mitra and Namson (1989) should be consulted for more accu-
rate depth-to-detachment calculations.
Box and lift-off folds are commonly found in association with each other. In the Pre-Alps,
Mosar and Suppe (1988) observed that lift-off structures form in the leading or the trailing posi-
tion relative to fault propagation folds. In addition, they observed that fault propagation folds may
transform laterally into lift-off structures, and that the two structural styles may be related to each
other as the local cutoff angle steepens. Thus, at low cutoff angles (less than about 18 deg to 20 deg),
(a) (b)
(c) (d)
Figure 10-47 (a) Box and (b - d) lift-off structures. (From Namson 1981.
Published by permission of the Chinese Petroleum Institute.)
576
Chap. 10 / Compressional Structures: Balancing and Interpretation
Figure 10-48 Cross section of the Chuhuangkeng Anticline, Taiwan, showing a broad
region of near-vertical dips. (From Namson 1981. Published by permission of the
Chinese Petroleum Institute.)
Nonclassical Methods 577
fault bend folds may form in an area, whereas if the cutoff angle is greater than about 20 deg to
25 deg, fault propagation folds usually form instead of fault bend folds. If the cutoff angle
increases to over 60 deg along the strike of a structure, then the structure may transform into a
lift-off or box fold.
When mapping box or lift-off structures, apply the kink method in your mapping. When
applying this method to these symmetric structures, remember that the hanging wall decollement
is assumed to rise vertically above the basal detachment and to fold back upon itself (Figs. 10-47
and 10-48). Box folds can be distinguished from other structural styles in that the width of the
box fold, across the crest of the fold, maps as a region of constant width (Fig. 10-47). No other
fold style exhibits this geometry in map view.
Triangle Zones and Wedge Structures
Triangle zones and wedge structures are complex structures that exhibit both a lower and an
upper detachment. The basal detachment is often called the sole, or floor, thrust, whereas the
uppermost thrust is called the roof thrust (Fig. 10-49) (Boyer and Elliott 1982). In the case of tri-
angle zones, the roof thrust is a passive back thrust. The wedge moves above the sole thrust and
beneath the roof thrust, peeling off the shallow portions of the cover.
Gordy and Frye (1975) and Gordy et al. (1977) initially used the concept of a triangle zone
to explain the complex relationships associated with an anticlinorium located at the front of the
Canadian Rockies. Jones (1982) refined the concept and showed that the structure contained a
duplex and that it was responsible for the termination of the eastern-directed thrusting along the
Rocky Mountain thrust front. We have learned that during the orogenic process, the deformation
progresses (advances) toward the foreland. Therefore, a frontal portion that existed at a previous
time during the formation of the thrust belt would exist today hinterland of the thrust front. This
implies that fossil triangle zones can exist within the cores of mountain ranges, perhaps repre-
senting the frontal edge of the deformation at a previous time.
A simple triangle zone that uses the concept of a ramping monocline is illustrated in Fig.
10-49. Notice that the deformation terminates where the roof thrust meets the sole thrust, creating
a half-syncline. This monoclinally shaped syncline with only one limb lies foreland of the thrust
belt. Jones (1982) mapped a half-syncline along the Rocky Mountain front and concluded that a
wedge-shaped body of material must be thrust underneath the dipping beds of the half-syncline.
A seismic section of a complex triangle zone is imaged in Fig. 10-50. Notice the wedge-
shaped body represented by the duplication of the reflection located between sp 190 to sp 240
and at 1.2 sec to 1.5 sec. That reflection appears to correlate to the flat reflection at 1.5 sec in
front of the structure. On the seismic line, use the following procedures to locate the backthrust.
First, project the synclinal axial surface at the front of the triangle zone downward to where the
axial surface intersects the sole thrust, as in Fig. 10-49. At the point where the two surfaces inter-
sect, construct a line that is parallel to the monoclinally dipping beds. Project this line that rep-
resents the backthrust toward the surface. The backthrust conforms to the shape of the hanging
wall beds. Notice at sp 220 and at 1000 ms how the backthrust separates dipping beds in its hang-
ing wall from flatter beds in the footwall. This change in bed dips across the backthrust is indica-
tive of a decollement or faulting.
Medwedeff (1988, 1989) extends the concept of interactive sole and roof thrusts to single
structures, and he calls these interactive thrusts wedge structures. Figure 10-51a is an example
of an incipient wedge structure that has two bends on its sole thrust and a single bend in its roof
thrust. As the deformation progresses (Fig. 10-51b), motion along the sole thrust deforms the roof
thrust. Back and frontal dip panels form over what is essentially a fault bend fold that also has an
578 Chap. 10 / Compressional Structures: Balancing and Interpretation
TOPOGRAPHY
ROOF THRUST
SOLE THURST
upper roof detachment. Notice, however, that the overlying beds in effect ride up the roof thrust,
and they will also form fold panels above the upper detachment. This structurally higher fold is
caused by the bends in the roof thrust, so its dip panels terminate at the upper detachment (Fig.
10-51b). The result is two folds for the price of one, which are slightly offset from each other. As
the deformation progresses (Fig. 10-51c), the axial surfaces interfere and annihilate each other as
they form branch points. This example illustrates that the deformation process can be very tran-
sient, and that the introduction of additional fault bends results in folds that have more rounded
tops (Fig. 10-51c). Medwedeff (1988) uses wedge structures to model the complex stratigraphic
relationships present at Wheeler Ridge, California. The restored structure and the present struc-
ture, with the corresponding positions of the wells (with their well logs), are shown in Figs.
10-52 and 10-53 respectively. These figures demonstrate how well logs can be used to define the
complex relationships that exist within some structures. Precise correlations and balancing can
be effectively integrated to locate prospects that may not be recognized by normal mapping tech-
niques.
Interference Structures
Would you believe that anticlines can form over synclines with no evidence of an intervening
fault or evidence of more than one deformation? Nevertheless, clear evidence for this seemingly
contradictory relationship can be seen on seismic sections and in outcrops. In the previous sec-
tion on wedge structures, we saw that deformation on a lower level can modify the shape and the
form of dip panels of structures located on a higher structural level. Where structural modifica-
tion of this type results from a single deformation along one thrust surface, as illustrated in Fig.
10-54a, the resulting structures are called interference structures (Suppe 1988).
Interference structures are commonly present where the spacing between ramps is relative-
ly narrow, causing the back dip panel of the leading structure to interfere with the frontal dip
panel of the trailing structure (Fig. 10-54a and b). The interference tends to produce chevron folds
and conjugate kink structures (Weiss 1972; Suppe 1988).
The resulting interference patterns that are created by the deformation are dependent on
ramp spacing, initial cutoff angle, and the total amount of slip. Two model patterns are useful in
Nonclassical Methods
Figure 10-50 Seismic section imaging a triangle zone to the right of sp 250, Raton Basin,
Colorado. (After Applegate and Ross, in Gries and Dyer 1985. Published by permission of
the Rocky Mountain Association of Geologists and the Denver Geophysical Society.)
579
580 Chap. 10 / Compressional Structures: Balancing and Interpretation
mapping these types of structures, although this does not deny the usefulness of other types of
patterns. In the first example, the leading fault bend fold has run up a ramp and the frontal dip
panel of the trailing fold (a monocline) occupies a portion of the lower flat (Fig. 10-54a). The
resulting deformation creates a structure in which a frontal fold lies beneath a structurally high-
er anticline formed by the trailing fold. Flat dips and a syncline exist just above the lower flat,
directly in front of the trailing monocline.
As the deformation progresses, the frontal portions of the trailing monocline will start to run
up the leading ramp (Fig. 10-54b). If both ramps have about the same cutoff angle, then as the
trailing monocline runs up the second ramp, the beds in the frontal dip panel of the trailing mon-
(a)
(b)
(c)
Figure 10-52 Wedge structure in its initial, or restored, state as defined by well logs;
Wheeler Ridge, California, USA. (Published by permission of Don Medwedeff 1988.)
Figure 10-53 Wedge structure in its present state as defined by well logs, Wheeler Ridge,
California, USA. (Published by permission of Don Medwedeff 1988.)
582 Chap. 10 / Compressional Structures: Balancing and Interpretation
ocline will flatten. One result of the deformation is to create a region of nearly flat dips and a nar-
row syncline over the leading ramp as the monocline unfolds. This example once again stresses
the progressive nature of the deformation. Structures were not cast in their present-day positions;
they move, and thus strata bend and rebend. Knowledge of which regions of a fold have been
subject to refolding should aid in the prediction of fracture porosities and better well site loca-
tions, which can result in greater productivity. In our example, some of the strata in the back dip
panel of the leading fault bend fold were first bent backward, and then forward, by the advanc-
ing monocline. As the active axial surfaces sweep through the structure, particular regions with-
in these folds will be subject to repeated deformation and bending that enhances fracture poros-
ity. One can study the refolding by applying the kink method and by modeling increasing amounts
of slip into the structure to study how the deformation progresses.
We make two more points before leaving this subject. First, as the initial cutoff angle
decreases, the structurally higher anticline will move vertically away from the structurally lower
syncline, but at the same time shift to a position where it is located almost directly above the syn-
cline. Second, the two examples presented here are for a clockwise shear within the interfering
frontal and back dip panels. In other words, the beds within the interfering dip panels exhibit a
“Z” vergence (Suppe 1988). An example of an “S” vergence (counterclockwise shear), in which
the frontal dip panel of the trailing monocline passes through the upper anticline, is shown in Fig.
10-54c.
(a)
Figure 10-54 Interference structures (a) and (b) for clockwise deformation and increasing
slip (c) for counterclockwise shear. (Modified after Suppe 1988. Published by permission
of John Suppe).
Nonclassical Methods 583
(c)
EXTENSIONAL STRUCTURES:
BALANCING
AND
INTERPRETATION
INTRODUCTION
In the first edition of this text we mentioned that the balancing of extensional structures was in
the initial stages of development. Since then, several major oil companies have successfully ap-
plied extensional balancing concepts in the northern Gulf of Mexico, Nigeria, Indonesia and
elsewhere (Dula 1991; Nunns 1991; Withjack et al. 1995; Shaw et al. 1997). Accordingly, we
now have a higher degree of confidence in extensional structural and balancing methods and
techniques. As with compressional structural geology and balancing, a common theme to this
chapter is the relationship between hanging wall fold geometry and fault geometry.
This chapter is divided into two parts. The first is the origin of hanging wall anticlines
(commonly called rollover anticlines or rollovers), antithetic and synthetic faults, and keystone
structures, and how knowledge of the genesis of a structure can help find additional hydrocar-
bons (Tearpock et al. 1994). We address such subjects as (1) why some growth faults die in both
the upward and downward directions and what this means to exploration, (2) where rollover anti-
clines are likely to exist or be positioned along major listric normal faults (i.e., faults that exhibit
large vertical expansion across the fault surface), and (3) how faults form prospects. The second
part of the chapter is the study of compaction effects along growth normal faults and how sand-
stone/shale ratios are utilized to project growth faults into poor data areas or, conversely, how
fault shape is used to predict percent sand in the footwall of the fault.
584
Origin of Hanging Wall (Rollover) Anticlines 585
Rollover anticlines have been successfully drilled for many years, yet questions remain concern-
ing their origin and how rollovers form prospects. A major insight into the origin of rollover was
initiated by Hamblin (1965) when he recognized that these strange “reverse drag folds” are the
natural consequence of motion along listric normal faults. He reasoned that if the hanging wall
block separates from the footwall block, a small void opens between the two blocks. Collapse
due to gravity would instantaneously close the void and, as the hanging wall block collapses
onto the footwall block, a reverse drag structure forms. However, Hamblin did not exactly spec-
ify how this gravitational collapse occurs. Gibbs (1984) later recognized from North Sea data
that extensional structures (as an analogy with compressional structures) seemed to form du-
plexes, complete with horses and transfer zones, etc. Yet a number of questions remain to be
fully answered, such as:
1. Can the shallow portion of rollover structures be used (a) to predict the deeper structure
(e.g., subfault structure), or (b) to extrapolate into regions of poor or nonexistent data (e.g.,
into poorly imaged seismic zones)?
2. What is the precise origin of rollover structures, and is it possible to predict the amplitude,
style, and position of a rollover from observable and mappable geologic features?
3. What are the origins of the antithetic, or compensating faults and the keystone structures,
and how do these structures influence rollover geometry?
4. What causes some (perhaps most) antithetic growth faults to exhibit displacements along
their surfaces that die in both the upward and downward directions in a seemingly contra-
dictory relationship?
Answers to at least some of these questions should aid geoscientists in reducing the time
and expense in (1) locating rollovers; (2) generating better interpretations, maps, and prospects;
and (3) isolating trapping mechanisms. Our studies and those of others suggest that compaction,
fault shape, antithetic and synthetic faulting, and cross structures affect the geometry of rollover
structures.
When developing a theory for a structure that is as complex as a rollover, the initial theories
are likely to be overly simplistic. Our position is that a theory is only as good as its ability to
quantitatively explain the observations. Therefore, we expect that in the future, modifications
and refinements are likely to make these theories more exacting.
(a)
(b)
Figure 11-1 Comparison of (a) Brazos Ridge seismic data to (b) computer model, utilizing the
Coulomb breakup theory. (Published by permission of H. Xiao and J. Suppe.)
M
AJ
OR
FA
UL
T
(a)
Figure 11-2 (a) - (e) Coulomb failure or shear collapse model, showing different stages in the de-
velopment of a simple rollover. A bend in the major fault subjects the hanging wall to deformation at
the active axial surface, which is fixed to the bend in the footwall. Increased slip causes the inactive
axial surface to migrate away from the surface of active deformation. (Modified after Suppe 1988.
Published by permission of John Suppe.)
Origin of Hanging Wall (Rollover) Anticlines 587
b
b
VOID
(b)
a
a
a
b
b b
(c)
Figure 11-2 (continued )
588 Chap. 11 / Extensional Structures: Balancing and Interpretation
b
VOID
(d)
(e)
Figure 11-2 (continued )
Origin of Hanging Wall (Rollover) Anticlines 589
As the hanging wall block moves over the footwall block, a small void opens between the
hanging wall and the footwall blocks, as described by Hamblin (1965) (Fig. 11-2b). Gravita-
tional forces cause the hanging wall block to instantaneously collapse into the hole (created by
the sliding) along the Coulomb failure surfaces. In Fig. 11-2b the Coulomb collapse angle is
70 deg with respect to the horizontal. Beds in the hanging wall shear parallel to the Coulomb
failure surfaces, and the material fills the hole, causing the beds to extend (Fig. 11-2c).
Observe in Fig. 11-2b and c that material experiences rotation as it passes through the
Coulomb shear surface that is fixed to the concave upward bend in the major fault. This is a
locus of deformation that extends upward through the strata and, for that reason, this shear sur-
face is called an active axial surface (Fig. 11-2c). Deformation occurs only along active axial
surfaces, which are affixed to the bends in listric faults.
Initially, the material that passes through the bend in the normal fault is deformed at the ac-
tive axial surface and is translated basinward. This material lies adjacent to the inactive axial
surface, basinward of which the strata are undeformed (Fig. 11-2c). The inactive axial surface
rides passively along the straight portion of the fault surface as movement progresses. As this
surface is not affixed to the bend in the fault surface that creates the void, the inactive surface is
passive and does not cause the hanging wall beds to dip toward the fault surface. Notice that the
slip on the fault surface is the distance between the active and inactive fault surfaces, parallel to
the fault surface.
This process is more readily understood if the fault model is subject to another increment of
sliding (Fig. 11-2d). Of course, in nature these increments are infinitesimal. The sliding opens
another void between the hanging wall and the footwall blocks, and the active axial surface that
formed in Fig. 11-2c translates basinward (Fig. 11-2d). However, the void instantaneously closes
by gravitational shear failure, pinning the active axial surface to the bend. The resultant structure
shown in Fig. 11-2e contains a graben-like feature adjacent to the steepest portion of the major
fault and is a monoclinally shaped rollover structure. Sedimentary compaction within the basin
creates basinward dip and closes the structure in the direction to the right in the figure.
You can better understand these processes by redrawing Fig. 11-2e and cutting the hanging
wall block from the footwall block with scissors. Then subject the major fault to another incre-
ment of motion and collapse the suspended material onto the footwall parallel to the Coulomb
failure surfaces.
Growth Sedimentation
In areas like the Gulf of Mexico, most major normal faults are active growth faults. This means
that sedimentation occurs simultaneously with fault slip, and thus the stratigraphic intervals are
subject to vertical expansion across the fault surface (i.e., the intervals thicken on the down-
thrown side of the growth fault). How does this syndepositional sedimentation affect the struc-
ture, and can an analysis of this growth aid us in finding hydrocarbons?
Referring back to Fig. 11-2c, let us assume that a layer of sediment is deposited across the
hanging wall and footwall portions of our structure (Fig. 11-3a). The graben-like feature con-
tains more accommodation space than the top of the rollover, so the sediments will be thickest
over the graben and thinner over the footwall and the crestal portions of the monoclinal rollover.
We saw in Fig. 11-2b through e that the active axial surface is fixed to the bend in the major
fault, and that it represents the locus of deformation of the hanging wall beds. The inactive axial
surface is passive and migrates basinward. Notice in Fig. 11-3a that the inactive surface does not
extend upward into the recently deposited growth sediments. The active axial surface, however, is
a locus of active deformation that affects the entire body of growth sediments. When deformation
590 Chap. 11 / Extensional Structures: Balancing and Interpretation
is viewed as an incremental process, a growth axial surface must connect the active axial surface
to the inactive axial surface, as shown in Fig. 11-3a. You can convince yourself of this statement
by visualizing a thin layer of additional growth sediments deposited above layer 1. This layer
will be horizontal. Another small increment of sliding along the major fault causes Coulomb col-
lapse that deforms this recently deposited layer in the region where the active and the growth
(a)
(b)
Figure 11-3 (a) - (b) Rollover development showing deformation during growth sedimentation. The
most recently deposited sediments are deformed at the point where the active and growth axial sur-
faces converge. (Modified after Suppe 1988. Published by permission of John Suppe.)
Origin of Hanging Wall (Rollover) Anticlines 591
axial surfaces converge (Fig. 11-3a). An additional increment of sliding, combined with growth
sedimentation, produces the geometry observed in Fig. 11-3b. The growth wedge expands as the
growth sediments move through the active axial surface.
For larger rollover structures, sands are more likely to be deposited in the graben, whereas
suspended sediments are more likely to be dominant across the top of the rollover. Therefore, fa-
cies changes and stratigraphic traps are predicted to occur in the vicinity of the growth axial sur-
face. A more realistic example of growth faulting is presented in the following section on
projecting large growth faults to depth.
The growth axial surface can be located on a seismic section by using dip domain analysis
(Tearpock et al. 1994). This surface is located between the steeper-dipping beds at the front of
the structure and the flatter-dipping beds located at the crest. After you locate the growth axial
surface, you can then determine the stratigraphic interval that was deposited as the structure
grew. This is important, as no structure exists to trap hydrocarbons prior to the growth phase. To
determine the growth phase of the structure, follow the growth axial surface downward. The
growth axial surface subparallels the main fault during the time of growth, thus delimiting the
history of the structural growth. The growth axial surface terminates at the inactive axial surface,
and this point indicates the top of the pre-growth strata (Fig. 11-3a and b). No structure existed
during the pre-growth phase of sedimentation to trap hydrocarbons (Fig. 11-2a).
Example of a Rollover Structure. Some listric faults are observed to dip more gently with
depth and then, at greater depth, to increase in dip (Fig. 11-12). This creates a concave down-
ward bend in the major fault. Above this bend, the Coulomb collapse theory predicts that the de-
formation takes place along basinward-dipping conjugate shear surfaces. Shear can occur along
two conjugate surfaces (Billings 1972) and, in our example, this shear occurs in the clockwise
direction (Fig. 11-4a).
The Brazos Ridge, in the northern Gulf of Mexico, can be used to demonstrate the Coulomb
breakup theory for a more realistic case. The Corsair Fault is the major listric normal fault in the
Brazos Ridge area. In the most general sense, the Corsair fault dips at about 45 deg, flattens to
about 10 deg, and then steepens to approximately 20 deg, as represented in a computer model in
Fig. 11-4c. The Brazos region is subject to synchronous deformation and sedimentation, and
thus the sediments deposited in the hanging wall block are growth sediments. If the simple Cor-
sair fault model described in Fig. 11-4c is subject to gravitational (basinward) sliding, two voids
would be created, one above the concave-up portion and the other adjacent to the concave-down
portion of the master fault. Clockwise collapse above the deeper concave downward bend (Fig.
11-4a) and counterclockwise collapse above the concave upward bend (Fig. 11-4b) would pro-
duce the geometry observed in Fig. 11-4c. Notice that there is a relationship between the shape
of the fault and the shape of the rollover in the hanging wall.
If deposition rates are high, as in the Brazos Ridge area, the sediments will fill the low area
adjacent to the shallowest portions of the major fault, thin over the top of the rollover structure,
and then thicken basinward (Fig. 11-4c). The vertical expansion of stratigraphic intervals, as ob-
served in this example, demonstrates why the sediments are referred to as growth sediments.
Also notice the growth axial surface in Fig. 11-4c. It subparallels the growth fault, and it termi-
nates with depth at an inactive axial surface that dips toward the fault (compare to Fig. 11-3b).
The model in Fig. 11-4c contains rollover amplitudes that are higher than observed along the
Brazos Ridge (Fig. 11-1a). However, compaction combined with synthetic and antithetic crestal
faulting reduces the amplitudes of rollovers.
The thickness of the crestal growth section can vary with the shape of the fault. In the case
of the Brazos Ridge, the large expansion fault that creates the structure dips at about a 20-deg
592 Chap. 11 / Extensional Structures: Balancing and Interpretation
(a)
(b)
(c)
Figure 11-4 Deformation of growth sediments at the bends in normal faults. (a) A downward bend
in the fault activates basinward dipping or synthetic shear. (b) An upward bend activates landward or
antithetic shear. (c) Simple generic model of Brazos Ridge rollover development, not including com-
paction. (Published by permission of Xiao and Suppe.)
angle at depth. Thus, as the hanging wall moves over the footwall, the hanging wall slides down
a 20-deg dipping fault surface. This slope causes the hanging wall to drop in elevation relative to
the footwall, and the downward motion creates ample space for the accumulation of thick growth
section over the crest of the structure (Figs. 11-3a and 11-12). Later we show that this fault geom-
etry creates productive four-way closures (Fig. 11-21).
However, if a large expansion fault flattens at depth and is subhorizontal, then as the hanging
wall moves over the footwall, the crest of the rollover structure does not change elevation with re-
spect to the footwall (Fig. 11-5). This fault geometry creates the classic half-graben structure,
with accommodation space developing between the fault and the flank of the half-graben struc-
ture, but not over the crest of the structure. In this case the growth sediments are more likely to
thin onto the flank of structure, creating stratigraphic traps. However, on actively growing struc-
tures, this thinning results from an interaction of stratigraphic and structural processes, classified
under the heading of tectonostratigraphy (Chapter 13). Figure 11-5 shows a seismic profile of a
half-graben structure from the Central Sumatra Basin, Indonesia. As the structure grows, growth
sediments deposited in the half-graben structural low will be subject to axial surface deformation,
as shown in Fig. 11-5. The growth sedimentary section thins dramatically across the growth axial
surfaces and onto the flank of structure (Suppe et al. 1992; Shaw et al. 1997). As growth sedimen-
tation is typically episodic, alternating between high and low sedimentation rates, the local subsi-
Origin of Hanging Wall (Rollover) Anticlines
Figure 11-5 Seismic profile of a half-graben structure from Central Sumatra Basin, Indonesia, showing axial surfaces emanating from bends in the
fault surface. Growth axial surfaces on the flank of the structure (subhorizontal curved lines) define the position where growth section thins onto struc-
ture. The syntectonic sedimentation created potential facies changes and unconformities at the position of growth axial surfaces. (From Shaw, Hook and
Sitohang 1997; AAPG1997, reprinted by permission of the AAPG whose permission is required for further use.)
593
594 Chap. 11 / Extensional Structures: Balancing and Interpretation
dence rate resulting from fault motion may temporarily exceed the sedimentation rate. In this case
coarser sediments may fill the accommodation space in the half-graben (Xiao and Suppe 1992).
As the sedimentation versus subsidence rates fluctuate, the tectonostratigraphic process creates
facies changes and unconformities along strike and down-dip of the growth axial surfaces.
Provided that information exists on the shape of a fault that has created a rollover, the
Coulomb breakup theory can be used to predict the rollover angle (θ) or bed dips. Based on
Coulomb deformation, the rollover angle (θ) is related to the change in fault dip (φ) through a
complicated set of trigonometric formulae, which can be represented by graphs (Fig. 11-6a and
b). The assumptions used in the graphs are that the material in the hanging wall is subject to
Coulomb collapse and that the structure has not experienced sedimentary compaction, as de-
scribed later in this chapter in the section Compaction Effects Along Growth Normal Faults. Thus,
these diagrams strictly apply to the nongrowth phase of the sedimentation and will approximate
the rollover angle within growth sediments. Let us assume that a seismic section images the shal-
low and deeper portions of a listric fault, but poorly images the bed dips. If the above conditions
are met and if the initial fault dip (β) and the fault dip at a deeper level (α1) can be observed on a
depth-converted seismic section, then the rollover dip can be estimated from Fig. 11-6a or b. Fig-
ure 11-6a assumes a Coulomb breakup angle of 60 deg, whereas Fig. 11-6b assumes a 70-deg
breakup angle. For example, the Brazos Ridge generally has a major fault that initially dips at
about 45 deg and flattens to about 15 deg. Therefore, the change in fault dip φ = 30 deg. From Fig.
11-6a and b, the predicted bed dips at the front of the rollover structures are estimated to be 18
deg to 22 deg. Even though the Brazos Ridge growth sediments have been subject to compaction,
these results compare favorably to observed bed dips, which average 20 deg. For presentation pur-
poses, Fig. 11-6a and b can also be used to generate generic or idealized models of real rollover
structures. The generic structure shows how the rollover structure formed.
(a)
Figure 11-6 (a) Theoretical prediction of rollover angle (θ) from fault dips (φ and α1) for a Coulomb
shear angle of 60 deg. Diagram does not include compaction. (b) Same as (a), but for a Coulomb
shear angle of 70 deg. (Published by permission of R. Bischke.)
A Graphical Dip Domain Technique for Projecting Large Growth Faults to Depth 595
(b)
Figure 11-6 (continued )
collapse angles. The Coulomb collapse angle formula requires knowledge of fault throw and has
application to rollover structures in extensional environments anywhere in the world. We also
present a graphical method for predicting Coulomb collapse angles in this section.
Figure 11-7 Rollovers form as the hanging wall collapses onto the footwall. Hanging wall
collapse could occur as inclined (Coulomb) collapse or as vertical collapse. (Modified after
Hamblin 1965. Published by permission of the Geological Society of America.)
A Graphical Dip Domain Technique for Projecting Large Growth Faults to Depth 597
Figure 11-8 Listric normal faults cause the hanging wall beds to diverge toward the fault as the
beds move through bends in fault surfaces. The fault bends define the dip domain boundaries. Young
growth sediments dip at gentle angles relative to the deeper beds, as the shallow beds have not
moved through many fault bends. The older and deeper beds, having moved through many fault
bends, dip at higher angles. (Modified from Xiao and Suppe 1992. Published by permission of the
Gulf Coast Association of Petroleum Geologists.)
inverse technique and calculation errors accumulate. These errors originate from uncertainties re-
sulting from depth correction, geologic and stratigraphic variations, and simple measurement er-
rors. However, both techniques are important tools in extensional structural interpretation.
Projecting Large Normal Faults to Depth
We present a graphical method for projecting large normal faults to depth using hanging wall
beds that have been subject to noticeable rollover. We also present methods for estimating fault
shape and position from the shape of a well-constrained rollover structure.
In order to apply this method, two features must be known: (1) a marker bed that exhibits
noticeable rollover, and (2) the true dip of the fault at the hanging wall cutoff level of the marker
beds. A profile is constructed that must be a dip profile, so that it is perpendicular (as close as
possible) to strike of the fault and strike of the beds. A fault surface map can help you determine
the dip direction of the profile. The structural information can come from well log data, depth-
corrected seismic profiles, or cross sections. If possible, start the projection process by locating a
marker bed above the known level of the main expansion fault. Then use the known portions of
the fault to calibrate and test the projection procedure. If satisfied with the results, apply the
method to the unknown levels of the fault.
Procedures for Projecting Large Normal Faults to Depth
1. If using a seismic profile, depth-correct the seismic section over the interval of concern.
Depth-correct the section on the workstation using local checkshot data.
2. If secondary faulting offsets the marker bed, restore the bed to its unfaulted position before
proceeding further. This is readily done by scanning or digitizing the marker bed into a
graphics program and restoring the upthrown and downthrown cutoffs of the marker bed
(Bischke and Tearpock 1999).
3. In the graphics program, locate a deep marker bed on a level that intersects the main fault.
The bed should exhibit obvious rollover and intersect the imaged portion of the main fault
(Fig. 11-9a).
4. Fit tangents to the marker bed in order to segment the rollover into dip domains (see section
on cross-section construction in Chapter 10). Each dip domain contains a region of roughly
598 Chap. 11 / Extensional Structures: Balancing and Interpretation
(a)
(b)
(c)
(d)
Figure 11-9 (a) - (d) Graphical procedures for projecting a normal fault to depth, using the shape
of a generic rollover. For explanation, see text.
uniform bed dips (Shaw et al. 1997). Curved rollovers will possess several dip domains
(Figs. 11-8 and 11-9a). In our generic example there are four important dip domains: dip
domains 1, 2, 3, and 4 (Fig. 11-9b). Where the tangent lines intersect, construct five dip do-
main boundary lines (S1, S2, S3, S4, and S5) to each of the dip domain panels (Fig. 11-9b).
The dip domain boundaries are either drawn at an angle of 68 deg or 112 deg. In Fig.
11-9b, dip is measured clockwise, with 0 deg to the right and 180 deg to the left. If the
marker bed dips toward the main fault, then these domain boundaries dip at about 112 deg
with respect to the horizontal, and dip in the antithetic direction toward the main fault. If
A Graphical Dip Domain Technique for Projecting Large Growth Faults to Depth 599
the marker bed dips away from the main fault, then these domain boundaries dip at about
68 deg in the synthetic direction, or in the same direction as the main fault. As an alterna-
tive to the above assumed angles, you can derive the collapse angle from the geometry of
rollover structures, as described later in the chapter (Bischke and Tearpock 1999).
5. Project the first dip domain boundary S1 downward to where it intersects the straight-line
projection of the main fault surface (Fig. 11-9c). Where the two surfaces intersect at point
(1), the fault dip decreases and the fault becomes more listric.
6. As the first step to determine the fault position beyond the first bend in the main fault, con-
struct horizontal lines between the right and the left dip domain boundaries, or between
boundaries S1 and S2, S2 and S3, etc. (Fig. 11-9b and c). Then, parallel to the dip domain
boundaries and between the marker bed and the horizontal line, measure the distances D1,
D2, and D3. In other words, measure distances D1 and D2 along the 112-deg inclined dip
domain boundaries. Similarly, measure distance D3 along the 68-deg inclined dip domain
boundary S5 (Fig. 11-9c).
7. Next, project the known portion of the main fault forward, from the left domain boundary
at point (1) to the projection of the right dip domain boundary at point (2) (Fig. 11-9c).
Where the straight-line projection of the main fault intersects the right domain boundary
S2, measure the distance D1 from point (2) upward and in line with the right domain bound-
ary S2. This procedure determines the position of point (3). The position of the main fault is
now known at point (1) and at its projected position at point (3), defined by distance D1.
Next, construct the projected portion of the fault surface between points (1) and (3), estab-
lishing the position of the fault beneath dip domain 1 (Fig. 11-9c).
8. Project the main fault forward from point (3) to point (4) (Fig. 11-9c) and repeat step 7 for
the next fault bend (Fig. 11-9d). For this fault bend, measure distance D2 from point (4) at
the right domain boundary S3 (Fig. 11-9d).
9. If the rollover has a horizontal crest, this means that the fault does not change dip beneath
the flat crest of the rollover. Thus, the main fault dips at the same angle beneath the crest of
the rollover (Fig. 11-9d). In other words, the fault is straight beneath dip domain 3 in Fig.
11-9b. Project the fault beneath the flat crest to dip domain boundary S4 at point (5). At
point (5) the fault turns down and becomes convex where the beds turn down, but what is
the dip of the fault beyond dip domain boundary S4?
10. Repeat step 7 again, but in this case, the fault turns down at the axial surface where the
marker bed turns down. So subtract distance D3 from point (6). Lastly, we construct the
convex projection of the fault surface beneath dip domain 4 (Fig. 11-9d).
This technique is very precise on ideal or generic rollovers, but it suffers from measurement
errors and geologic variations on real rollover structures, where these measurement errors add
through each fault bend. Therefore, the projection technique may deteriorate after projecting the
fault surface through three or four fault bends.
Field Example. The method has application to all extensional regimes. A field example of a
rollover structure from the Gulf of Mexico basin illustrates the procedures used to employ the
fault projection method.
The example is from a 3D depth-corrected profile from the Burgentine Lake Field, onshore
Texas (Fig. 11-10) (Bischke and Tearpock 1999). The horizons and fault surface were tied to
well log data and mapped. The profile was depth-corrected using checkshot data. A graphical
technique for determining the Coulomb collapse angle (described in the next section) derives a
600 Chap. 11 / Extensional Structures: Balancing and Interpretation
(a)
(b)
Figure 11-10 (a) Graphical procedures for determining the Coulomb collapse angle on a rollover
anticline at Burgentine Lake, Texas, USA. For detailed explanations, see text. (b) Predicted fault sur-
face conforms to depth-corrected fault surface [dashed in (a)].
collapse angle ψ of 63 deg for the structure (Fig. 11-10a). This angle is slightly less than the
67-deg to 68-deg collapse angles observed on some Gulf of Mexico rollovers (Xiao and Suppe
1992). Employing the steps described in the proceeding section on procedures, and using a ψ of
63 deg, we generate the cross section shown in Fig. 11-10b. The projected portion of the fault
surface lies over the depth-corrected portion of the fault surface. In this example, the agree-
ment between theory and observation is very good. Therefore, we have shown that the method
has application in working with real rollover structures.
A Graphical Dip Domain Technique for Projecting Large Growth Faults to Depth 601
(a)
(b)
Figure 11-11 (a) - (b) Graphical method for predicting Coulomb collapse angles from rollover
structures. For detailed explanations, see text.
602 Chap. 11 / Extensional Structures: Balancing and Interpretation
dipping beds adjacent to the fault surface, and another tangent line is fit to the flat crest of the
rollover. The technique requires the determination of two displacement parameters: (1) the throw
T on the fault and (2) the vertical distance V between the marker bed’s footwall position and the
position of the marker bed at the crest of the rollover (Fig. 11-11b).
At the position of the marker bed’s hanging wall cutoff, construct the vertical throw vector
line T on the depth-corrected profile (Fig. 11-11b). Determine the vertical distance V. Next, ex-
tend the throw line T by the distance T–V (Fig. 11-11b). Determine length (ᐉ̂) by first locating
the point of intersection between tangent line ᐉ at the front and the tangent line at the crest of the
rollover. Length ᐉ̂ is the distance between this intersection point and the vertical line T (Fig.
11-11b). Measure length ᐉ̂ at the same structural level as the tangent line drawn at the flat crest
of the rollover structure. Next, position a line of length (ᐉ̂) adjacent to the marker bed’s footwall
cutoff position (Fig. 11-11b). The final step in the process is to determine the value of the
Coulomb collapse angle ψ (Fig. 11-11b). The angle ψ is equal to the inclination of the dip do-
main boundary (Fig. 11-11a). From the right-hand termination of line (ᐉ̂), positioned adjacent to
the marker bed’s footwall cut off, construct a dashed line downward to the lower extension of the
T–V line (Fig. 11-11b). Lastly, measure the ψ angle from the horizontal, using a protractor. At
Burgentine Lake, the ψ angle is 63 deg if measured counterclockwise, as dip (Fig. 11-10b).
In conclusion, Hamblin’s (1965) inclined collapse mechanism of rollover formation com-
bined with dip domain analysis provides a graphical method for projecting normal faults into
poor data regions. The fault projection method, when compared to depth-corrected profiles of
fault shape, yields reasonable results. The technique is sensitive to measurement errors and may
deteriorate after projecting a fault through more than a few fault bends. The method is also de-
pendent on the collapse angle (Bischke and Tearpock 1999), so we presented a graphical method
for calculating the Coulomb collapse angle for rollover structures.
603
604 Chap. 11 / Extensional Structures: Balancing and Interpretation
sea floor. The fact that the displacements are negligible near the sea floor is not surprising, as little
time has lapsed since the most recent sediments were deposited. Our examination of the seismic
lines in this area has resulted in two seemingly confusing relationships concerning the slip along
the antithetics and the master synthetic fault: (1) the slip along the antithetics generally terminates
at the master synthetic fault; and (2) the slip along the master synthetic fault in Fig. 11-12 at first
increases from shallow depths but then dies at greater depths. Where does all this slip go, and
does it really go to zero? A close examination of Fig. 11-12 shows that these relationships of
downward-decreasing slip continue along the deep portions of the main listric growth fault shown
at sp A. However, the conservation of fault displacement and the linking of faults, discussed in
Chapter 10, suggest otherwise. In Chapter 10, we make the point that slip along a thrust fault need
not remain constant and that the slip could totally die in the core of a fault propagation fold. We
do, however, account at all times for the changes in slip and how the slip is consumed, or dissi-
pated. In extensional tectonic areas such as shown here, this slip, which appears to vanish so
abruptly, apparently follows the bedding planes. How does this process work?
Backsliding Process
One can envision these large listric growth faults as being within a large, slowly moving land-
slide (Xiao and Suppe 1992). As the hanging wall fault block slips, a void opens between the
hanging wall and the footwall, and the overlying material instantaneously collapses along
Coulomb shear surfaces, producing the rollover (Figs. 11-2 and 11-4). We learned from Fig. 11-6a
and b that the greater the concave bend (φ) in the major normal fault, the greater the rollover angle
(θ). If the dip along the major fault decreases with depth, then eventually the rollover angle may
increase to an angle such that the bed dip (θ) exceeds the dip on the major fault (Fig. 11-13). Fric-
tional failure could then occur along the bedding surfaces that form the front limb of the rollover
structure, and not just along the Coulomb shear surfaces. The higher the dip of the rollover struc-
ture, the closer the sedimentary beds come into parallelism with the antithetic Coulomb failure di-
Figure 11-13 Minimum dip (ω) on Corsair Fault (U.S. Gulf of Mexico) versus maximum bed dips
(θ) at the front of the rollover and above its base. The frontal limb dip on the rollover structures gen-
erally exceeds the gentle dips on the Corsair Fault. (Published by permission of R. Bischke.)
Origin of Synthetic & Antithetic Faults, Keystone Structures, & Downward Dying Growth Faults 605
rection. At some angle of inclination, perhaps less than the Coulomb shear angle, the bedding
planes are likely to present less frictional resistance than a sequence of sedimentary layers. Thus,
frictional failure along the bedding surfaces would be favored over the Coulomb shear surface (of
60 deg to 70 deg), which cuts across the layering at a high angle (Fig. 11-2b). These potential
bedding plane slip surfaces are likely to occur in weak layers, such as overpressured shale zones,
which would present the least possible frictional resistance.
We therefore conclude that backsliding along bedding surfaces is a mechanism that can ac-
count for the apparent termination of the slip along the downward dying synthetic faults and an-
tithetic faults. Using the examples of the Brazos Ridge, the backsliding appears to initiate where
the rollover angle exceeds about 10 deg. The backsliding is a second-order effect relative to the
total amount of slip along the major normal fault – one part of bedding plane slip to about every
seven parts of the slip along the Corsair Fault. Although the bedding plane slip within the frontal
limb of the rollover structure is small compared to the total slip, we shall see that it can have a
marked effect upon the amplitude of the rollover structure.
Backsliding Model. Exactly how does this backsliding mechanism operate? We present two
examples, the first of which outlines the backsliding process, and a second, more realistic model,
which assumes that the backsliding consumes about 1 part in 10 of the amount of slip that occurs
along the major detachment fault. Depending upon geologic conditions, however, backsliding
may at times be the dominant process.
Figure 11-14 illustrates the backsliding mechanism. Initially, slip along the master fault
must occur to an extent such that a critical rollover failure angle is exceeded (Fig. 11-14a). This
critical failure angle depends upon local geologic conditions, such as the pore pressures, and is
likely to vary from area to area. As the pore pressure is unlikely to be large in pre-growth sedi-
ments, backsliding is most likely to occur within the growth sedimentary package. At some time,
a weak rock, such as an overpressured shale, moves through the concave upward bend in the nor-
mal fault surface. When this occurs, an increment of backsliding along the bedding surfaces that
comprise the frontal portions of the rollover structure fills the lower part of the void created by
the basinward sliding (Fig. 11-14b). Initially, the plane of detachment propagates along the bed-
ding plane surface and, in the case of the Brazos Ridge, up to the master synthetic fault. At this
location, the synthetic shear associated with the concave downward bend in the major fault turns
the strata downward in the basinward direction and forms the upper portions of the rollover
(Figs. 11-4c and 11-12).
At the position of the master synthetic fault, the shear within the weak horizon can follow one
of four possible paths. First, the propagating bedding plane slip (overpressured) could also turn
downward and follow the bedding surfaces across the top of the rollover structure. This possibility
seems unlikely, as large amounts of energy would be required. Second, the propagating slip could
cut across the more gently dipping bedding surfaces located near the top of the rollover to a pre-ex-
isting antithetic fault (plane of weakness), and then follow the older antithetic fault to the surface.
This would create small triangular-shaped structures. Third, at the position of the master synthetic,
the growing bedding plane fault could follow a Coulomb failure surface up to the sea floor, creat-
ing a new antithetic fault (Fig. 11-14b). This deformation path appears to be the mechanism of least
resistance and the path most favored by many compensating faults. This mechanism would also re-
sult in antithetic faults that appear to terminate along the master synthetic, which are normally ob-
served on our Brazos Ridge seismic section example. Fourth, the hanging wall block above the
bedding plane detachment and the major fault and master synthetic fault could detach and slump
toward the major fault (not shown in Fig. 11-14). This mechanism would produce a surface of de-
tachment along the master synthetic. This mechanism could be a dominant process in certain areas.
606 Chap. 11 / Extensional Structures: Balancing and Interpretation
(a)
(b)
(c)
Figure 11-14 Backsliding model in growth sedimentary section. (a) Uniform shear along
bedding surfaces causes the hanging wall to detach from the undeformed footwall wedge.
(b) The backsliding forms an antithetic fault and opens a void along the newly formed fault
surface. (c) The upper void is filled by synthetic Coulomb collapse along a basinward-
dipping shear band. The balanced model forms a keystone and downward dying antithetic
and synthetic faults. (Published by permission of R. Bischke.)
We favor the third, and to a lesser extent, the second mechanism to explain the observed antithetic
faulting that we have studied in examples such as the Brazos Ridge.
The third mechanism would cause another void to open along the newly formed anti-
thetic fault (Fig. 11-14b). This hole can only be filled by collapse along the basinward-
dipping Coulomb failure surfaces (Fig. 11-14c), forming a keystone structure. According to
Origin of Synthetic & Antithetic Faults, Keystone Structures, & Downward Dying Growth Faults 607
the balanced deformation model just described, another increment of backsliding will result in
the formation of another antithetic, which forms above the master synthetic and the previously
formed antithetic.
Another consequence of this model is that the bedding plane slip is likely to occur along
shale horizons that could be related to fluctuations in sea level (Payton 1977). As a result, the in-
tervening horizons are more likely to contain sands, or reservoir rock. Notice in Fig. 11-12 that
the antithetic faults die downward into a unit of semicoherent reflections between 2.1 sec and
(a)
(b)
(c)
Figure 11-15 (a) - (i) Generic model of the Corsair Fault illustrating the progressive development
of downward dying antithetic and synthetic faults through growth stages 1 to 7. (Published by per-
mission of R. Bischke.)
608 Chap. 11 / Extensional Structures: Balancing and Interpretation
(d)
(e)
(f)
(g)
Figure 11-15 (continued )
Origin of Synthetic & Antithetic Faults, Keystone Structures, & Downward Dying Growth Faults 609
(h)
(i)
2.4 sec. This unit is an overpressured shale, and we have observed normal faults to die into over-
pressured shale intervals in other structures.
Example from Corsair Trend. A more realistic, generic example is illustrated in Fig. 11-15.
Here the major fault initially dips at 45 deg, decreases to 10 deg on the flat, and then steepens
basinward to 20 deg. This fault configuration is similar to the shape of the Corsair Fault in
Fig. 11-12. Also, the Coulomb failure angle is assumed to be 70 deg and the growth phase of the
sedimentation to begin at Horizon 1 (Fig. 11-15a). In Chapter 13, we describe how to distinguish
the growth interval from the nongrowth interval using growth plots. Since the sediments de-
posited beneath Horizon 1 are pre-growth sediments, and since the backsliding mechanism is
likely to initiate in an overpressured horizon, the backsliding process is not likely to begin until a
critical dip angle is reached within the growth sediments. At this stage, the first increment of
backsliding occurs and a keystone structure forms (Fig. 11-15b) at Horizon 1. Notice in this ex-
ample that the distance between the concave upward bend and the concave downward bend in
the fault is greater than in the example in Fig. 11-14.
Additional sliding along the major fault produces the geometry present in Fig. 11-15c, and
another increment of backsliding results in Fig. 11-15d. The backsliding has the effect of creat-
ing keystone blocks that become younger upward as the blocks grow larger. The basinward slid-
ing along the major fault has deactivated the antithetic fault that formed in Fig. 11-15b, and this
610
Chap. 11 / Extensional Structures: Balancing and Interpretation
Figure 11-16 Jebco Seismic, Inc. migrated line of Brazos Ridge showing antithetic faults that die with depth into the sedimen-
tary basin. Corsair Fault is imaged at sp C and two other less active faults at sp A and B. (Interpretation by R. Bischke. From Xiao
and Suppe 1990. Published by permission of John Suppe.)
Origin of Synthetic & Antithetic Faults, Keystone Structures, & Downward Dying Growth Faults 611
antithetic fault stopped growing after Horizon 2 was deposited (Fig. 11-15c) and was buried by
the more recent growth sediments. The backsliding also has the effect of reducing the rollover
amplitude. After the deposition of Horizon 4, the geometry appears as shown in Fig. 11-15e. At
this stage, the lower two antithetic faults, which formed during the interval of time that Horizons
l to 3 were deposited, have moved through the locus of deformation (active axial surface) that is
associated with the concave downward bend in the major fault. The clockwise shear associated
with this deformation has the effect of slightly rotating the antithetic faults clockwise (Fig. 11-15e).
During the interval of time that Horizon 5 is deposited (Fig. 11-15f), slip along the major fault has
advanced to the stage that the newly forming antithetics initiate to the right of the clockwise
shear-active axial surface. Thus, the antithetics that form after the deposition of Horizon 4 will
not be rotated clockwise by the concave downward bend in the major fault (Fig. 11-15g). Addi-
tional increments of backsliding are shown in Fig. 11-15h and i.
Figure 11-15i is a generalization of the deformation that occurred during the deposition of
Horizons 1 through 7. Dashed lines are drawn at the base of the antithetic faults where the slip
entered the bedding planes, and at the top of the antithetic faults where they ceased to grow and
became inactive (Fig. 11-15i). These lines can be considered axial surfaces that are associated
with the growth phase of the antithetic crestal faulting and the formation of the keystones (Tear-
pock et al. 1994).
Again, notice that the more recent antithetic faulting and keystones form to the left of the
older rollovers. If this model is correct, then it can be tested by data. Figures 11-16 and 11-17 are
two seismic lines from the northern Gulf of Mexico and Brunei (SE Asia) respectively. On both
of these lines, the antithetic faults have a tendency to be younger upward and to be older in the
deeper sediments. The pattern of downward dying faults and keystone blocks is similar to our
example in Fig. 11-15.
Figure 11-17 Seismic line from Brunei (Borneo) showing antithetic faults that appear to terminate
along bedding surfaces. The antithetic faulting also dies with depth into the basin. (Published by per-
mission of Muzium Brunei.)
612 Chap. 11 / Extensional Structures: Balancing and Interpretation
1. Growth antithetic faults form basinward and above synthetic faults located near the crests
of rollovers.
2. The antithetics become younger upward with the older antithetics being positioned basin-
ward and terminating at a deeper level.
3. Slip along the master synthetic may die with depth, and slip along the antithetic faults ap-
pears to terminate at the position of the master synthetic.
4. The deformation mechanism forms a keystone structure, or a graben, which with the anti-
thetic faults has the effect of reducing the amplitude of the rollover structure.
5. The deformation and sedimentary mechanisms form faults that die in both the upward and
downward directions.
6. The process may be controlled by the sedimentary cycle.
613
614 Chap. 11 / Extensional Structures: Balancing and Interpretation
presently active, although at much lower slip rates (Fig. 11-16, sp A and B). In Fig. 11-16, the
Corsair Fault is seen to surface at sp C. On the strike lines, these deeper faults can be observed to
intersect the Corsair Fault (Fig. 11-18, sp B), to offset the fault, or to form cross structures.
These cross structures trend subperpendicular to the Corsair Fault and deform and partition
the footwall into a series of high-gradient and low-gradient areas (Fig. 11-19). On a regional
scale, the series of highs and lows on the Corsair Fault, which are caused by the major subfault
deformation, are bounded by the cross structures (Fig. 11-20). The low-gradient, shelf-like, or
“flat” areas on the Corsair Fault produce what we term bows on the fault surface. Bows are
known in the U.S. Gulf of Mexico to be a key indicator of a petroleum trap. We propose the term
chute for the higher gradient regions of indentation into the fault surface. The strike line (Fig.
11-18) images chute 1 (Fig. 11-20) at sp B.
Finally, we demonstrate that the general position of the rollover structures can be predicted
from the fault surface map of the Corsair Fault (Fig. 11-19), with the bows (lowest gradient
areas) on the fault surface corresponding to the position of the structural highs within the
rollover structure, and the chutes (higher gradient areas) corresponding to the position of the sad-
dles between the structural highs. These relationships are shown in Fig. 11-21, which is a re-
stored time map of a horizon within the rollover structure. The map was restored to a common
level by simply closing the faults. A comparison of Fig. 11-21 to Figs. 11-19 and 11-20 demon-
strates that the bows correlate to the position of the structural highs, whereas the chutes corre-
late to the saddles. Thus, rollover closures can be located from fault surface maps, which further
demonstrates the value of constructing these maps (Chapter 7).
What does the saddle geometry look like in 3D? Fig. 11-22 is a 3D perspective view of the
flanks of two structures within a large rollover structure, based on a study of axial surface defor-
mation in the Central Sumatra Basin, Indonesia (Shaw et al. 1997). The fault surface contains a
central low, or chute, and the chute is bordered on its flanks by two fault surface highs. As the
hanging wall block moves down the fault surface, the displaced horizons conform to the shape of
the 3D low in the fault surface, forming a saddle and the flanks of the adjoining structures, as de-
picted in the inset in Fig. 11-22. The geometry resembles that observed on the strike line (Fig.
11-18) between sp D and sp E, where the flank of a closure in the hanging wall block is imaged.
The structures flanking the low dip toward the chute in the fault surface (Fig. 11-22).
STRIKE-RAMP PITFALL
We discuss the necessity of constructing accurate maps throughout this book, and we emphasize
that here by presenting examples of the pitfall in interpreting strike-ramps between en echelon
normal faults, which are common in extensional terranes (Tearpock et al. 1994). We demonstrate
how to recognize, locate, and avoid this costly pitfall. We have observed this problem and its re-
sultant dry wells on a number of 2D and 3D data sets. The examples illustrate the necessity of
constructing accurate maps of subsurface faults.
Strike-ramps form along arrays of normal faults that contain en echelon offsets, or stepovers
(Morley et al. 1990; Peacock and Sanderson 1991). Strike-ramps are commonly ubiquitous fea-
tures on extensional data sets. Examine the air photograph shown in Fig. 11-23, taken from the
Yucca Mountains, Nevada (Ferrill et al. 1999). Notice how the Boomerang fault almost connects
to the uppermost extension of the Fatigue Wash fault, but that no connecting fault links the two
fault systems to form a two-way fault closure. On 3D data sets, the seismic data may become
semicoherent, or deteriorate, where two faults overlap, perhaps due to the fault shadow effect. If
this occurs, then structural horizon mapping may result in the mapping of two fault surfaces as a
single fault surface. In addition, on 2D and 3D seismic data sets, structural aliasing can result in
Strike-Ramp Pitfall 615
Figure 11-19 Fault surface map of a portion of the Brazos Ridge. The structural contours deepen
toward the lower portion of the diagram. The cross structures segment the fault surface map into a
series of low-gradient areas (bows) and high-gradient areas (chutes). (Prepared by W. L. Keyser.
Published by permission of Texaco USA, Eastern E & P Region.)
Figure 11-20 Location of chutes and bows. The chutes exist as intrusions into, and the bows exist as
protrusions upon, the fault surface map (Fig. 11-19). Five bows and chutes exist along trend, with chutes
3 and 4 being a double chute. (Published by permission of Texaco USA, Eastern E & P Region.)
Figure 11-21 Partially restored horizon time map. The structural highs correspond to the position
of bows, whereas the saddles correlate to the chutes. (Published by permission of Texaco USA,
Eastern E & P Region.)
616 Chap. 11 / Extensional Structures: Balancing and Interpretation
Fault Contours
Active axial surfaces
Fault bends
Z
igh
lt h
fau
ow
lt l bed strike
fau
& dip
Z
igh
X
lt h
fau
Intermediate Y
fault panel
Constant horizontal
extension
Z Z
fold low
fault low
inset not shown to scale
Figure 11-22 Perspective view of a 3D structural saddle based upon a fault surface map from
Central Sumatra Basin, Indonesia. The low area on the fault surface generates a structural low
(Modified after Shaw, Hook and Sitohang 1997; AAPG1997, reprinted by permission of the APPG
whose permission is required for further use.)
mapping the two overlapping faults as one fault (Fig. 11-24a and b). Furthermore, this overlap
has a limited lateral zone. Even if the data are good, en echelon faults may be overlooked if the
interpretation is based on insufficiently close seismic lines.
Notice on Fig. 11-24a that the strata downthrown to the overlapping faults form a structural
low relative to the upthrown strata. Where Fault 1 dies out, or loses displacement to the east,
Fault 2 increases in displacement to the east. This reversal of displacements between the offset-
ting fault surfaces causes the strata within the overlap area to dip to the west, forming a strike-
ramp. The pitfall is that this ramp may or may not contain a fault that would form a three-way
fault closure (Brenneke 1995).
Let us assume that the strike ramp shown on Fig. 11-24a is within a 3D survey area, that a
fault shadow effect causes the data to deteriorate between the offset fault surfaces, and that the
data were interpreted along equally spaced lines. If on the 3D data set the horizontal distance be-
tween the two overlapping fault surfaces is large, then seismic horizon mapping should detect
the strike-ramp. On narrower strike-ramps, however, structural horizon mapping of the offset
fault surfaces is subject to possible fault miscorrelation problems. Geoscientists may assume that
the fault surfaces are throughgoing and, in this case, the deterioration of the seismic data set may
not resolve that two faults are present. If the geoscientist connects Fault 1 to Fault 2, then the
Strike-Ramp Pitfall 617
Figure 11-23 En echelon normal fault surfaces from Yucca Mountains, Nevada, USA, showing
faults and fault offsets in bold lines. An unfailed block, or strike-ramp, separates the Boomerang fault
from the Fatigue Wash fault. Failure of the block between the Fatigue Wash and Northern Windy
Wash faults creates the West Ridge cross fault. In the subsurface, a structural geometry identical to
the Yucca Mountains normal fault system would trap hydrocarbons in the block downthrown to the
West Ridge connecting fault system, but not in the block between the Boomerang and the Fatigue
Wash Faults. (From Ferrill, Stamatakos, and Simms 1998. Published by permission of Elsevier Sci-
ence, Journal of Structural Geology.)
result is a nonexistent prospect based on a three-way closure (Fig. 11-24b). In practice, the data
are commonly incoherent between the offset fault surfaces, so in such a case, interpreting every
line on a 3D data set may not even resolve this problem.
How does one locate, recognize, and resolve fault correlation problems? Figure 11-25
shows a fault surface map that contains two en echelon faults, which are incorrectly mapped as a
single fault. Can you detect the strike-ramp between the two faults? The strike-ramp exists as a
minor bend or curve on the fault surface map (Brenneke 1995).
We are aware of two methods to prevent this costly pitfall, both of which involve fault sur-
face mapping. Geoscientists can analyze 3D coherency data over many time slices (Fig. 11-26).
Coherency data commonly image the en echelon offsets associated with strike-ramps, particu-
larly on the shallow structural levels. When interpreting seismic data, assign a fault segment to
every fault that exists on the multilevel time slices. These interpreted fault segments on the time
slices represent strike (trace) segments relative to the dip profiles. Thus, faults observed on time
slices easily tie to faults observed on dip profiles.
Lacking coherency data on 3D data sets, or when using 2D data, you can locate en echelon
offsets by constructing fault surface maps, as described in Chapter 7. If the fault surface map
contains a kink or a bend, as in Fig. 11-25, then en echelon faults may be present. Two overlap-
618 Chap. 11 / Extensional Structures: Balancing and Interpretation
(a)
(b)
Figure 11-24 (a) Two en echelon faults die out in opposite directions, creating a strike-ramp. The
unfaulted ramp is incapable of trapping hydrocarbons. An improperly spaced 2D data set, structural
aliasing, or poor 3D seismic data in the area of fault overlap could result in the prospect shown in
(b). (b) If fault surface maps are not constructed, then a combination of incomplete seismic control,
structural aliasing, or poor 3D seismic data can create a nonexistent prospect. (Published by per-
mission of R. Bischke.)
ping fault surfaces, separated by a significant horizontal distance, will contain a contouring bend
if mapped as a single fault surface. However, two closely spaced faults may exhibit only a subtle
kink or bend in the fault surface. When using 3D data, we can readily solve the problem of dis-
tinguishing between a gentle bend in a single fault from an en echelon offset of two fault sur-
faces. Construct several arbitrary lines across the kink or the bend in the fault surface, at oblique
Strike-Ramp Pitfall 619
Figure 11-25 Fault surface map of two en echelon fault surfaces mapped as a single fault surface.
Bend in fault surface map suggests the possible presence of two faults that do not intersect, rather
than one fault. Map is in two-way time (msec). It was generated using a triangulation algorithm that
does not smooth through bends in fault surfaces. Compare to the coherency cube display in Fig. 11-26
that clearly shows the presence of two separated faults. (Published by permission of R. Bischke.)
angles to the kink. Where strike-ramps are present, these arbitrary lines will commonly en-
counter regions of coherent data in the areas where the two fault surfaces do not overlap. Ac-
cordingly, two faults, and not one fault, will image on the arbitrary lines.
In order to solve the strike-ramp problem on 2D data sets, properly oriented oblique or dip
profiles must exist that cross both faults. These lines may not exist. Thus, be suspicious of any
bend, kink, or curve in a normal fault surface that acts as a potential trapping fault for a prospect.
Risk this bend in the fault accordingly. An additional line or two may need to be acquired to re-
solve the problem.
Figure 11-26 Strike-ramp on coherency cube data taken from St. Charles Ranch, Texas. Just
above the dashed line, the en echelon pattern of the faults is evident. The dashed line represents a
fault surface contour on Fig. 11-25, where the curvature of the contours is a clue to possible en ech-
elon faults. (Published by permission of R. Bischke.)
620 Chap. 11 / Extensional Structures: Balancing and Interpretation
Alternatively, additional motions along the strike-ramp may generate faults that link two
originally separate faults, creating a fault trap (Tearpock et al. 1994). The probability of an accu-
mulation in this type of trap depends on the timing of these younger faults. Chapter 13 in this
book and Tearpock et al. (1994) present methods for studying fault timing.
Geoscientists should exercise care when interpreting faults and constructing fault surface
maps. Geoscientists loop-tie fault surfaces maps for the same reason that they tie horizon sur-
faces: to make sure they remain on the same fault. Also, consider that on 3D data sets, randomly
placed arbitrary lines may or may not locate the strike-ramps. Strike-ramps that exist along faults
that constitute the same fault system are often subtle features on 3D data sets. Geoscientists must
know where to place the arbitrary lines in order to demonstrate that a trapping fault exists. Cor-
rectly interpreted and loop-tied fault surface maps show geoscientists where to place the arbi-
trary lines. Fault surface mapping also reduces the horizon mis-tie problem, often saving time.
The final product is a higher quality interpretation and prospects.
After burial, column A will dewater and compact and take on the shape of column B (Fig. 11-27).
Column B has a buried width, height, fault dip, and porosity of ∆X′, ∆Y′, θ, and φ respec-
tively. Therefore, the fault dip at B is defined as
However, the compaction primarily affects the height and not the width; therefore,
∆X = ∆X′ (11-3)
Compaction Effects Along Growth Normal Faults 621
Figure 11-27 Compaction and burial along a growth normal fault. Element A at the sea floor com-
pacts to the geometry present at B. These relationships are taken relative to the compacted footwall.
(From Xiao and Suppe 1989; AAPG1989, reprinted by permission of the AAPG whose permission
is required for further use.)
As porosity controls the extent of compaction (Baldwin and Butler 1985), this mass conser-
vation formula written in terms of solid volume is
Therefore, if we are able to predict the initial fault dip (θo) and initial porosity (φo) along
with the porosity (φ) of the rock at any given depth level, we can determine the fault dip (θ) at
any level of interest (Fig. 11-27).
The process of dewatering and compaction is rapid, occurring within about the upper 700 ft
(Baldwin and Butler 1985), so any compaction equation is controlled primarily by the porosity,
which is in turn related to the sandstone/shale ratio. Thus, if the depths of a normal fault are
known in each of several neighboring wells, and if the sandstone/shale ratios are known from
local or adjoining footwall well logs, then the fault shape and its location can be extrapolated be-
tween the adjoining wells using Eq. (11-6). Case histories of the process are presented later in
this section.
622 Chap. 11 / Extensional Structures: Balancing and Interpretation
Before applying Eq. (11-6) to an area, we must be able to calculate the amount of porosity
to be expected in sandstone and shale horizons at any given depth. Porosity/depth equations are
dependent on the region being studied (Baldwin and Butler 1985). These equations represent the
average porosity of sandstone or shale at any given depth, so determine these relationships by
using local data. The porosity/depth equations represent averages, so results will also represent
averages.
For the United States Gulf Coast, the following empirically derived sandstone and shale
porosity/depth equations can be applied (Xiao and Suppe 1989). For shale,
and
Alternatively, you may use the following equation (Atwater and Miller 1965).
From Eq. (11-9) we can determine that the initial sandstone porosity is about 39 percent by
setting the depth to zero. Porosity/depth equations for areas other than the Gulf of Mexico are
presented by Baldwin and Butler (1985) and Sclater and Christie (1980).
The technique can now be applied. First, the sandstone portions of a well are determined by
locating the sandstone and the shale base lines from SP logs. Common industry practice suggests
that sandstones exist wherever the SP log exceeds 25 percent of the distance between the sand-
stone and shale baselines (Fig. 11-28). Values of less than 25 percent on the divided SP log
are taken to be shale (Schlumberger 1987). Second, a “splicing method” is employed, using
Fig. 11-29. In those depth intervals that have been determined from the SP log to be sandstone,
use Eqs. (11-8), (11-9), or (11-10), and use Eq. (11-7) in those depth intervals that represent shale.
Equation (11-6) is also dependent on the initial fault dip and initial porosities. As an exam-
ple, for the northern Gulf of Mexico these values have been empirically determined to be 67.5 deg
for initial fault dip, and 39 percent and 68 percent porosity for sandstone and shale, respectively
(Xiao and Suppe 1989). You may wish to confirm these initial values from local data prior to ap-
plying Eq. (11-6) to your area of interest. We have now determined the sandstone and the shale
portions of the sedimentary section from the SP logs and can calculate their average porosity at
any given depth from Eqs. (11-7) to (11-10). Therefore, we can now calculate the average fault
Compaction Effects Along Growth Normal Faults 623
Figure 11-28 Relationships utilized to determine sand/shale ratios. Sand is present wherever the
SP log deviates more than 25 percent off the shale base line. (From Xiao and Suppe 1989;
AAPG1989, reprinted by permission of the AAPG whose permission is required for further use.)
dip (θ) at any depth from Eq. (11-6) and the “splicing method.” These formulae were tested with
good results in a number of examples in the Gulf of Mexico, such as the one shown in Fig. 11-30.
In areas of poor seismic data, or in areas with well log data and limited seismic data, the
technique can be used to help interpret the shape of important growth faults. The shape of a fault
with depth may be important in generating a prospect, designing a prospective well location, or
the determination of reserves based on the areal extent of a reservoir.
Perhaps a prospect is under evaluation and the location of the trapping fault is of critical im-
portance. The method can be used as a “quick look” technique (Tearpock et al. 1994) to evaluate
the presented interpretation and, in particular, the shape and location of the trapping fault.
Prospect Example
A good application of this method arises when analyzing a prospect. Examine Fig. 11-31, show-
ing Well No. 1, which penetrates a large growth fault, and Well No. 2, which has a discovery at
the D Sand level on a rollover structure. Seismic data from this area are not of the best quality,
but they image gentle south dips. The proposed well, to be drilled up-dip of the discovery well,
looks like a lead-pipe cinch! Would you participate in the proposed well? If not, why not?
624 Chap. 11 / Extensional Structures: Balancing and Interpretation
Figure 11-29 The “splicing” method for estimating fault dips from sand/shale horizons determined
from SP logs. (From Xiao and Suppe 1989; AAPG1989, reprinted by permission of the AAPG
whose permission is required for further use.)
Let’s do a quick examination of the cross section in Fig. 11-31. We first observe that the
normal fault is planar in shape, dipping at an angle of about 55 deg. The discovery and proposed
wells appear to be in the hanging wall rollover anticline formed by the fault, which is penetrated
by Well No. 1.
Several questions come to mind. Why is the fault in Fig. 11-31 not listric? Why do the beds
dip or roll toward a fault surface that is a planar surface? Geoscientists know that rollover struc-
tures are associated with listric fault surfaces; i.e., rollover structures form above fault surfaces
that are not planar. In the section on the origin of rollover in this chapter, we demonstrate that
only above listric fault surfaces do hanging wall strata form a rollover anticline.
Let’s take a look at the fault surface shown in Fig. 11-31 and refer back to Fig. 11-29. Fig-
ure 11-29 allows us to estimate fault dips from lithology. The lithology in the footwall of the
fault in Well No. 1 allows us to predict the dip and shape of the fault. Well No. 1, which pene-
trates the fault near –4000 ft, contains about 50 percent sand in the footwall block between the
–4000-ft and –8000-ft levels (–1200 m to –2400 m). The average depth of the sandy section is
–6000 ft (–1800 m). On Fig. 11-29, a fault at a depth of –6000 ft (–1800 m) in a 50 percent sand
section will dip at an angle that lies halfway between the fault shape in pure sandstone and the
fault shape in pure shale. We can fit a protractor to this figure to obtain a fault dip of about 55
deg. The fault shown on Fig. 11-31 dips about 55 deg between –4000 ft and –8000 ft, exhibiting
good agreement between observation and theory.
Notice that Well No. 1 penetrates a predominantly shale section below –8000 ft (–2400 m).
Thus, the footwall section below –8000 ft should be subject to significant compaction. Faults in
compacting shales will dip at gentler angles than faults traversing sand-rich sections. Let’s take
another look at Fig. 11-29, but this time we assume a pure shale section between –8000 ft and
–20,000 ft (–2400 m to –6100 m). The average depth of this pure shale section is about –14,000
ft (–4250 m). Fit a protractor to the fault shape on the pure shale curve (Fig. 11-29) at an aver-
age depth of –14,000 ft. Figure 11-29 predicts that the fault will dip at an average of 40 deg
below –8000 ft. If the fault dips at an average of 40 deg below –8000 ft, then the fault should fol-
Compaction Effects Along Growth Normal Faults 625
Figure 11-30 Test example for projecting fault dips between wells. The SP log from Well No. 1, the
fault dip (Fig. 11-27), and the porosity/depth equations are utilized to predict the shape and position
of Fault A with depth. (From Xiao and Suppe 1989; AAPG1989, reprinted by permission of the
AAPG whose permission is required for further use.)
low the path shown in Fig. 11-32. The listric bend in the fault causes the strata to roll toward the
fault and to form a rollover anticline in the hanging wall block (see section on origin of rollover).
Furthermore, the listric shape of the fault causes the strata to dip at a higher angle than those
shown in Fig. 11-31 and causes the fault to be farther to the south than previously interpreted.
The result of our quick analysis is to predict the possibility that the productive D Sand will be
faulted out in the proposed well (Fig. 11-32). What appeared at first to be a sure thing may not
be. An alternative to the quick look method is to use a spreadsheet and the procedures outlined in
this section to write a program using Eqs. (11-6), (11-7), (11-8), (11-9), and (11-10). The pro-
gram can be used to predict the fault dip through a varying sandstone/shale section.
626 Chap. 11 / Extensional Structures: Balancing and Interpretation
Fig. 11-31 Proposed well position located up-dip of the producing Well No. 2, and downthrown to a
planar fault surface that intersects Well No. 1 at –4000 ft. Quick look techniques can determine
structural viability of the proposed production well. (Published by permission of R. Bischke.)
Fig. 11-32 The SP log in Well No. 1 and the compaction formula result in a listric fault surface that
dips at a lower angle below –8,000 ft, causing the interpreted fault to displace to the south. Listric
shape of the fault surface causes the beds to roll toward the fault surface. Given this structure and
the relocated fault, a good probability exists for the reservoir being faulted out in the proposed well.
(Published by permission of R. Bischke.)
Compaction Effects Along Growth Normal Faults 627
The importance of this example is to stress that there is often a geometric relationship of
fault shape to fold shape, and it can be used when generating geologic interpretations. The cross
section shown in Fig. 11-31 does not conform to certain basic geologic principles. Therefore, a
good possibility exists that the interpretation is in error. The error may be small or large. The
quick look technique indicates that the error, in this case, may be large enough to result in a dry
hole. This quick analysis should direct the prospect generator to additional work to either sub-
stantiate the interpretation, as shown in Fig. 11-32, or modify the interpretation based on the
principles shown in this section on the effects of compaction on the shape of growth normal
faults.
The compaction work presented by Xiao and Suppe (1989) can result in accurate predic-
tions of fault shape where compaction is the predominant process. The method is stable and
tends to average the well log data. As the method predicts the average fault dip based on the av-
erage lithology over an interval, the method averages through intervals that do not obey the ide-
alized compaction formulae. Thus, the method tends to average out local and aberrant secondary
cementation intervals that may affect the compaction curves. Deformation of the fault surface,
either by salt or by another fault, may affect the results of the method.
porosity depth curves. This information can be derived from the nearest wells, which in frontier
areas could be many miles from the area of interest.
First we must determine a sandstone/shale ratio formula. This can be accomplished using
Fig. 11-33. In practice, the observed height of the figure is the top and the bottom of a portion of
an interpreted growth normal fault on a depth-corrected seismic section. This distance represents
a column of interbedded sandstone and shale. Using a deck of cards analogy, the sandstone lay-
ers are removed from the deck and placed at its bottom.
We now calculate
where
Figure 11-33 An alternating sequence of sand and shale beds can be generalized into a shale in-
terval of thickness hsh and a sandstone layer of thickness hss. As shales have a higher initial porosity
than sandstones, shales compact more than sandstones. (Published by permission of R. Bischke.)
Compaction Effects Along Growth Normal Faults 629
r–calc = hss/hsh = [sin (θobs – θsh) sin (θss)]/[sin (θss – θobs) sin (θsh)] (11-11)
or
r–calc = [(tan θobs /tan θsh ) – 1.0]/[1.0 – (tan θobs/tan θss)] (11-12)
where
Equation (11-12) is less sensitive to measurement errors than the preceding Eq. (11-11),
which was presented in the first edition of this book.
Sandstone/shale ratios can be converted to percent sand by the following equation:
As θsh and θss can be obtained from Eq. (11-6) and the porosity from Eqs. (11-7) to (11-10),
we can calculate the average porosity, or sand/shale, curve with increasing depth.
Before proceeding to a test case, we will outline a number of factors that affect the accuracy
of our methods.
1. The method applies only to growth normal faults, and thus to areas that are experiencing
both active sedimentation and extension.
2. Although the method can be applied wherever velocity information is available, the veloc-
ity function is critical to obtaining reliable results because it is used to determine depth of
the fault. Velocity determinations are discussed in Chapter 5.
3. The method in its present stage of development involves only compaction, and thus defor-
mation of normal faults by other processes, such as salt flow, may invalidate the method.
4. The method is restricted to the depth to which the seismic reflections can be resolved.
5. The method applies only to sandstone/shale lithologies, and not to carbonates.
To test the method, we compare a depth-corrected growth normal fault to well log data. In
order to compare observed well log results, which resolve horizons to within feet, to seismic
data, in which the resolution is a function of depth, the well log data must be converted to longer
wavelengths. This is accomplished by first determining the sand and shale horizons from SP
logs, as discussed earlier (Fig. 11-28), and then measuring the wavelength of coherent reflectors
directly on the seismic section. The frequency content of the seismic section is thus determined.
As the frequency content decreases with depth, the wavelength of the reflections increases with
depth. This wavelength is then depth-corrected and passed over the SP data as a moving average,
with the length of the operator increasing with depth. The result of this averaging process from
an example well is the sandstone/shale depth curve shown in Fig. 11-34. The data are plotted at a
frequency of two seismic wavelengths. This long wavelength well log data can be directly com-
pared to normal fault calculations based on seismic data. Typical results, calculated using Eq.
(11-11) and based on a growth fault intersected by the well, are shown in Fig. 11-35. The results
630 Chap. 11 / Extensional Structures: Balancing and Interpretation
Figure 11-34 Processed sandstone/shale depth curve obtained by averaging sand/shale horizons
taken from an SP log. The data are plotted at a frequency of two seismic wavelengths and range in
depth from –3000 ft to –10,000 ft. (Published by permission of R. Bischke.)
approximate the well log data. The seismic data was processed utilizing stacking velocities. The
normal fault, which terminates at a depth of –7000 ft, has been used to predict the higher sand-
stone/shale ratios present in the –5000-ft to –6000-ft level.
Elements of the sandstone-shale theory were known to industry for many years and were
applied worldwide. As a final example, we examine a field in Texas where an early form of the
method was applied in the 1960s.
Example from Segno Field, Polk County, Texas. The Segno Field rollover structure
formed downthrown to an east-west trending normal fault (Fig. 11-36a). Examine the footwall
portion of cross section A-A′ constructed from dense well control from the eastern flank of the
field (Fig. 11-36b). This section from Gow (1962) through the Tertiary Jackson, Cook Moun-
tain, Sparta, and Weches shale sections shows that the normal fault dips at angles of 30 deg to
40 deg. Gow wrote on the profile, “Note that fault plane dip steepens in sand formations”
(Gow 1962). In the Yegua and Wilcox sand sections the fault surface steepens to about 50 deg
to 55 deg (Fig. 11-36b).
Compaction Effects Along Growth Normal Faults 631
Figure 11-35 Plot showing predicted sandstone/shale depth curve derived from fault dips inter-
preted on seismic profile. The fault is adjacent to the well data shown in Fig. 11-34. The xs curve
uses the Xiao-Suppe compaction equation, whereas the at curve denotes the Atwater equation.
(Published by permission of R. Bischke.)
A change in fault dip angle from 30-40 deg to 50-55 deg is typically apparent on depth-
corrected seismic profiles. Indeed, Roux (1978) presents an example of a fault that changes dip
through a thick sand and shale section on a vintage seismic time profile. This profile is in Low-
ell’s book on structural styles (Lowell 1985). Industry recognized, and at times used, this
“change in fault dip technique” to locate sand sections during the 1960s and 1970s. Subsequent
bright spot analysis apparently replaced the technique, but geoscientists are again using the tech-
nique where direct hydrocarbon indicators are not applicable or not in support of AVO analysis.
We obtained a seismic line that crosses the crest of the structure. The line was digitized and
depth-corrected using interval velocities. Interval velocities approximate true velocities and are
always available to the geoscientist. On the profile (Fig. 11-36b), the fault dips at high angles
through the Yegua and Wilcox sands and at less steep angles through the Jackson, Cook Moun-
tain, and Sparta shale sections (Bischke and Tearpock 1993). After picking the fault on the vin-
tage seismic line, the inflection points on the fault surface were depth-corrected for input into a
632 Chap. 11 / Extensional Structures: Balancing and Interpretation
(a)
(b)
Figure 11-36 (a) Map of Segno Field, Texas, USA, with location of Cross Section A-A′. (b) Section
A-A′ shows growth fault steepening in sand intervals. As Gow (1962) noted on the section, “Note
that fault plane steepens in sand formations.” (From Bischke and Tearpock 1993. Redrawn from Gow
1962. Published by permission of the Houston Geological Society.)
sand percentage prediction program. This program uses formulas presented in the previous sec-
tion. Program results shown in Fig. 11-37 are comparable to Well No. 8 in the footwall block
(Fig. 11-36a), located adjacent to the seismic line (Bischke and Tearpock 1993). In Fig 11-37,
the diamond-shaped bar on the left shows the resulting computations from the Atwater/Miller
sand compaction formula (Eq. 11-10), whereas the open box bar on the right defines the Xiao/
Suppe compaction curve results (Eqs. 11-8 and 11-9). The Atwater/Miller sand compaction
Compaction Effects Along Growth Normal Faults 633
Figure 11-37 Percent sand plot for Segno Field growth fault, based on changes in dip of fault.
Fault steepens through Yegua and Wilcox sand sections, generating a high sand percentage re-
sponse from about 5000 ft to 7000 ft and below about 9000 ft. Compare calculated sand percent-
ages to sand percentages based on log of footwall Well No. 8 (closed box bar on graph). (From
Bischke and Tearpock 1993. Published by permission of the Houston Geological Society.)
formula is more conservative than the Xiao/Suppe formula. The closed box bar (middle) is the
percent sand based on the Well No. 8 log.
When subject to computation, the interpreted fault trace predicts a sand response on the
level of the Yegua sand between the depths of about –5000 ft and –7000 ft. The Atwater/Miller
and Xiao/Suppe bars bracket the sand percentage in Well No. 8 and predict a sand percentage of
about 60 percent (Fig. 11-37). Between the –5000-ft and –6700-ft depths in Well No. 8, the sand
percent is about 60 percent (Bischke and Tearpock 1993). An examination of Fig. 11-37 indi-
cates that a shale section should exist at about the –7000-ft to –8800-ft level and that sand should
exist below –8800 ft. Well No. 8 contains shale between –6700 ft and –8200 ft before encounter-
ing the Wilcox sand (Fig. 11-36b). These estimates, which are a good match to the well data, are
based on imprecise interval velocities, and better velocity estimates should improve the depth
predictions even more.
The method works best on seismic sections where you can pick the fault trace accurately
and where thick sand and shale sections are present. Fault surfaces also respond to interbedded
sand/shale sections, and the section need not be predominantly sand (Bischke and Tearpock
1993; Tearpock and Bischke 1991). Why drill a well to a depth of 16,000 ft if there is no direct
evidence for sand at that level, without first using fault dip analysis to help reduce your sand
risk?
This method has broad application to petroleum exploration, and we have found that the
method is robust in that useful results are obtained even from imprecise velocity data, such as in-
634 Chap. 11 / Extensional Structures: Balancing and Interpretation
terval velocities (Bischke and Suppe l990a; Bischke and Tearpock 1993). Normally, when a well
is drilled, one of the least known parameters is the sand/shale ratio of the reservoir. As the sec-
tion in the Segno Field example contains shale from –7000 ft to –8800 ft (Fig. 11-37), the
method also produces a positive prediction for seal. Such information can be used by a geoscien-
tist or engineer to better determine the target depths for wells and to improve the calculations of
potential reserves prior to drilling. The method can be applied to the third dimension, so 3D per-
cent sand maps can be constructed to better select well locations.
We have seen examples of exploration well locations being moved by as much as 1000 ft,
based on this analysis. If a prospect has good structural closure over a large lateral area, the spe-
cific choice of the actual well location may be based on where the prospective intervals have the
highest percent sand. This technique can be used, under the circumstances discussed, to make
such determinations. All that is required to apply the method is an array of small growth normal
faults. The usefulness of the method increases as well control becomes limited and the distance
between wells increases.
CHAPTER 12
STRIKE-SLIP FAULTS
AND
ASSOCIATED STRUCTURES
INTRODUCTION
Strike-slip displacements occur along near-vertical faults that offset basement (Harding 1990).
Displacements along strike-slip faults are predominantly in the strike direction of the fault, and
the vertical separations of the horizons along strike-slip faults may alternate between normal and
reverse separations. Along active strike-slip faults, dip-slip components of displacement are also
common (Clark et al. 1984). As they move, the crustal blocks typically encounter curves or
bends in the near-vertical fault surfaces. Material moving into these fault bends may generate
structures that can trap hydrocarbons. Commonly, areas may be subject to strike-slip, compres-
sional, and extensional displacements, and the compressional or extensional displacements need
not be contemporaneous with the strike-slip faulting (Wright 1991; Shaw and Suppe 1996). This
complex style of displacements, combined with the 3D structural development and the progres-
sive nature of the deformation, which changes with time, makes the interpretation of strike-slip
faults and their related structures difficult. These complexities often result in the misidentifica-
tion of strike-slip faults and the misinterpretation of structural styles (Harding 1985 and 1990).
Strike-slip deformation is truly a four-dimensional problem, which requires an understanding of
how the predominantly horizontal displacements occur through time. In this chapter and in
Chapter 13 we address the 4D strike-slip problem and propose methods to solve the complexities
associated with strike-slip deformation, thus providing ways to improve interpretations used to
explore for and develop hydrocarbon resources.
Problems concerning the interpretation of strike-slip faulted structures involve the issue of
recognizing empirical evidence for strike-slip faulting (Harding 1985 and 1990). According to
Harding (1990), “Many workers do not provide evidence for their assertions of strike-slip defor-
mation.” More specifically, some strike-slip interpretations lack direct evidence in support of
635
636 Chap. 12 / Strike-Slip Faults and Associated Structures
pretation, such as the presence of two faults rather than one fault or intersecting faults. Branch-
ing strike-slip faults intersect along a line of bifurcation (Bischke 2002).
In a typical exploration or development study, after mapping the faults we integrate chosen
horizons into the fault surface maps, as discussed in Chapter 8. We are unaware of any tech-
niques, other than those already discussed in this book, required to integrate strike-slip faults
into horizon maps. Thus, if a strike-slip fault exhibits normal separations, then employ the tech-
niques described in the section Techniques for Contouring Across Normal Faults in Chapter 8. If
a strike-slip fault exhibits reverse separations, then consult the discussion on reverse faults and
compressional tectonics in Chapter 8. A discussion of other geometries concerning strike-slip
faults and their associated features and styles are in the section on strike-slip faults in Chapter 8
and in the bulk of this chapter.
The Problem of Strike-Slip Fault Interpretation
Our review of strike-slip fault interpretations and the associated structures and prospects from
around the world suggests that some interpretations of strike-slip faults do not present direct evi-
dence for horizontal displacements. Often, geoscientists do not generate fault surface maps in
support of strike-slip faulting. Instead, the working group may rely too heavily on stress/strain
models or may interpret strike-slip faults to be within no-data zones on seismic sections. Some
of the observed geometries associated with strike-slip faults are inconsistent with simple models
of stress or strain (Sylvester 1988). For example, some textbooks on structural geology teach
that strike-slip faults are oriented at about a 30-deg to 45-deg angle to the direction of maximum
principal stress or strain (the compressional direction), as in Fig. 12-1. This simple model of
Figure 12-1 Simple model for compression and strain. Strike-slip faults lie at a 45-deg angle, and
reverse faults and fold axes form at about a 45-deg angle, from the contraction direction (the direc-
tion of maximum principal stress). The simple double-couple model for stress and strain assumes a
continuous (unfractured) material body. This model is inconsistent with observed geometrics associ-
ated with strike-slip faults.
638 Chap. 12 / Strike-Slip Faults and Associated Structures
strike-slip faulting assumes a continuous (unfractured), isotropic, and homogeneous body. How-
ever, faults and joints introduce discontinuities into rock that invalidate the continuous material
body assumption. The crust contains numerous fractures and is not a continuous body (Pollard
and Segall 1987). In the following section, we review evidence that this simple model of defor-
mation is inconsistent with the regional orientation and distribution of Neogene fold axes rela-
tive to strike-slip faults (Mount and Suppe 1987 and 1992).
Strain Ellipse Model
A common model used in prospect generation is one in which strike-slip displacements cause
contractional folds to form along and near strike-slip faults. Some structural geology texts teach
that these folds form along strike-slip faults according to a simple model of strain, as shown in
Fig. 12-1. The folds and the throughgoing faults are thought to form as a result of the maximum
shear strain oriented parallel to the fault surface. When applying the simple model of strain to
strike-slip faults, geoscientists may orient the model so that a shear couple parallels the direction
of the throughgoing fault (Fig. 12-1). This assumes a continuous and homogenous medium and
that the direction of maximum contractional strain or stress is at a 45-deg angle to the through-
going fault. However, rock fractures at about a 30-deg angle from the maximum principal stress
(Ramsay 1967). Some texts infer from an analysis of the simple shear model that (1) the model
accounts for the origin of folds in the vicinity of strike-slip faults, and (2) the model requires the
fold axes to trend at high angles (about 30 deg to 45 deg) to the strike-slip faults. This analysis
assumes that the plane of maximum shear stress or strain lies near, or in the plane of, the
throughgoing strike-slip fault (Fig. 12-1). However, actual folds trend at much lower angles to
strike-slip faults (Sylvester 1988), perhaps as a result of rotational displacements.
In the section on balancing strike-slip fault interpretations, we discuss how folds form adja-
cent to and along strike-slip faults. Folds adjacent to strike-slip faults may or may not exhibit
Figure 12-3 Borehole elongations measure the direction of maximum principal stress across the
Semangko Fault, Indonesia. Borehole breakout directions rotated 90 deg (short thin lines), Neogene
fold axes (dashed lines), and thrust fault earthquake focal mechanism solutions (long thin lines) are
shown. Arrows show direction of relative plate motion. Maximum compressive stress direction is
subnormal to, and Neogene fold axes are subparallel to, the Semangko fault. Neogene fold axes ex-
tend about 300 km from fault zone. (From Mount and Suppe 1992. Published by permission of the
American Geophysical Union.)
fold axes that trend at 30-deg to 45-deg to the fault surfaces. Figures 12-2, 12-3, and 12-4 show
regional Neogene fold axes relative to the San Andreas, Semangko (Indonesia), and Philippine
Faults. Notice on the figures that most of the fold axes trend subparallel to the strike-slip fault
zones. One would assume that these fold axes lie in a plane that is subnormal to the maximum
contraction direction, which in these examples would be subnormal to the surface trace of the
strike-slip faults. This orientation of the maximum contraction direction is inconsistent with the
simple strain model, and it suggests that the compression subnormal to the faults may be inde-
pendent of the strike-slip displacements. Furthermore, the Neotectonic folds are not concentrated
along the fault zones, but rather they exist as far as 50 km to 300 km from the fault zones. Thus
empirical data obtained from areas such as the San Andreas, Semangko and Philippine Faults do
not support the simple model of stress and strain, resulting in the “stress paradox” (Sylvester
1988). The simple theory, if applied to geologic or geophysical data, could affect regional and
prospect interpretations and your understanding of the petroleum system under study.
The fold geometry present on Figs. 12-2, 12-3, and 12-4 seems to be in conflict with deduc-
tive reasoning concerning continuous material behavior taught in many texts on structural geol-
ogy. However, some of these observations are consistent with or predicted by more advanced
640 Chap. 12 / Strike-Slip Faults and Associated Structures
the coordinate system. The tensor components of stress or strain interact with discontinuities to
cause the state of stress or strain within a body to be complicated. Rock mechanics textbooks pre-
sent numerous examples of complicated stress trajectory patterns, related to simple structures,
that are certainly not intuitive (e.g., Obert and Duvall 1967). Chinnery (1963) solved the problem
of the state of stress along a strike-slip fault using elastic dislocation models. Elastic dislocation
models show complicated stress trajectory patterns involving simple structures (Fig. 12-5), partic-
ularly at the ends of fractures (Bischke 1974; Xiahoan 1983). The simple model of stress or strain
cannot predict the complicated stress patterns associated with faults and fractures.
Geoscientists sometimes attempt to infer the state of stress from geologic features or struc-
tures. Inferences concerning the state of stress are complicated by several factors, including the
rotational property of the stress and strain ellipsoid. As the stress tensor may rotate through time,
the finite strain observed in rocks need not result from a unique stress direction (Flinn 1962;
Ramsay 1967). Furthermore, geoscientists must measure the state of stress; they cannot directly
observe stress. No one has ever seen a stress, certainly not a stress in the distant past. If a struc-
ture formed in the distant past, then the stresses that formed the structure may not be recover-
able. Thus, we believe that inferences, deductions, and speculations concerning stress often lead
to incorrect conclusions concerning structures and prospects. It is not our intent to discuss all the
ramifications of stress or strain or their field measurements, which are presented in texts on rock
mechanics (e.g., Jaeger 1962; Obert and Duval 1967; Jaeger and Cook 1969). Our point is that
speculations concerning stress have little value during the interpretation and prospect-generation
process, and that these speculations may cause more harm than good. Accordingly, we con-
centrate on interpretation methods that involve the displacement of stratigraphic units. This
approach has an advantage in that displaced horizons are subject to direct observation, as op-
posed to stresses that are at best measured.
Stress Measurements Across Strike-Slip Faults
When attempts were made to measure the stress on the San Andreas (California) and other large
strike-slip faults, the results were puzzling (Zoback et al. 1987; Mount and Suppe 1987 and
1992). Figures 12-2, 12-3, and 12-4 show the direction of the maximum horizontal stress trajec-
tories, as determined from borehole breakout measurements and earthquake focal mechanisms
solutions, for the San Andreas (California), Semangko (Indonesia), and Philippine Faults (Mount
and Suppe 1992). The maximum horizontal stress lies in the same plane as the maximum princi-
pal stress σ1. These measurements record the state of stress during the Neogene. The borehole
breakouts obtained from deep wells react to stresses at depth, below the influence of surface
topographic effects. Hydro-fracturing experiments, conducted during enhanced recovery efforts,
show that wellbore breakout data record the direction of the minimum principal stress σ3
(Zoback et al. 1985; Zheng et al. 1989; Dart and Swolfs 1992). The direction of maximum prin-
ciple stress σ1 is 90 deg from the σ3 direction, and lies in a plane oriented at right angles to the
borehole breakout direction. For three of the world’s major strike-slip faults, stress directions de-
rived from borehole breakout data, as shown on Figs. 12-2, 12-3, and 12-4, suggest that the max-
imum horizontal stress is predominantly subnormal to the strike-slip faults. This is in contrast to
the 30-deg to 45-deg angles as predicted by the simple model of stress. This may suggest that
major strike-slip faults are low shear stress, or weak, faults. Their fault surfaces apparently lie
near a principal plane of stress (Ramsay 1967), where the shear stress, and thus the frictional
stress, is low. These stress measurements are consistent with the lack of a heat flow anomaly on
the San Andreas Fault (Brune et al. 1969).
From this discussion, we conclude that the simple models of stress and strain are inconsis-
tent with the contractional direction implied by the regional Neotectonic fold axes and borehole
breakout directions (Figs. 12-2, 12-3, and 12-4). The simple models of stress and strain do not
include material discontinuities in the analysis, so these simple models often fail to predict the
trends and the distribution of the observed tectonic features. The fact that the simple theory fails
to predict natural features is consistent with advanced theories on discontinuous material behav-
ior and elastic deformation (Pollard and Segall 1987; Obert and Duval 1967). The basic problem
of identifying or mapping strike-slip faults is not a mechanical problem, but rather a geometric
problem. Perhaps it is time to rethink the strike-slip fault problem, particularly as strike-slip
faulting has been difficult to document when interpreting subsurface data.
Applying the appropriate model to describe the stress and strain along fault surfaces is im-
portant. Often when confronted with possible strike-slip motions, some geoscientists may model
the observed fold axes at about 30 deg to 45 deg angles to the throughgoing strike-slip fault
(Fig. 12-1). This approach, which conforms to the assumption of high shear stress, can cause
several mapping and interpretation problems. We have seen maps derived from 2D data in which
faults and structures were forced to conform to the simple strain assumption shown in Fig. 12-1.
An incorrect conclusion derived from the application of the simple model may cause several in-
terpretation problems, which are discussed below. It is not our intent to single out specific inter-
pretations, but to improve the understanding of strike-slip structures in order to help
geoscientists generate high-quality prospects. Thus, we treat strike-slip faulting as a purely geo-
metric problem that involves standard mapping techniques and methods.
One example of interpretation problems is fitting the simple strain model to nonexistent
faults, and thereby forcing these nonexistent faults into interpretations. Also, little or no consid-
eration may be given to the exploration potential of structures that exist at a distance from a
strike-slip fault. Alternatively, confusion may exist on the part of management as to why the
Criteria for Strike-Slip Faulting 643
structures do not exist at 30-deg angles to the fault zone, or why structures exist at a distance
from the fault zone. Management may assign a higher risk to the area due to these apparent
“strange structural complications.”
Geoscientists may believe that folding must have been accompanied by strike-slip faulting
in an area under study. Figure 12-1 implies that folds and strike-slip faults exist in common asso-
ciation, which occurs in many areas (Figs. 12-2 and 12-3). However, if strike-slip faults do not
exist where folds are present, which is common to many folded terrains, then geoscientists may
force strike-slip faults through data in order to satisfy a 30-deg to 45-deg angle assumption.
Nonexistent faults may be drawn downward through poor seismic data zones to converge into a
deep master basement fault, creating a thick-skinned strike-slip environment, where a thin-
skinned compressional environment and hydrocarbon migration model is the appropriate model
for the area. The interpretation of an intensely fractured structure may cause management to
abandon a good prospect (Tearpock et al. 1994).
In cases concerning en echelon folded structures, interpreters may force strike-slip faults
through coherent seismic data that constitute lateral ramps (Chapter 10), or force faults along
axial surfaces in an attempt to satisfy the strain ellipse assumption (Fig. 12-1). These practices
may result in incorrect seismic correlations, incorrect interpretations and maps, incorrect models
of faulting and the petroleum system, and in numerous disappointments and dry holes. This may
lead to further confusion concerning strike-slip faulting when evaluating the results of a drilling
program and the local petroleum system.
5. The en echelon offsets are the sites of the compressional restraining bends and the exten-
sional releasing bends, such as the rhombochasms and tipped wedge basins (Crowell 1974a
and b). These bends are recognizable on fault surface maps constructed along strike-slip
faults, where they appear as bends in the fault surface. Criteria 4 and 5 are discussed in de-
tail in the following sections.
If the preceding criteria are met within an area, then the application of a strike-slip fault
model is warranted (Harding 1990) but not proven. Across normal and reverse faults, the hang-
ing wall and footwall cutoffs of correlatable beds record the sense and the amount of the vertical
and horizontal separations. If the cutoffs are recognized, then vertical and horizontal displace-
ments are proven. However, across strike-slip faults, commonly no cutoffs exist to record the
horizontal separations along straight-line segments of strike-slip faults. Unfortunately, 2D fault
patterns or sets of divergent or convergent fault patterns, as interpreted on seismic profiles, do
not provide direct evidence for lateral displacements. These divergent or convergent fault pat-
terns provide evidence for vertical separations, but not actual horizontal displacements. Thus,
strike-slip faulting requires a 3D analysis in support of horizontal displacements. The problem
with these criteria, which are suggestive of strike-slip faulting, is the inability to directly address
the issue of horizontal displacements. Economics obligates geoscientists to present viable inter-
pretations rather than concepts “drawn on a seismic background,” as Bally (1983) correctly
recognized.
across the fault surface. Alternatively, reconstructions based on sediments deposited in a starved
basin environment will contain a displacement error commensurate with the size of the initial
offset. If the strata correlate across the fault surfaces, then these errors are likely to be small.
Faults displace distinctive stratigraphic horizons, diapirs, dike swarms, mountain ranges or
basins. If one of these features is cut by a fault, then the feature may be restorable to its initial
position. To determine the approximate slip on a fault, simply move the strata back along the
fault until the displaced feature restores to its approximate initial, or intact, position (Sylvester
1988). The feature restores back to its initial configuration at a corresponding piercing point or
line. For example, if a normal fault displaces a horizon, then the hanging wall cutoff restores
back into its corresponding footwall cutoff. A 2D seismic profile would intersect the hanging
wall and footwall cutoffs at two points. If the profile aligns in the direction of fault slip, then the
structure restores back to its corresponding piercing points.
Some features provide better piercing point evidence than others. For example, stratigraphic
pinchouts or subcrops provide better piercing point evidence than isopach or isochore maps con-
structed from syntectonic sedimentary intervals. The stratigraphic pinchout or subcrop informa-
tion represents a “line in space,” whereas an isopach map represents thickness information.
Thus, problems concerning thickness changes related to syntectonic sedimentation often arise
where the stratigraphic units change thickness across active fault surfaces. Stratigraphic thick-
ness changes can result from a variety of causes, such as growth faulting and its associated
syntectonic sedimentation, or from paleotopographic slopes that cause changes in basin configu-
ration and environments. Geoscientists have documented growth sedimentation across all known
geologic structural styles, including compressional features and strike-slip faults (see Chapter 13).
Isopach or isochore information, if based on pre-growth strata that change stratigraphic
thickness, provides good piercing point evidence. A discussion of methods for distinguishing be-
tween pre-growth and growth strata are in Chapter 13.
Other features that represent good piercing point evidence include offset zoned diapirs,
mountain ranges, volcanic belts, and basins. Zoned diapirs and basins contain walls and flanks
that represent offset surfaces where faulted. These features are useful for recognizing strike-slip
faulting if the contacts are nearly vertical and correlatable. Mountain ranges and volcanic belts
contain structures and trends that may be restorable, although mountain ranges and basins could
contain pre-existing offsets such as the salients and en echelon folds common to fold-thrust
belts. Offsets along salients and en echelon folds can be large and can introduce large errors into
reconstructions. Subsequent fault motion may occur along these potentially weak, pre-existing
offsets.
We discuss two types of piercing point evidence: first, the less definitive regional restoration
process, followed by the more definitive local and balanced restoration process.
(a) (b)
Figure 12-8 (a) - (b) Restoration of Philippine Fault, using long-wavelength features such as ophio-
lite belts, sedimentary basins, and volcanic chains. (From Bischke, del Pilar, and Suppe 1990. Pub-
lished by permission of Tectonophysics.)
648 Chap. 12 / Strike-Slip Faults and Associated Structures
vide insight into the approximate amount of horizontal slip on strike-slip faults. These approxi-
mate restorations, in support of strike-slip motion, are more convincing when supported by the
criteria discussed in the section on criteria for strike-slip faulting.
Local Restoration. Commonly, the best available evidence in support of strike-slip faulting is
the presence of the ubiquitous restraining bends and releasing bends (Crowell 1974a and b).
Restraining and releasing bends document horizontal displacements, and the restoration of these
features, using piercing point and piercing line evidence, can determine the approximate slip
along fault surfaces (Hill and Dibblee 1959).
The known strike-slip faults of the world are not perfectly straight faults, but contain small
to large en echelon offsets, called stepovers, that form restraining and releasing bends (Aydin
and Nur 1985) (Fig. 12-9). Table 12-1, taken from Aydin and Nur (1982), lists constraining and
releasing bends from various areas of the world.
Alternatively, restraining and releasing bends form at bends in strike-slip fault surfaces
(Crowell 1982). If slip along the linked strike-slip fault system moves material away from the
stepover or bend in a fault surface, then the resulting extension forms releasing bends (Fig. 12-9a).
If, however, slip along the stepover or bend moves material into the stepover or bend in the fault
surface, then the resulting compression causes restraining bends (Fig. 12-9c) (Crowell 1974a
and b; McClay and Bonora 2001).
(a) (b)
Figure 12-9 (a) - (b) Geometry of releasing and restraining bends (Aydin and Nur 1985). (a) If ma-
terial moves away from a stepover, then extension and resultant normal faults develop a releasing
bend. Structural lows exist in the area of the stepover. (b) If material moves into a stepover, then
compression and resultant reverse and thrust faults develop a restraining bend. Structural highs
exist in the area of the stepover. (From Aydin and Nur 1985. Published by permission of the Society
for Sedimentary Geology.)
Analysis of Lateral Displacements 649
Table 12-1 Size of restraining and releasing bends along strike-slip faults. (Aydin and Nur
1982. Published by permission of the American Geophysical Union.)
Releasing Bends. If motion at a stepover or bend in a strike-slip fault creates extension, then
a basin bounded by normal faults develops (Fig. 12-9a and b). Two examples of this type of mo-
tion are segments of the Hope Fault, New Zealand, and the San Jacinto Fault, California, USA
(Suppe 1985) (Fig. 12-10b and c). The San Jacinto Fault is part of the San Andreas Fault system.
Releasing bends form as material within a stepover is subjected to extension. The amount of
extension is related to the amount of slip on the strike-slip fault system. According to the releas-
ing bend model, strike-slip faults bound the basin on two parallel sides and normal faults bound
650 Chap. 12 / Strike-Slip Faults and Associated Structures
Uplift
e
ng
Ra
rse
nsve
Tra
Figure 12-10 (a) - (e) Examples of releasing and restraining bends (Suppe 1985).
the basin on the other two sides. The strike-slip faults form the walls of the basin between the
stepover, and normal faults form basin margins at the ends of the stepover (Fig. 12-9a). As the
basin extends, material slumps into the void created by the extension parallel to the direction of
strike-slip fault motion, and thus the extension records the amount of displacement that formed
the basin. To determine the approximate amount of motion that formed the bend, restore the
bend by moving the correlatable strata back in a direction that is opposite to the direction of
strike-slip displacements.
If the subsidence rate in the basin exceeds the sedimentation rate, then the syntectonic sedi-
ments deposited concurrent with strike-slip displacements may contain initial offsets across the
normal fault surfaces. Although releasing bends provide direct evidence for horizontal displace-
ments, the restoration of releasing bends may overstate the amount of the strike-slip displace-
Analysis of Lateral Displacements 651
ments, if the upthrown block does not contain growth sediments. Sequence stratigraphic evi-
dence, based on high-stand or low-stand evidence, can minimize the amount of error encoun-
tered during the restoration process. However, if the stratigraphy easily correlates across fault
surfaces, then any error in restoration is likely to be small.
The normal faults that form the margins of the basin contain hanging wall and footwall cut-
offs that form piercing lines (Fig. 12-11a and b). These piercing lines represent the hanging wall
and footwall cutoffs mapped in three dimensions. The piercing lines form as the blocks at the
(a)
(b)
(c)
Figure 12-11 Motion along a strike-slip fault can be restored by (a) mapping the hanging wall and
footwall cutoffs. (b) A profile taken parallel to the surface trace of the strike-slip fault cuts the piercing
lines, creating piercing points. (c) The piercing points are restored back to an undeformed position.
(Published by permission of R. Bischke.)
652 Chap. 12 / Strike-Slip Faults and Associated Structures
edge of the extensional basin slump into the basin subnormal to the surface trace of the strike-
slip fault. Thus, if we can assume that the direction of strike-slip motion is parallel to the sur-
face trace of the strike-slip fault (i.e., in the strike direction of the mapped fault surface), then
we can restore, or close, the basin by moving the strata in the opposite direction of fault dis-
placements and along any number of profiles that parallel the surface trace of the strike-slip
fault (Fig. 12-11b and c). The hanging wall cutoffs restore back into the footwall cutoffs at their
corresponding piercing points located along the piercing line (Fig. 12-11c).
Restraining Bends. If material moves into a stepover or bend, compression occurs in the re-
straining bend (Crowell 1974a and b; Aydin and Nur 1985). The compression forms reverse
faults, thrust faults, and pressure ridges, or pop-ups (Fig. 12-9b). McClay and Bonora (2001)
present several examples from Nevada, Wyoming, Chile, and the Netherlands. The Transverse
Ranges of California, which exist on the Great Bend, a stepover in the San Andreas Fault, are an
example of a large restraining bend (Fig. 12-10d). Contraction also occurs at restraining bends
on continuously curved strike-slip fault surfaces. We discuss an example along the Loma Prieta
bend in the San Andreas Fault in a later section in this chapter.
Let’s restore a small restraining bend to illustrate how the process confirms the presence of
strike-slip displacements, restores the initial Pliocene stratigraphic trends, and allows us to esti-
mate the amount of Pliocene displacements. This knowledge will allow interpreters to converge
on the geometry and history of the Pliocene structures and the associated petroleum system. The
bend is located on the flank of the Long Beach Anticline in the Newport-Inglewood Trend,
southern California, USA. The Newport-Inglewood Trend is a classic zone of “transpressional”
deformation (Harding 1973). The strike-slip faults along the trend form left-stepping, en eche-
lon offsets, and the Cherry Hill and Northeast Flank Faults are major faults in the Newport-
Inglewood Trend. Along the southern flank of the Long Beach Anticline, the Northeast Flank
Fault steps over to the Cherry Hill Fault (Wright 1991) (Fig. 12-12a). Trenching indicates that
the Cherry Hill Fault dies out to the southeast of the map area. Motion along the Newport-
Inglewood Trend is right-lateral and, therefore, the bend should be subject to compression. If we
assume that material enters the bend parallel to the surface trace of the Northeast Flank Fault or
the Cherry Hill Fault, then the resulting contraction is restorable. This contraction forms the Sig-
nal Hill Promontory, or pressure ridge (Fig. 12-12b), and a reverse fault that dips at about 50 deg
to the southeast (Taylor 1973) (Fig. 12-13). The method of implied fault strike (Tearpock et al.
1994), when applied to Taylor’s map of the bend in the Northeast Flank Fault (Fig. 12-12a),
shows that the fault strikes about N65E beneath Signal Hill. The reverse fault motions cause the
hanging wall beds and footwall beds to form piercing lines that strike northeast-southwest be-
neath Signal Hill.
Some textbooks seem to imply that strike-slip and compressional displacements are causi-
tively related, and that transpressional strike-slip displacements commonly generate compres-
sional folds adjacent to strike-slip faults (Fig. 12-1). The Signal Hill pressure ridge formed on
the flank of the Long Beach Anticline, but strike-slip displacements need not be the cause of the
Long Beach Anticline itself (Wright 1991). In this case, a decoupling of the strike-slip displace-
ments from the compressional displacements may be more appropriate.
Several empirical observations support a decoupling process, which could change the inter-
pretation of the Newport-Inglewood Trend. For example, according to the transpressional expla-
nation shown in Fig. 12-1, the axis of the folds would initiate at 30-deg to 45-deg angles to the
strike-slip fault. The strains required to rotate a fold axis through a 30-deg to 45-deg angle sug-
gest large amounts of strike-slip displacements. As the axis of the Long Beach Anticline is paral-
lel to the Cherry Hill Fault (Fig. 12-12a), the strike-slip displacement on the Cherry Hill Fault
should be substantial. Harding (1973) makes the observation that fold axes are presently offset
Analysis of Lateral Displacements 653
(a)
only by 200 m to 800 m, but some or most of this displacement could be an initial offset of the axes
of compartmentalized folds (Chapter 8, Fig. 8-74). The front limb of the Long Beach Anticline is
not offset from its crest (Wright 1991) (Figs. 12-13 and 12-14). A second observation is that the
structural contours in the hanging wall of the Northeast Flank Fault (Fig. 12-12a) are compa-
tible with the footwall structural contours. This structural compatibility also implies small
Plio-Pleistocene displacements (see Chapter 8). Thus, some or all of the displacements on the
Northeast Flank Fault appear to be younger than the Long Beach Anticline, which supports
Wright’s (1991) observations that the Newport-Inglewood Trend exhibits a complex structural
and stratigraphic history not easily reconciled with a simple strike-slip origin.
If you construct a fault surface map for the Northeast Flank Fault along the southeastern
flank of the Long Beach Anticline, it shows that the fault surface curves beneath Signal Hill
(Taylor 1973), where the fault strikes at about N65E (Fig. 12-12). To the southeast of the anti-
cline, the fault dips at high angles and strikes at about N60W (Fig. 12-12). A seismic or geologic
profile, such as cross section C-C′ taken in the NW-SE direction (and across the curved portion
of the fault surface), will confirm if the beds are thrust, reverse, or normally faulted. In this case
they are reverse faulted (Fig. 12-13) and, referring to Fig. 12-9, we can deduce the correct sense
of strike-slip motion. That confirms the Signal Hill restraining bend to be a small but obvious re-
straining bend.
It is obvious from this discussion that in order to locate bends in fault surfaces, it is neces-
sary to construct accurate maps of the fault surfaces, as described in Chapter 7. Typically, a 2D
seismic grid is sufficient for constructing general fault surface maps and permits the detection of
gentle bends in fault surfaces. Bends or offsets in fault surfaces have important consequences
concerning the correct interpretation of data, prospect generation, or the absence of prospects
(Tearpock et al. 1994). This is another reason for constructing loop-tied fault surface maps. As
curved fault surfaces create releasing and restraining bends along strike-slip faults, maps of these
fault surfaces may readily solve difficult structure problems.
The Signal Hill restraining bend is restorable along any number of profiles aligned subpar-
allel to the surface trace of the Cherry Hill or Northeast Flank Faults (Figs. 12-13 and 12-14).
The Northeast Flank Fault exhibits about 150 m of vertical separation and about 170 m of hori-
zontal separation in the Pliocene Lower Wilber and Alamitos horizons. These displacements are
in general agreement with Harding’s (1973) estimates. The Northeast Flank Fault cuts the
southeastern limb of a pre-existing Long Beach anticline (Figs. 12-12 and 12-14) and appears
younger than the compressional folding that formed the Long Beach anticline. The strike-slip
motion may decouple from the compressional motions that formed the Long Beach Anticline
(Shaw and Suppe 1996). Therefore, the compressional and strike-slip motions may not be di-
rectly related.
A misconception concerning piercing points is that offset fold axes, once thought to repre-
sent a continuous line, are good piercing point evidence because these offset fold axes restore to
a single line. However, fold compartmentalization caused by tear faults are known to exist
(Chapter 8, Fig. 8-74), and associated folds may form with an initial offset, and not along a sin-
gle unbroken fold axis. These initial offsets can be large. In addition, folds forming along tear
faults may grow during the deformation process, contributing to the offset. Unfortunately, we
find that most types of piercing point evidence are rare or absent from most subsurface data sets.
In summary, the releasing and restraining bends, which are common to the known strike-
slip faults throughout the world, often provide the best evidence in support of strike-slip faulting
(Aydin and Nur 1985). If you are working a suspected strike-slip fault, locate a bend to confirm
the strike-slip motions. Partially linked strike-slip fault systems typically contain stepovers.
Often, a quick look at a structure map containing a strike-slip fault interpretation can resolve the
presence or absence of strike-slip faulting in a matter of minutes. If strike-slip faults are thought
to be in the area, examine a structure map for the presence of en echelon stepovers. Next, inquire
as to the direction of fault motion. If fault motion moves material into the bend, then a structural
high should exist on the map in the area of the restraining bend (Figs. 12-9b and 12-12). If, how-
ever, fault motion moves material out of the bend, then a low should exist on the map in the area
656 Chap. 12 / Strike-Slip Faults and Associated Structures
of the releasing bend (Figs. 12-9a and 12-10b and c). An exception to this Quick Look Technique
is structural inversion, which can also produce high and low areas along en echelon inversion
faults. If a regional strike-slip fault interpretation does not contain restraining and releasing
bends, perhaps another style of faulting is present.
The documentation of strike-slip motion may be no more difficult than the documentation
of other types of fault motion, but specific technical work must be done. First, fault surface
maps, constructed in support of the deformation, should contain geometries that are consistent
with the highs and lows on horizon structure maps. Second, in other tectonic regimes, whether it
be compressional, extensional, or salt-related deformation, geoscientists and management re-
quire direct evidence of the type and style of the deformation. Strike-slip faults are not two-
dimensional, and thus they require a 3D analysis. In many cases, the answer lies in a 4D analysis
of the problem (Wright 1991). We presented techniques that can rapidly resolve the interpreta-
tion of strike-slip faulting that are no more difficult than techniques required to confirm other
types of faulting. These techniques are supported by empirical evidence that establishes 3D
structural validity.
Modern explorationists place emphasis on the petroleum system and its associated risk fac-
tors. To better understand how structures form and how faulting affects structural development
and hydrocarbon migration, a good understanding of fault timing and geometry is necessary. We
know of no way to address these issues concerning risk without fault surface maps, correctly in-
terpreted from loop-tied data. If fault surface maps do not exist, then interpreters working an area
have imprecise knowledge as to how the local structures formed and how the recognized faults
may have channeled hydrocarbons. How can geoscientists correctly identify and understand the
type and style of faulting, and generate viable prospects, without constructing viable maps based
on loop-tied data? Furthermore, these maps of the fault surfaces may contain subtle bends that in-
dicate small restraining and releasing curves, thus generating additional prospective structures.
Fault surface maps also may record the strike-slip motions, as we show in a following section.
If accurate fault surface maps are not available, then it is possible to misidentify the struc-
tural style. On 2D seismic data, one may factor flower structure fault geometries (Harding 1985)
into the risk analysis, where in reality inversion is the correct structural style. The petroleum sys-
tem model may erroneously contain thick-skinned vertical faults, whereas in reality the faults
are thin-skinned, low-angle faults. This can have a major effect on the interpretation of how hy-
drocarbons migrate and enter structures. Thus, differences in structural style can impact the un-
derstanding of the petroleum system, the interpretation model, the determination of risk, and the
ultimate success of an exploration or development program.
Strike-slip faulting is sometimes over-interpreted and confused with other structural styles
(Harding 1985), and some interpretations may lack positive evidence in support of horizontal
displacements. Too often, strike-slip fault interpretation and the related structures are supported
by a lack of data, no-seismic-data zones, high bed dips, axial surfaces, misapplied seismic inter-
pretation rules, etc. (Tearpock et al. 1994). Certainly, strike-slip faulting should be subject to the
same scientific methods and subsurface mapping standards that apply to other styles of faulting.
Your success as a geoscientist depends on high-quality interpretations, maps, and prospects.
Only solid scientific work, as described here, can ensure the best chance of success.
of strike-slip fault bends. Their study has implications regarding the development of strike-slip
faults that may not be totally understood. Aydin and Nur investigated 11 major strike-slip faults
in various areas around the world, and they measured the length and the width of 70 restraining
and releasing bends associated with those faults (Table 12-1). The width of bends was not inde-
pendent of the length of bends, as one would assume from the model shown in Fig. 12-15. In-
stead, the lengths (L) of bends are proportional to the widths (W) by the relation
L ≅ 3.2 W (12-1)
Equation 12-1 is at the 95 percent confidence level. The formula implies that the widths of
bends widen as the slip on the faults increases. This implies that smaller bends may coalesce and
cluster into larger bends (Aydin and Nur 1982), or that fault zones grow wider as the faults grow
in length.
The consequences of their study are multifold. Strike-slip bend widths can be used to esti-
mate the lengths of bends. Large strike-slip faults have wide bends (Fig. 12-10d), whereas small
strike-slip faults have narrow bends. It is possible, however, that a large strike-slip fault may de-
activate and a new fault may replace a major fault, which would start the process over again. As
large strike-slip faults have wide bends, these wide bends should be easy to locate during a re-
gional analysis, and help to confirm the strike-slip interpretation. Narrow bends on small strike-
slip faults may be more difficult to locate. In this case, however, the strike-slip process is less
important and other processes may dominate, such as extensional or compressional faulting or
folding.
We know of no well-documented example by which large amplitude anticlines, domes,
structural highs, or basins form as a result of small amounts of strike-slip motion. If the amount
of strike-slip deformation is small, the exploration emphasis must shift from processes associ-
ated with strike-slip faulting to processes associated with other types of faulting. Another conse-
quence of the study of releasing and restraining bends is that the bends are relatively common
along strike-slip faults (Table 12-1). Lastly, as releasing bends grow wider with increasing slip,
smaller basins that form within each bend may widen to form larger basins.
Figure 12-15 A simple model of a strike-slip fault that has a releasing bend that grows in length as
the slip on the fault increases. According to this model, the width (w) of the bend should be indepen-
dent of the length (l) of the bend. Strike-slip faults do not behave according to this simple model, and
real strike-slip faults exhibit a relationship in which length is proportional to width. (From Aydin and
Nur 1982. Published by permission of the American Geophysical Union.)
658 Chap. 12 / Strike-Slip Faults and Associated Structures
Figure 12-17 Geologic map of Loma Prieta epicentral area showing Glenwood syncline emanat-
ing from bend in surface trace of San Andreas Fault. Balanced cross section B-B′ is shown in
Fig. 12-19. San Andreas Fault turns from 320 deg to 310 deg, forming a restraining bend. (From
Shaw, Bischke, and Suppe 1994, United States Geological Survey Publication.)
restraining bend. Other strike-slip faults should exhibit similar geometries, but we have not been
able to locate public domain fault surface maps constructed from seismic or other data that are
published, in order to evaluate their geometries.
Before entering the bend, the Pacific Plate moves in the N40W (320 deg) direction parallel
to the surface trace of the San Andreas Fault. As material enters the bend to the north of Wat-
sonville (Fig. 12-17), the Pacific Plate moves up the ramp formed by the southwesterly dipping,
high-angle reverse-strike-slip fault surface (Fig. 12-18a). The upward motion generates com-
pressional synclines, similar to synclines that form at the base of compressional thrust fault
ramps (Chapter 10). In this example, the syncline should emanate from where the fault surface
bends and departs from its general N40W (320 deg) trend. In the south, hypocentral solutions
define a San Andreas Fault that dips at 82 deg and strikes N40W (320 deg) (Profile 3 of Fig.
12-18b). In the restraining bend and at Profile 2 of Fig. 12-18b, the fault changes in dip to 70 deg
at 8 km depth and strikes at N50W (310 deg). As predicted, material entering this bend in the
fault surface generates the Glenwood syncline (Fig. 12-17). This syncline terminates at the
southern bend in the fault surface, northeast of the Rapp Well, where the San Andreas Fault de-
parts from its general N40W (320 deg) trend (Fig. 12-17). Ground-surface dip data present on
the surface geologic map and well log data constrain the geometry of the Glenwood Syncline
(Figs. 12-17 and 12-19).
660 Chap. 12 / Strike-Slip Faults and Associated Structures
(a)
(b)
Figure 12-18 (a) Fault surface map for Loma Prieta restraining bend showing locations of cross
sections 1, 2, and 3. (b) Cross sections of San Andreas Fault as defined by hypocentral activity. The
fault bends at cross section 2 at a depth of 8 km and it dips at a higher angle at cross section 3.
(From Shaw, Bischke, and Suppe 1994.)
In the brittle regime of the earth’s crust, folds form as hanging wall beds move over nonpla-
nar fault surfaces (Bally et al. 1966; Suppe 1983). Thus, we can use surface and well data related
to the Glenwood Syncline to generate balanced models of strike-slip compressional folding
along the San Andreas Fault. Cross section balancing allows us to predict the subsurface fault
geometry from the surface and well data. In practice, this exercise allows geoscientists to predict
the vertical and lateral dimensions of the hanging wall geometry (Tearpock et al. 1994). Hanging
wall geometry is important in positioning wells, in understanding the size of a prospect, and
Balancing Strike-Slip Faults 661
in generating admissible interpretations of strike-slip faults (see the Generic Example in this
section). We use the balanced model to predict subsurface fault geometry and compare the
model predictions against the observed fault surface geometry as defined by hypocentral earth-
quake activity.
Sometimes fault surfaces image on seismic data sets, but the hanging wall structure does
not clearly image. Balancing techniques can predict fold shape from fault shape (Tearpock et
al. 1994) and help constrain interpretations. We balance profile B-B′ on Fig. 12-17 (profile 2 in
Fig. 12-18b) and present it as Fig. 12-19. One of two approaches can be employed in order to
balance a profile. If the subsurface fault geometry is known from depth-corrected seismic sec-
tions and well control or, in this case, hypocentral aftershock activity, then geoscientists can
generate a balanced, or generic, model of the hanging wall structure. Alternatively, if the shal-
low hanging wall geometry variables γ and β are known from depth-corrected seismic sections,
outcrop data, or well control, then geoscientists can estimate the fault surface geometry vari-
ables θ and φ (Fig. 12-19), where
662 Chap. 12 / Strike-Slip Faults and Associated Structures
In this case, we use surface dip data and shallow well control to determine γ, the angle be-
tween the axial surface and the adjacent beds of the Glenwood Syncline (Fig. 12-19). The hang-
ing wall cutoff angle β, in Fig. 12-19, can be determined from local bed dips and fault geometry.
By definition,
(β) = dip of shallow fault – dip of beds adjacent to fault surface (12-2)
The flank of the Glenwood Syncline adjacent to the fault dips at an average of 35 deg to the
southwest (Figs. 12-17 and 12-19), whereas regional dip is 8 deg southwest in the area west of
the synclinal axis. Well log and surface dip data from along the trend of the Glenwood Syncline
constrain the regional dip. Thus, to determine β, subtract the dip of the flank of the Glenwood
Syncline from the observed surface dip of the San Andreas Fault within the restraining bend. The
surface dip of the fault is about 80 deg (Brabb 1989), and thus β = 80 deg – 35 deg = 45 deg.
From inspection of Fig. 12-19, the axial surface angle γ can be determined from the Kink Law
(Chapter 9), or from the following equation.
2 γ = 180 deg – dip of synclinal limb adjacent to fault + regional dip
Therefore,
γ = (180 deg + regional dip – dip of synclinal limb adjacent to fault)/2 (12-3)
γ = (180 deg + 8 deg – 35 deg)/2
γ = 77 deg
Dip data at the ground surface is used to position the axial surface (Fig. 12-19). The axial
surface is drawn downward at an angle of 69 deg (77 deg – 8 deg) until it reaches the fault. This
point determines the position of the bend in the fault. Horizons on the synclinal limbs can now
be constructed, with the fold hinge bisected by the axial surface.
We next consult the fault-bend fold graph (Fig. 10-34) to generate a balanced solution to the
fault surface problem. The correct graph to use is the diagram for synclines (right graph). The
values required are γ = 77 deg, recorded on the y-axis of the graph, and β = 45 deg. The values
for β are recorded by the right-sloping bold diagonal lines. Project a horizontal line into the
graph from the value of γ = 77 deg. Where the γ = 77 deg line intersects the β = 45 deg curve, the
change in fault dip angle φ is read off the graph from the thin, near-horizontal and downward-
sloping curve on the graph. The value of φ is slightly less than 10 deg.
Therefore we conclude that the Glenwood Syncline is created by a 10-deg subsurface bend
in the San Andreas Fault. The axial surface of the Glenwood Syncline emanates from this sub-
surface bend in the San Andreas Fault, where the previously undeformed beds moved over the
bend in the fault surface (Fig. 12-19). As described above, the depth to the bend in the fault sur-
face was determined by projecting the axial surface of the Glenwood Syncline downward to
where it intersects the 80-deg dipping fault. The axial surface intersects the San Andreas fault at
a depth of about 7 km (Fig. 12-19). Below this depth, the fault takes a 10-deg bend and the bal-
anced model predicts a fault dip of about 70 deg below 7 km (Fig. 12-19).
Balancing Strike-Slip Faults 663
We can now compare the model-generated values and the predicted depth of the fault bend
to profile 2 in Fig. 12-18b. The strike-slip fault bend fold model predicts that the San Andreas
Fault makes about a 10-deg angle bend at a depth of about 7 km and that the fault dips at about
70 deg below 7 km (Fig. 12-19). These values compare favorably with the hypocentral data that
suggests a 10-deg fault bend at a depth of 8 km. Below this depth, the fault dips at about 70 deg,
as shown in Fig. 12-18b, Profile 2. Given the good agreement between theory and observation,
strike-slip fault bend fold theory may allow geoscientists to make viable predictions of the struc-
tural geometry along strike-slip faults. Balanced cross sections of compartmentalized displace-
ments along strike-slip faults may help geoscientists generate higher quality prospects in a
tectonic environment that has proven to be difficult to quantify. When millions of dollars are at
stake, deterministic models may add value to conceptual interpretations of prospects generated
in strike-slip faulted tectonic regimes.
The San Andreas Fault is subject to hundreds of kilometers of slip, which created the Glen-
wood Syncline with a broad south-dipping limb (Fig. 12-17). This limb could form a three-way
closure against the San Andreas Fault. Other strike-slip faults that contain smaller amounts of
slip should succumb to an analysis similar to what we have presented here. Balanced solutions
of these structures can improve prospect viability, perhaps solve complex structural problems,
and thereby reduce prospect risk.
Generic Example of Strike-Slip Compressional Folding. If fault surface maps are ob-
tainable from well log or seismic data, then balanced models of hanging wall geometries are
possible. Balanced models of hanging wall geometries lead to viable, low-risk prospects. A
generic example is shown in the cross section presented in Fig. 12-20. Perhaps your working
group was subject to the following type of problem that employed well log and seismic data. Let
us assume that two companies find hydrocarbons in Wells No. 2 and 3, in the A, B, and C Hori-
zons adjacent to a vertical fault. Outcrop and seismic data suggest that the near-vertical fault,
containing restraining and releasing bends, is located between Wells No. 1 and 2 (Fig. 12-20).
The two companies construct cross sections of the new field in order to propose additional devel-
opment wells. In Fig. 12-20, well log data constrain the fault geometry, but what is the limit of
the field, and is it necessary to drill additional wells?
We consider two interpretations of the data: a qualitative interpretation presented by Com-
pany A and a quantitative interpretation presented by Company B. We discuss the Company A
interpretation first. We make the assumption that geoscientists from Company A do not have a
solid background in structural geology and therefore have not been trained in volume conserva-
tion or structural balancing concepts. They construct a cross section through the field that em-
ploys conceptual, but not volume, conservation concepts. On the other hand, geoscientists from
Company B construct a balanced cross section of the existing data. How may these two working
groups and their interpretations differ? Can the difference affect future exploration and success?
We will assume that the 3D seismic data that crosses the area is of reasonable quality but suffers
from the usual problems, such as surface statics and the inability to image steeply dipping beds.
After examining the data present in Fig. 12-20, the Company A geoscientist interprets a sec-
ondary fault along the western flank of the structure. Two independent sources of evidence exist
for this fault. The first piece of evidence is the no-data seismic zone on the flank of the structure
(Fig. 12-20). The second piece of evidence is the change in thickness of the stratigraphic units
above the D Horizon, between Well No. 2 and Well No. 3. The interpretation is shown in Fig.
12-21. The proposed secondary fault exhibits normal, strike-slip, and reverse separations, and it ex-
plains the bed dip and thickness variations between the B and C Horizons and the C and D Hori-
zons in Wells No. 2 and 3. This fault geometry also accounts for the slip reversal between the B and
D Horizons. Strike-slip faults can exhibit normal, reverse, and lateral separations. The interpreted
fault turns down and merges with the large, master strike-slip fault at depth.
Lastly, Company A completes the seismic interpretation by correlating the reflections from
the crest of structure into the off-structure flank position located to the west of the no-data zone
on the seismic data set. In the low dip area to the west of the no-data zone, the seismic images a
gently folded structure in which the reflectors turn up beneath the secondary fault (Fig. 12-21).
The turned-up beds are interpreted to be caused by fault drag on the secondary fault, as is shown
in many structural geology textbooks (Billings 1972).
The Company A team present the cross section of the folded and faulted structure shown in
Fig. 12-21. Using cross sections and maps of the field, they propose two additional wells to de-
fine the western limits of the field. These wells will test the upturned A and B Horizons located
beneath the fault.
The company A interpretation makes three basic assumptions: first, that the no-data zone on
the seismic data set results from faulting; second, that faulting causes changes in thickness of
stratigraphic intervals as seen in well logs, perhaps associated with repeated or missing section;
third, that folds tend to be gently curved, rounded features, as shown in some textbooks on struc-
tural-metamorphic geology (see Chapter 10). Does another interpretation of the data exist that
applies to the low temperature portions of mobile belts?
The Company B geoscientist team, having knowledge of structural geology concepts and
techniques, attempts to balance the data shown in Fig. 12-20 using procedures developed in Chap-
ter 10 and in this chapter. They notice that the D Horizon is on the same structural level in Wells
No. 2 and 3, and thus this horizon is not subject to possible faulting or folding between these two
wells. Procedures outlined in the section on compressional folding along strike-slip faults suggest
that the D Horizon may be near a bend in the vertical fault. Using the fault cuts located below the
Balancing Strike-Slip Faults 665
Ver tical
Strike-
slip
Fault
Figure 12-21 Company A geoscientists solve the folded-structure problem by using the concept of
a secondary fault, related to strike-slip displacements, to explain the apparently high bed dips and
vertical thickness changes interpreted from the wells. They propose two additional wells to test the A
and B Horizons. (Published by permission of R. Bischke.)
D Horizon in Wells No. 2 and 3, the Company B geoscience team reasons that the cutoff angle (θ)
between the D Horizon and the deeper part of the fault is 60 deg (Fig. 12-22). Thus, the difference
(φ) in dip angle of the fault is 30 deg.
The geoscientists can now complete the cross section shown in Fig. 12-22, using the meth-
ods outlined in the previous section on compressional folding along strike-slip faults. Strata are
deformed above the bend in the fault surface, so an axial surface emanates from this fault bend.
They determine the dip of the axial surface by using the method described in the preceding sec-
tion. Knowing that θ = 60 deg and φ = 30 deg, they use Fig. 11-34 (right graph) to determine the
axial surface angle γ to be 60 deg. As the regional dip is 0 deg, the actual dip of the axial surface
is 60 deg. The axial surface projects upward at 60 deg from the bend in the fault (Fig. 12-22).
The axial surface bisects the hinges of a fold, so by construction the A, B, and C Horizons dip at
60 deg through the no-seismic-data zone. The axial surface correlates to the western limit of the
no-data zone on the seismic, and thus the no-data zone can result from changing bed dips and
not from faulting, as assumed in the first interpretation.
Using similar reasoning with respect to Horizons A and B, the up-dip limit of the no-data
zone can define a second axial surface. The axial surfaces are positioned as shown in Fig. 12-22.
As described earlier, seismic character correlation suggests the structural position of Horizons A
and B are to the west (Fig. 12-20).
666 Chap. 12 / Strike-Slip Faults and Associated Structures
Figure 12-22 Company B geoscientists attempt to balance the structure and generate a strike-slip
fault bend fold model that explains the anomalous bed dips and interval thickness changes at and
below horizon C. In this model, the high bed dips occur at axial surfaces that define the flank of a
monocline. Geoscientists propose a single shallow well at the A Sand to define the western limits of
the field. This model shares properties with box and lift-off structures. (Published by permission of
R. Bischke.)
The Company B team assumes that the structure can be modeled using the Kink Method de-
scribed in Chapter 10, and thus their cross section has more angular folds than the Company A
cross section shown in Fig. 12-21. Petroleum-scale structures, exhibiting a kink fold structural
style, are common to the low temperature portions of fold and thrust belts (Suppe 1985; Boyer
1986). Furthermore, the Company B team are suspicious of the upturned beds imaged below the
no-data seismic zone. The folding elevates high density and high velocity strata to a structural
level that is higher than the equivalent beds located to the west of the no-seismic-data zone.
Thus, the area below the region of high bed dips may be exhibiting velocity pull-up. In addition,
as Company B concluded that the no-data zones results from folding and not faulting, there is no
reason to postulate a fault-drag effect. They model the beds to the west of the no-data zone to be
nearly flat surfaces.
The balanced interpretation presented by Company B also makes three assumptions. First,
the no-data zone on the seismic data set results from high bed dips caused by folding. Second,
folding causes changes in bed dips as well as in the different well log thicknesses. Third, folds in
the low-temperature portions of mobile belts commonly exhibit kink geometry. The interpreta-
tion presented in Fig. 12-22 has properties similar to lift-off or box folds (Chapter 10). The two
Balancing Strike-Slip Faults 667
dip domain analysis (Chapter 10) typically locates direct evidence for the faulting in the
form of bed dip discontinuities. If another explanation is as probable, or more probable,
then caution is prudent. Risk the uncertainty accordingly. Good scientific work should yield
better interpretations, resulting in more success at lower costs.
A structure similar to Fig. 12-22 is shown in Fig. 12-23 from Pecos Country, New
Mexico, USA (Kelley 1971). The geologic map of the region shows several long, arcuate
northeast-southwest trending faults attributed to strike-slip displacements. This structure,
called the Y-O Buckle, is bound by a near-vertical fault that contains a bend. A monoclinal
fold forms above the bend in the fault, and below the bend the beds are not folded. The
folded structures along these faults are subparallel to the surface trace of the faults, and
they tend to exist at gentle bends in the faults. Thus, some of the folds appear to form at re-
straining bends. Field mapping demonstrates that the faults contain vertical separations that
commonly change from normal to reverse separations along strike (Kelley 1971). The
strata are deformed adjacent to the fault surfaces, often as disharmonic folds with near-ver-
tical limbs (Figs. 12-22 and 12-23).
Figure 12-23 Box-like detachment fold in Pecos County, New Mexico, USA along a long, arcuate
fault system that appears to contain restraining and releasing bends. In the photograph, the strata
are folded above the bend in the fault surface, similar to Fig. 12-22. (From Kelly 1971. Published by
permission of the New Mexico Bureau of Mines and Mineral Resources.)
Balancing Strike-Slip Faults 669
Basin, southern California, that was carefully mapped by Crowell and his colleagues and stu-
dents (Crowell 1950, 1982, 2002a, b, c). Crowell (1974a and b) formulates the concept of re-
straining and releasing bends along strike-slip faults. Bischke (2002), Crowell (2002c), and Link
(2002) describe the detailed structure and stratigraphy of the basin along with the geologic
complications.
Ridge Basin Geology. Ridge Basin formed along the northeastern boundary of the large San
Gabriel strike-slip fault (Fig. 12-24), which was active during the late Miocene. The San Gabriel
Fault is an inclined fault surface that strikes at about N40W and dips 50 deg to 70 deg to the
northeast (Crowell 2002c; Bischke 2002). Piercing point evidence, consisting of offset belts of
pre-Late Miocene rocks and structures, suggests that the San Gabriel Fault experienced about 60
km of strike-slip displacement (Crowell 1982, 2002c). A seismic line presented by May et al.
(1993) trends across the northwestern portions of Ridge Basin (Figs. 12-24 and 12-25). This line
appears to image a fault that branches (splays) off the San Gabriel Fault (Bischke 2002), and we
call this proposed branch fault the Hungry Valley Fault. Most importantly, the line may image a
rollover structure that forms downthrown to a fault that exhibits normal separations (Fig. 12-25).
Typically, rollover structures dip toward listric normal faults at about right angles, and therefore
the rolled-over beds may dip toward a north-striking, branch normal fault that splays off the San
Gabriel Fault (Bischke 2002; Link 2002).
Crowell (1974a and b) proposed that Ridge Basin formed as a result of a restraining bend
located to the northwest of Ridge Basin and beneath the Frazier Mountain Thrust (Fig. 12-24).
This bend would uplift rocks located to the northwest, subjecting the rocks to the southeast of
the bend to extension and providing a source for Ridge Basin sediments. This structural configu-
ration favors a releasing bend, subjecting Ridge Basin to a component of both dip-slip and
strike-slip deformation. Furthermore, the Hungry Valley Fault imaged in Fig. 12-25 has normal
separations that create accommodation space for the accumulation of the Ridge Basin sediments.
This depocenter beneath Hungry Valley trends about north-south (Link 2002), so the Hungry
Valley Fault also trends at about N0E. The Hungry Valley Fault was apparently responsible for
the accumulation, rotation and shingling of a section of Ridge Basin syntectonic sediments that
are 14 km thick. The basin, however, was only about 5 km to 6 km deep (Crowell 2002c; Link
2002), and some deformational process was responsible for the rotation of these strata.
A geologic map of the basin constructed by Crowell (1982, 2002a, b, c) shows the follow-
ing major features (Fig. 12-24). The geology adjacent to the San Gabriel Fault is dominated by
the syntectonic Violin Breccia, deposited throughout the Late Miocene (Crowell 1982). Paral-
leling the breccia is Ridge Basin syncline. The close association of the Violin Breccia with the
Ridge Basin syncline suggests that the two features are related and that Ridge Basin syncline is
a syntectonic feature. As Ridge Basin Group sediments onlap the flanks of the syncline, the
stratigraphy supports this interpretation that the syncline is syntectonic (Link 2002). In the
vicinity of Hungry Valley, the Pliocene sediments dip at gentle angles of 5 deg to 10 deg. These
gently warped sediments were subject to late-stage Pliocene to Holocene folding that post-
dated the structure in areas north and northwest of dip domain 1 (May et al. 1993). Thus, the
area surrounding Hungry Valley represents a gently dipping domain that was subsequently
folded (Fig. 12-24).
Surface bed dips obtained by Crowell define two large dip domains that dominate the struc-
ture to the southeast of Hungry Valley (dip domains 1 and 2 in Fig. 12-24). The Ridge Basin syn-
cline partitions these domains. The largest of the domains (domain 1) dips to the west and lies
about 1 to 3 km to the northeast of the San Gabriel Fault (Crowell 1982, 2002c; Wood 1981).
The beds in this domain contain average dips of about 25W. In map view, the strike of the beds
varies from 45 deg to 90 deg relative to the surface trace of the San Gabriel Fault. We interpret
670
Chap. 12 / Strike-Slip Faults and Associated Structures
Figure 12-24 Generalized geologic map of Ridge Basin, California, modified after Crowell (1982). Dip domain 1, located 1 to 3 km from the San
Gabriel Fault, strikes north-south and dips to the west. Dip domain 2, adjacent to the San Gabriel Fault, dips to the north and is syncline-
separated from dip domain 1. The two domains form the Ridge Basin syncline that subparallels the San Gabriel Fault. The gentle dips beneath
Hungry Valley (in the west) are slightly folded by later compressional forces, suggesting that this region may represent the frontal, undeformed
portions of a rollover structure (Fig. 12-25). Posted on the figure are the average values of the bed dips in the individual subdomains, along with
the dip and strike directions. Average bed dip and strike in domain 1 are 25W and N10E, and plunge of the synclinal axis is 24NW. (Published by
permission of R. Bischke.)
Balancing Strike-Slip Faults 671
Figure 12-25 Tracing of depth-corrected seismic line across Ridge Basin that images an apparent
half-graben rollover structure. The oblique seismic profile images reflections that contain apparent
dips to the NW. Where the profile crosses the low dip area of Hungry Valley, the hanging wall reflec-
tions directly above and adjacent to the fault surface exhibit low apparent dips. Seismic line crosses
into domain 1 at about sp 250. See Fig. 12-24 for the line location. (Modified after May et al. 1993.
Published by permission of the Geological Society of America.)
this domain to dip toward the Hungry Valley Fault. The average bed dips in smaller subdomains
are shown on Fig. 12-24.
Dip domain 2 lies adjacent to the San Gabriel Fault and is syncline-separated from do-
main 1 (Fig. 12-24). In the northwest, this domain extends about 3 km to the northeast of the San
Gabriel Fault, but to the southeast of Fisher Spring, the domain narrows to within 1 km of the
fault zone. In domain 2 and in map view, the beds generally dip at high angles to the north and
uniformly dip away from the San Gabriel Fault at oblique angles to the fault zone. Bed dips in
this domain average 34 deg to the north. On typical rollover structures, from different areas of the
world, beds dip toward the normal fault at about right angles, and not away from the fault surface
at oblique angles. Thus, dip domains 1 and 2 form the Ridge Basin Syncline. The oblique seismic
line extends near dip domain 2, located adjacent to the San Gabriel Fault, before entering dip do-
main 1, located about 5 km from the fault zone along the seismic line (Fig. 12-25).
Any model of the gross structural features in Ridge Basin should qualitatively and quantita-
tively explain the following structural features: (1) the approximate bed dips and the direction of
the dips within the two domains, (2) the close association of the Ridge Basin syncline with the
syntectonic Violin Breccia and the San Gabriel Fault, (3) the processes that cause the bed dips to
rotate up to 90 deg near the San Gabriel Fault zone, forming the Ridge Basin synclinal kink fold,
and how these processes operate, and (4) how Ridge Basin was filled with 14 km of sediments
that dip at an average of 25 deg to the west. In addition, the model must be compatible with the
stratigraphy of Ridge Basin (Link 2002) and with Goguel’s Law of volume conservation (Chap-
ter 10). The dip domains and the syncline mapped by Crowell provide insight into the major
672 Chap. 12 / Strike-Slip Faults and Associated Structures
tectonic processes associated with the San Gabriel Fault and the formation of Ridge Basin. As
the subsurface geology is not well-constrained, we consider a solution to the Ridge Basin geol-
ogy that assumes an extensional origin. Other explanations may be possible, especially if more
data are acquired.
Geometry of Strike-Slip Extensional Folding. The deformation of the hanging wall beds
along a releasing bend differs from the deformation along a normal fault, as is seen in the per-
spective view diagram presented in Fig. 12-26. An inclined strike-slip fault and a branch normal
fault are shown. The dip of the branch normal fault surface decreases with depth. For purposes of
convenience only, in Fig. 12-26 we show this deep fault surface to be virtually horizontal. Dis-
placement of the hanging wall block over the footwall causes the beds to be downthrown to the
branch normal fault. These displacements also create accommodation space for growth sedi-
ments. According to the model, the beds adjacent to the strike-slip fault rotate downward at high
angles. Beds located above the deep branch fault surface rotate toward the branch normal fault.
The process generates a synclinally folded structure downthrown to the branch normal fault,
with a synclinal axis that subparallels the strike of the strike-slip fault.
The detailed kinematics of the process are best shown in Fig. 12-27, which is drawn such
that the near left face of the block is in the same vertical plane as the synclinal axis GH in Fig.
12-26. In Fig. 12-27a, an inclined strike-slip fault surface, ABCD, forms a right-stepping releas-
Figure 12-26 Perspective view of a synclinal, rollover structure forming along an inclined strike-
slip fault. The strike-slip and branch normal faults form a releasing bend, and the strike-slip fault may
be throughgoing. Slip vector is parallel to the strike direction of the strike-slip fault surface. The dip
domain panel adjacent to the strike-slip fault dips away from the strike-slip fault, as the domain situ-
ated across the synclinal fold axis dips toward the branch fault surface. This fold style occurs along
the San Gabriel Fault as the Ridge Basin Syncline (Fig. 12-24). Synclines may emanate from in-
clined, releasing bends subject to large components of strike-slip motion. Figure is not drawn to
scale. (Published by permission of R. Bischke.)
Balancing Strike-Slip Faults 673
(a)
(b)
Figure 12-27 The kinematics of rollover folding above a releasing bend in a strike-slip fault. Slip
component SH parallels the strike of the fault surface. (a) Slip from point A to point B causes the
hanging wall cutoff BC to separate from footwall cutoff AD. The resultant motion opens a void be-
neath surface EBC, and the hanging wall will collapse by Coulomb shear onto the footwall. Point D′
collapses onto point D along collapse angle ψa (b) The resultant deformation causes the beds to dip
away from the strike-slip fault surface in a direction perpendicular to the surface trace of the
Coulomb collapse surface BGC. Stereographic projection methods can determine the plunge of
the line of bifurcation AD, based on knowledge of the strike and dip of the beds adjacent to the
strike-slip fault and the collapse angle ψa. Figure is not drawn to scale. (Published by permission of
R. Bischke.)
ing bend fault surface ADF that exhibits normal separations. The hanging wall block slips a
small distance SH over the inclined strike-slip fault from point A to B, parallel to the strike of the
strike-slip fault. This motion opens a small void beneath the hanging wall, a void which extends
to the hidden, dashed line BC. The line BC was originally coincident with the line of bifurcation,
AD, of the two fault surfaces (Fig. 12-27b). So this void exists above the surface ABCDF. As fault
ADF is a normal fault subject to oblique-slip, this void is instantaneously filled by collapse (along
Coulomb failure surface ψ) of the hanging wall material onto the surface ABCDF (Chapter 11).
674 Chap. 12 / Strike-Slip Faults and Associated Structures
In Figs. 12-26 and 12-27b, line GH defines the synclinal axial surface. Volume conservation
therefore dictates that point D′ collapses onto the footwall in the plane containing line GH. As
points D and D′ lie in the plane defined by the apparent Coulomb collapse angle ψa, point D′
collapses onto point D, forming a fold (Bischke 2002). The collapse of the hanging wall onto the
footwall causes the hanging wall beds to rotate toward point E in Fig. 12-27b, by bending along
axial surface trace line BG. Axial surface trace line BG is the surface expression of the inactive
Coulomb collapse surface, BGC. The strike of the beds adjacent to the strike-slip fault define
this deformation, which in the case of Ridge Basin is domain 2 in Fig. 12-24. This deformation
causes the beds to rotate away from the inclined strike-slip fault surface, forming one-half of a
synclinal fold.
The remaining half of the synclinal fold is shown in Fig. 12-26 as the surface above the fault
flat, dipping toward the branch normal fault. For simplicity, that surface was not shown in Fig.
12-27. Collapse along Coulomb shearing surfaces causes that surface to dip toward the branch
fault. In the case of Ridge Basin, this surface is in dip domain 1.
If a strike-slip fault and branch fault form a fault flat as shown in Fig. 12-26, then two dip
domains form as the hanging wall block slides over the footwall. The beds adjacent to the strike-
slip fault dip away from the strike-slip fault and the beds above the fault flat dip toward the
branch normal fault. This deformation results in a synclinal structure with an axial surface that
trends subparallel to the surface trace of the strike-slip fault (Fig. 12-26). The Ridge Basin Syn-
cline exhibits this type geometry (Fig. 12-24).
Geometry of the Hungry Valley Fault. After determining the strike and dip of the Hungry
Valley Fault, which branches off the San Gabriel Fault, we model the dip of the deep normal
fault beneath Ridge Basin. The modeling uses stereographic projection methods and extensional
inverse techniques described in Chapter 11. As stated previously, near Hungry Valley the San
Gabriel Fault strikes at about N40W and dips to the northeast (Crowell 2002c; Bischke 2002)
(Fig. 12-24). In addition, the trend of the Hungry Valley depocenter of about N0E, and the N10E
strike of the beds in domain 1, constrain the strike of the Hungry Valley branch normal fault.
For the case of pure strike-slip faulting, any vertical profile, oriented parallel to the surface
trace of the master strike-slip fault, contains a zero-deg (αa = 0.0) apparent dip for the fault
surface. This apparent dip is independent of the dip on the strike-slip fault surface. This means
that a general solution to the problem of strike-slip extensional folding is independent of the dip
on the San Gabriel Fault surface (Bischke 2002). To determine the dip of the Hungry Valley
Branch Fault, model-based variables include (1) the strike of the San Gabriel Fault trending at
N40W, (2) the strike of the branch fault, and (3) the Coulomb collapse angle that typically varies
from 60 deg to 70 deg (Xiao and Suppe 1992; Bischke and Tearpock 1999). Here we assume a
65-deg Coulomb collapse angle.
If we assume that the shallow portions of the proposed Hungry Valley Fault parallel the
strike direction of dip domain 1, then the Hungry Valley Branch Fault may subparallel the pre-
sent trend of Hungry Valley, or N10E. In the model, the branch fault forms the western flank of
the Hungry Valley depocenter. This assumption seems reasonable, as the stratigraphy of the
basin (Link and Osborne 1982; Link 2002) predicts that a depocenter exists in the area of Hun-
gry Valley (Figs. 12-24 and 12-25). In addition, the direction of sedimentary transport into the
Hungry Valley depocenter is from the north (Link 2002). Using the observation that dipping beds
related to normal faulting dip at about a right angle to the strike of the normal fault (typical
of rollover structures), a specific solution to the structural problem is shown in Figs. 12-28 and
12-29. In this model, the beds strike at N10W and dip toward the Hungry Valley Fault. Inverse
modeling, to be discussed next, and the plunge angle of the Ridge Basin syncline of θa = 24 deg
toward N40W, constrain this solution. In this solution the Hungry Valley Fault trends at N10E
Balancing Strike-Slip Faults 675
Figure 12-28 Stereoplot showing interpreted surface and subsurface structure at the Hungry
Valley Branch Point (Fig. 12-30). Fault surfaces are bold lines and dip domains are thin lines on the
projection. Constraints on the interpretation include the structural cross sections of Crowell (2002c),
a seismic line, the surface trace of the San Gabriel Fault that trends at N40W, the average dip and
strike of beds in dip domains adjacent to Hungry Valley, the N10W trend of the Hungry Valley de-
pocenter (Link 2002), and the Sun Schmidt Well. The plunge of dip domain 2 and the Ridge Basin
syncline (θa = 24 deg at N40W) help constrain the apparent dip (βa = 55 deg at N40W) of the Hun-
gry Valley Fault that strikes at N10E. The Hungry Valley Fault flattens at depth to form a listric fault
surface. Dip domain 1 (25W/N10E) and dip domain 2 (36NE/N75W) intersect to form the Ridge
Basin syncline that trends subparallel to the San Gabriel Fault. Apparent bed dip of (θa = 24NW)
constrains bend in the normal fault surface, φa = 55 deg. (Published by permission of R. Bischke.)
and intersects the San Gabriel Fault at an apparent dip of 55 deg (βa on Fig. 12-28). Additional
subsurface data may result in alternative interpretations.
Model for Dip Domain 2. The kinematics of pure strike-slip folding shown in Figs. 12-26 and
12-27 permit a balanced and retrodeformable model of Ridge Basin. A profile in dip domain 2,
constructed parallel to the surface trace of the San Gabriel Fault, can resolve the rollover struc-
tural geometry (Fig. 12-29). If methods exist to estimate (1) the apparent collapse angle ψa (Bis-
chke and Tearpock 1999), (2) the apparent bed dip θa, and (3) the apparent dip αa of the fault at
depth, then the apparent bend φa in the fault surface is obtainable using either forward or inverse
modeling techniques, described in Chapter 11.
Notice in Fig. 12-27b that line GHE lies in the plane defined by the direction of fault slip,
and that material does not cross this plane. Also, notice that the Coulomb collapse plane BGC in-
tersects line GHE at an acute angle. This means that (1) not only should a balanced profile lie in
the plane defined by the direction of fault slip, but (2) the Coulomb collapse angle ψ must be
corrected for apparent dip (along with the other appropriate angles, such as bed dip θ).
Assuming a Coulomb collapse angle of 65 deg, we first calculate the apparent collapse
676 Chap. 12 / Strike-Slip Faults and Associated Structures
(a)
(b)
Figure 12-29 (a) - (b) Model for dip domain 2 adjacent to San Gabriel Fault. (Published by permis-
sion of R. Bischke.)
angle ψa from the strike direction of the Coulomb collapse surface. We obtain this angle from
the strike direction of the beds in domain 2 adjacent to the San Gabriel Fault (Fig. 12-24). In the
model in Fig. 12-27, strike is parallel to this axial surface trace line BG. The strike of dip domain
2 is N75W, as shown in Fig. 12-28. Accordingly, using stereographic projection methods, the ap-
parent dip ψa for a collapse plane trending at N75W and dipping at 65 deg, and lying in a profile
trending at N40W, is ψa = 51 deg (Fig. 12-28).
Second, the apparent bed dip θa at the intersection of domains 1 and 2 defines the plunge
angle of the Ridge Basin Syncline (Figs. 12-26 and 12-27), which is 24 deg NW. Therefore, the
apparent bed dip φa = 24 deg. This apparent dip lies in the same vertical profile as the strike di-
rection of the strike-slip fault, which in the case of the San Gabriel Fault is N40W (Figs. 12-24
and 12-28). Third, if a profile lies in the strike direction of an inclined strike-slip fault, then the
apparent fault dip αa = 0.0 deg. If the profile also crosses a branching fault that has an apparent
dip of βa in the profile, then the apparent bend φa in the fault surface is equal to βa (Fig. 12-29).
The value of φa, and hence βa, can be determined from a combination of inverse modeling and
stereographic projection techniques (Figs. 12-28 and 12-29).
The intersection of the Hungry Valley Branch Fault with the San Gabriel Fault defines the
apparent dip βa of the branch normal fault in the profile in Fig. 12-29. Using stereographic pro-
Balancing Strike-Slip Faults 677
jection methods, βa = 55 deg (Fig. 12-28). The value of φa is determinable from the apparent dip
θa of the hanging wall beds, taken in the direction of the plunge of the Ridge Basin Syncline. Re-
ferring to Fig. 12-29, if θa = 24 deg NW, then φa = βa = 55 deg. Thus, the dip and strike of the
shallow portions of the Hungry Valley Fault are 59E/N10E (Fig. 12-28).
Tests of Model. Stereographic projection methods determine the true bed dip θ for dip domain 2,
using the strike direction of the beds in domain 2 and the apparent dip of the beds θa. This appar-
ent dip angle θa lies in the same plane as the true bed dip (Fig. 12-28). Near Hungry Valley, the
strike of the beds in dip domain 2 is about N75W (Figs. 12-24 and 12-28). As the apparent dip of
the beds (θa = 24 deg) lies in a profile that trends at N40W, and in a plane that strikes at N75W,
the true dip of the beds is determined to be 38 deg in a N15E direction (Fig. 12-28). This calcu-
lated value of 38N/N75W for the dip and strike of the beds in domain 2 is close to the observed
average dips of 34N/N85W in dip domain 2 (Fig. 12-24) (Bischke 2002).
Plotted on Fig. 12-28 is dip domain 1, which dips and strikes at 25W/N10E. This plane in-
tersects dip domain 2 to form the Ridge Basin Syncline, which subparallels the surface trace of
the San Gabriel Fault (compare Figs. 12-24 and 12-28). In Fig. 12-28, the plane representing do-
main 1 (25W/N10E) intersects the plane representing dip domain 2 (36N/N75W) in a direction
that is subparallel to the surface trace of the San Gabriel Fault. Thus, there is a good correspon-
dence between the model and the observed surface bed dip data (Fig. 12-24).
Inverse Modeling of Dip Domain 1 and the Branch Normal Fault. Lastly, using the Extension In-
verse Theory described in Chapter 11 and in Bischke and Tearpock (1999), we determine the dip of
the deeper portion of the Hungry Valley Branch Fault, which forms the western boundary of Ridge
Basin. Inverse modeling, which is the reverse of forward modeling, employs surface bed dips or
depth-corrected seismic data to predict deep subsurface fault geometry. The consequence of decou-
pling known surface data from the unknown deep fault geometry is the introduction of uncertainty
into the subsurface depth predictions. This uncertainty results primarily from measurement errors
that are inherent in all inversion methods, including seismic and potential field methods. Although
inversion methods contain error, they have the advantage of being able to make predictions.
As an exercise, we invert the surface dip data and the 55-deg apparent dip on the Hungry
Valley normal fault to predict the gross crustal structure beneath Ridge Basin. This exercise re-
sults in a generic and retrodeformable model of Ridge Basin. Additional subsurface data, such as
seismic profiles, are required to confirm these predictions.
Inverse theory requires a number of assumptions that include knowledge of the Coulomb
collapse angle (typically between 60 deg and 70 deg), taken here to be 65 deg. If fault dip φ is
known within limits, then the surface bed dip θ predicts the subsurface fault dip angle α (Chap-
ter 11). This inversion uses a profile constructed parallel to the strike of the San Gabriel Fault, or
in the direction of fault slip, which is N40W.
Retrodeformable Fig. 12-30 represents any profile that trends in a direction of N40W across
dip domain 1, between Hungry and Peace Valleys in the northern portions of Ridge Basin
(Fig. 12-24). As the profile parallels the slip direction of the San Gabriel Fault, the bed dip angle
θa and deep normal fault dip angle αa are apparent dip angles. Domain 1 has an average bed dip
θ of 25 deg W, and it is separated from the gently dipping and folded Hungry Valley domain by
an axial surface (Fig. 12-24). Correcting the bed dips in domain 1 to the profile direction of
N40W results in an apparent dip θa of 20 deg. This profile intersects the Hungry Valley Fault at
an apparent dip of βa = 55 deg (Figs. 12-28 and 12-29). Determine the apparent dip αa of the
deeper part of the fault by using the method described in the section Procedures for Projecting
Large Normal Faults to Depth, in Chapter 11. As described in that section, use the distance D1
(Fig. 12-30) to establish the position of the deeper part of the branch fault. The dip angle αa then
678 Chap. 12 / Strike-Slip Faults and Associated Structures
Figure 12-30 Inversion model for dip domain 1, used to calculate dip αa of the normal fault at
depth. The listric normal fault causes the beds between Hungry and Peace Valleys to dip toward that
fault, thus forming the Hungry Valley depocenter. The interpretation based on the surface dip data
suggests that the 16-deg dipping normal fault may intersect the brittle-ductile transition beneath the
volcanic Soledad Basin. (Published by permission of R. Bischke.)
measures to be 12 deg (Fig. 12-30). Using the stereonet (Fig. 12-28), this 12-deg apparent dip of
the deeper fault, in a N40W (S40E) direction, establishes the plane representing the deep normal
fault. The true dip of the fault measures to be 16 deg in a S80E direction (Fig. 12-28).
A normal fault dip of 16 deg projects to the east to intersect the brittle-ductile transition at a
depth of about 20 km beneath the Soledad Basin that contains mid-Miocene volcanic rocks. This
prior volcanic activity may have weakened the crust, facilitating the crustal extension. A general-
ized fault surface map for Ridge Basin is shown in Fig. 12-31. According to the model, hanging
wall beds moving over the nonplanar fault surfaces deform at the bends in the fault surfaces. This
motion replicates the main structural features and bed dips observed in Ridge Basin. Thus, the
fault surface map represents a 3D fault surface model for the Ridge Basin structure (Fig. 12-24).
The faults are tied to cross sections constructed by Crowell (2002c) and to the Sun Schmidt Well
(Fig. 12-31). On Fig. 12-31, all the fault surfaces are planar and the lines of fault intersection are
determined from stereoplots. Where well control and seismic data are sparse, this procedure al-
lows for the projection of fault surfaces over long distances and provides strong constraints on an
interpretation (Bischke 2002). Methods involving curved line projection techniques require
dense data control and are not applicable to underconstrained data sets.
The generalized fault surface map shown in Fig. 12-31 is consistent with the major struc-
tural features and the bed dips observed in Ridge Basin. The generalized model shows that the
Ridge Basin syncline emanates from the branch point (BP) of the San Gabriel Fault with the
Hungry Valley Branch Fault. As the San Gabriel Fault dips at a higher angle in the southeast area
of the map, the intersection of the deep normal fault with the San Gabriel Fault migrates to the
Summary 679
Figure 12-31 Generic fault surface map for Ridge Basin, California. The interpretation of dip do-
main, seismic, and well data suggest that a branch fault forms a releasing bend in the area of Hun-
gry Valley. This fault geometry, when combined with Coulomb collapse theory, models the general
dip and strike pattern shown in Fig. 12-24. The synclinal fold structure emanates from branch points
located beneath Hungry Valley. Stereographic projection methods determine the fault plane inter-
sections. (Published by permission of R. Bischke.)
south. This may explain why the Ridge Basin syncline narrows to the southeast. The geometry of
the San Gabriel Fault controls the position of the Ridge Basin syncline.
SUMMARY
The balancing of strike-slip fault displacements is in its infancy and requires additional work in-
volving kinematic models that describe the structural styles observed along strike-slip fault
zones. In Ridge Basin, the good correspondence between model-based predictions and observa-
tion is encouraging. However, these kinematic models should be subject to additional tests utiliz-
ing well-constrained data sets and viable fault surface maps. Nilsen and McLaughlin (1985)
report that Ridge Basin contains some structures that are similar to Hornelen Basin in Norway
and to portions of Little Sulphur Creek Basin in northern California (Fig. 12-32). Two of these
basins contain a throughgoing strike-slip fault that bounds one flank of the basin, and talus and
debris flow deposits exist adjacent to the strike-slip fault. These syntectonic breccias demon-
strate growth through time. Paralleling the fault along the talus and debris flow deposits is a syn-
clinal fold structure (Fig. 12-32). Nilsen and McLaughlin (1985) conclude that many of the
world’s strike-slip extensional basins share similar tectonostratigraphic histories.
When examining a strike-slip structure, we should ask ourselves how the structure moved
into its present position. A kinematic description of displacement of the structure is essential
prior to embarking on the more theoretical analysis of how the structure reacted to stress and
680 Chap. 12 / Strike-Slip Faults and Associated Structures
Figure 12-32 Hornelen Basin (Norway), Ridge Basin (southern California), and Little Sulphur
Creek Basin (northern California), showing strike-slip fault and associated breccia and synclinal
structure that parallel the strike-slip fault zone. (From Nilsen and McLaughlin 1985. Published by
permission of the Society for Sedimentary Geology.)
strain. The Theory of Elasticity is clear on this subject: Kinematics is followed by dynamics
(Obert and Duval 1967; Ramsay 1967). Clearly, scientists must develop the kinematic equations
prior to formulating the differential equations for stress and strain. Structural studies that restrict
the analysis to stress, while ignoring displacement analysis, seem questionable and unrealistic.
CONCLUSIONS
We are unaware of any method obtained from seismic or well log correlation data, or from mod-
els concerning stress theory (or the strain ellipsoid), that leads to resolving the state of stress on
fault surfaces. Deductions concerning stress are not a viable exploration or prospect-generation
tool. Stress is a mathematical concept, is invisible, and its direction cannot be deduced from
Neotectonic or older faulting data. Furthermore, clay analog models of strike-slip faulting con-
Conclusions 681
tain a number of assumptions that affect the understanding of the structures and the styles of
faulting that form along strike-slip faults. Clay models do not scale to the real earth (Hubbert
1937) and suffer from near-field boundary condition problems that impart contractional motions
across a fault surface. The clay models do not correctly model the orientation or the regional dis-
tribution of real structures (Figs. 12-3, 12-4, and 12-5). Geoscientists are best served by applying
correct interpretation procedures and mapping techniques to their data and prospects, rather than
relying on theoretical models involving stress or strain.
Early in an exploration project, geoscientists should concentrate their interpretation on observ-
able and reliable data, and construct viable maps of recognizable fault surfaces. During the inter-
pretation and construction of admissible fault surface maps, geoscientists will gain insight into the
appropriate tectonic style affecting their area. When the mapped data lead geoscientists to a viable
tectonic style, then they may employ a model to better understand and interpret the area. If inter-
preters approach a data set with preconceived ideas of the structure or bias toward a particular
model, then the results of the exploration project are predictable – the interpretation and maps will
conform to the preconceived ideas based on the model. A data-first, models-second approach is a
more objective approach to any data set involving subsurface structure. Geoscientists then develop
the appropriate model from the data, rather than interpreting the data using a convenient model.
This approach is more likely to result in admissible interpretations and maps of the subsurface that
are valid in 3D. This approach will minimize structural risks where exploration prospects exist.
Strike-slip fault interpretations should present direct evidence for horizontal displacements.
As strike-slip faulting is a 3D problem, 2D seismic profiles can, at best, provide only suggestions
of strike-slip faulting. However, the construction of viable maps of the faults is the first step in
providing direct evidence for the horizontal displacements that are required for strike-slip inter-
pretations. These viable maps of fault surfaces should show evidence of curved, or bent, fault
surfaces. Strike-slip displacements at the bends will create restraining and releasing bends. Thus,
restraining and releasing bends seem fundamental to the existence and understanding of strike-
slip faults (Crowell 1974a and b). Their ubiquitous presence along the major strike-slip fault
zones presents direct evidence for strike-slip faulting. However, restraining and releasing bends
may not record the total amount of strike-slip displacements. The bends may be obvious, such as
the large bend along the Transverse Ranges of California (Fig. 12-10b). Alternatively, the bends
may be subtle like the bend at Loma Prieta along the San Andreas Fault (Fig. 12-18). Most im-
portantly, these bends contain information that can confirm the presence of the horizontal dis-
placements that accompany strike-slip faulting. If strike-slip faulting is present on a subsurface
map, one only needs to look for the presence of restraining and releasing bends. Restraining
bends should map as structural highs and releasing bends as structural lows. If evidence for re-
straining and releasing bends is not present on a regional scale subsurface map, then another in-
terpretation is probably warranted. If the strike-slip fault breaks the surface, then look for offset
rock types on the surface geologic map (Crowell 1974a and b).
We discussed several published techniques for recognizing strike-slip displacements. Three-
dimensional regional and local restorations of strike-slip faulting may be supportive of horizon-
tal displacements. Fault surface maps may show curved fault surfaces that form restraining and
releasing bends. Folds form as the hanging wall beds move over curved fault surfaces (Bally et
al. 1966; Rich 1934). Fault bend folding along strike-slip faults may prove useful in describing
and analyzing structures in fault bends and along mapped fault surfaces. Balanced cross sections
of restraining and releasing bends provide constraints on structural interpretation problems that
lead to 3D structural validity and to admissible interpretations of subsurface data. Balanced cross
sections of restraining and releasing bends should assist industry to generate high-quality
prospects with reduced risk. In short, strike-slip fault interpretations should be subject to the
same high-quality and rigorous interpretation techniques required of other styles of faulting.
CHAPTER 13
GROWTH
STRUCTURES
INTRODUCTION
Tectonic and sedimentary growth is important to petroleum exploration. For example, growth
along faults creates both structural traps and seals (Brenneke 1995), and this growth can be coin-
cident with the deposition of reservoir units. Growth faults may trap hydrocarbons during the
early stages of a hydrocarbon cycle. The expanded sections that typically form downthrown to
down-to-the-basin normal faults are the highest growth sections found in the petroleum regime
(Thorsen 1963). These expanded downthrown sections often contain large accumulations of hy-
drocarbons (Tearpock et al. 1994). Hydrocarbon accumulations correlate not only to high growth
intervals but often to the highest growth intervals within the petroleum system (Fisher and Mc-
Gowen 1967; Woodbury et al. 1973; Branson 1991; Pacht et al. 1992). Therefore, in order to find
large accumulations of hydrocarbons, we must understand and analyze growth.
We emphasized the importance of constructing a three-dimensional structural framework in
Chapters 7 and 8. In this chapter we address the fourth-dimensional time factor, which is important
when exploring for and developing hydrocarbons. We consider the time variable through the super-
position, or relative age, principle. In other words, the oldest rocks are the deepest rocks, except for
the inverted sequences discussed in Chapter 10. Rocks record cycles of changing sedimentary in-
flux that relate to systems tract and sequence boundaries. For example, if we define the thickness of
an interval of sediments as d, and time as t, then the sedimentation rate is d/t. The sedimentation
rate for that interval may change between two locations, and the change in rate is ∆d/t. We can take
the ratio of the change in sedimentation rate to the sedimentation rate, or (∆d/t)/(d/t). In this ratio,
time cancels out. This dimensionless growth (Suppe et al. 1992), or relative age, ratio ∆d/d is a
measure of growth (Bischke 1994b). When using subsurface data, information on growth and rel-
ative age is always available to evaluate growth and its impact on exploration.
682
Introduction 683
We demonstrate from correlation data that each sedimentary sequence, being a genetically
related unit, experiences different patterns of growth. Furthermore, sedimentary sequences in all
tectonically active regimes exhibit growth, not just in the extensional growth fault regime. Typi-
cally, linear or monotonic patterns of growth characterize sedimentary sequences. The linear
growth recorded in a sequence is typically punctuated by abrupt changes in growth rate. These
punctuated growth patterns may aid in defining systems tract and sequence boundaries, and they
can contribute to the resolution of a variety of practical problems involving changes in growth.
When geoscientists think of growth faulting, they are often referring to the large listric,
growth normal faults. However, growth faults can be small, with only tens of feet of vertical dis-
placements. Any normal fault is a growth fault if the stratigraphic section expands across the
fault (Thorsen 1963; Dawers and Underhill 2000). Furthermore, all structural styles can exhibit
growth, including growth reverse and growth thrust faults (Suppe et al. 1992) as well as
growth strike-slip faults (Shaw et al. 1994; Bischke 2002), and growth inversion structures
(Mitra 1993; Link et al. 1996). We discuss these and other complex structural styles that are
sometimes misidentified with well log and seismic data. Structural styles are important because
the recognition of the correct style for a region affects our interpretations and our understanding
of the petroleum system. We discuss techniques for recognizing structural styles using growth
analysis, with examples of growth structures from the four major tectonic regimes: extensional,
compressional, strike-slip, and salt. Large accumulations of hydrocarbons correlate to timing of
migration and fault movement, so it is important to understand the timing related to fault growth
and deposition.
An understanding of growth history, when integrated with geologic and geophysical data,
can help solve a number of existing or potential problems that include (1) rapidly distinguishing
faults from unconformities and confirming small faults or uncertain fault interpretations; (2) lo-
cating sequence boundaries and subtle stratigraphic traps and predicting erosional surfaces and
potential bald structures on the crests of anticlines; (3) solving general correlation problems; (4)
locating channel sands; (5) locating the highest growth or highest petroleum potential intervals;
(6) determining the time of structural growth and timing of faults; (7) checking interpretations
for consistency with the existing growth history; (8) recognizing unidentified problems; and
(9) quality-controlling the input values in a well log data base.
We shall discuss two techniques for studying growth: the Expansion Index Method
(Thorsen 1963) and the high-resolution ∆d/d technique and its related methods (Bischke 1994b;
Sanchez et al. 1997). We describe practical examples pertaining to sequence stratigraphic inter-
pretations, to strike-slip and compressional deformation, to unconformities and correlation prob-
lems related to salt, and to the extensional regime that typically contains growth sediments.
Thickness Downthrown
Expansion Index =
Thickness Upthrown
Where growth is recognized, an expansion index greater than 1.0 typifies growth. If the ex-
pansion index ratio is less than 1.0, then a correlation error has been made or an unconformity is
present in the hanging wall fault block.
Figure 13-1 is a simple example to illustrate the determination and application of the expan-
sion index for a generic growth fault. The left side of the figure contains a cross section of a
growth fault with one well in the upthrown block and one in the expanded, downthrown block.
For simplicity, the correlative horizons, a through k, in the upthrown fault block bound units that
are all of equal thickness (t = 1.0). The thickness of each unit in the downthrown block is used to
calculate the expansion index for each.
Thorsen presents the results in the form of a bar graph on the right side of Fig. 13-1, and the
fault movement with time is readily apparent. By analyzing the graph of the expansion indices,
we conclude that movement on the fault began at time j and reached a maximum period of
growth activity between times f and g. Fault movement ceased at time b. The maximum period
of fault growth between times f and g involved a 2.1 to 1 expansion of the downthrown strati-
graphic section. If the interval f to g contains thick sands, then this thick interval has the highest
potential for hydrocarbon accumulation.
Each growth fault appears to have its own unique fingerprint in reference to its expansion
indices. The rate of movement on a growth fault increases from the initial movement to some
Figure 13-1 Expansion index can be used to measure the timing and rate of movement along
a growth fault. Expansion indices are commonly illustrated in bar graph form (right side of figure).
(From Thorsen 1963. Published by permission of the Gulf Coast Association of Geological
Societies.)
Introduction 685
maximum rate, and then decreases again until fault movement ceases. During its life, a growth
fault may move sporadically, resulting in an irregular growth history. Each growth fault has its
own unique growth history as a result of fault movement, the rate of subsidence, and the rate of
sediment supply across the fault. Even over long horizontal distances, the fingerprint for a partic-
ular fault can usually be recognized along strike, as shown in Fig. 13-2, even though the values
of the expansion indices may not be constant along strike. The plots of the expansion indices for
this fault, calculated at separate locations two miles apart, are similar. A review of the individual
expansion indices and the overall pattern at each location shows that they are similar enough to
be identified as representing the same fault.
Because of the unique characteristics of the plot of the expansion indices for each growth
fault, the technique can be used in complexly faulted areas as a lateral correlation tool for the
identification of individual growth faults over long distances. The expansion index technique is
applicable for use with either electric well logs or seismic sections. Because of the greater strati-
graphic detail that is found in well logs versus seismic sections, expansion indices calculated
using well logs are more accurate. One important note needs to be made with regard to this tech-
nique and seismic sections. As time distorts vertical distance on seismic sections, always depth-
convert the seismic sections before calculating the expansion index.
The expansion index also serves as a vertical correlation aid for well log and seismic inter-
pretations. The thickness of the stratigraphic section in the upthrown block should almost always
be equal to or less than the thickness for the equivalent section in the downthrown block, so the
expansion index that is calculated for any specific unit is typically greater than 1.0. Therefore, a
calculated expansion index less than 1.0 using well logs or seismic data either indicates a bust in
Figure 13-2 Comparison of expansion indices, taken over 2 miles apart, indicates that a growth
fault’s fingerprint can be recognized over long distances. (From Thorsen 1963. Published by permis-
sion of the Gulf Coast Association of Geological Societies.)
686 Chap. 13 / Growth Structures
Figure 13-3 Expansion indices can be used to date growth faults with regard to their initial move-
ment, time of maximum growth, and last movement. (From Thorsen 1963. Published by permission
of the Gulf Coast Association of Geological Societies.)
the correlation of the individual units across the fault, or an unconformity in the hanging wall
fault block.
Figure 13-3 shows the analysis of three separate growth faults in South Louisiana. The plots
of the expansion indices indicate that the age of each fault is different and that they become pro-
gressively younger in a basinward direction (see map insert). The time of maximum fault
growth, indicated by dark arrows, corresponds to the age of maximum sedimentation in the area
across each fault. In certain geologic settings, such information is vital to the exploration for hy-
drocarbons. In the Miocene trend onshore and offshore Louisiana, for example, many of the
major structural features relate to the genesis of diapirs of the Jurassic Louann Salt, which di-
rectly ties to the timing of the sediment load across growth faults.
The analysis of a growth fault, based on the pattern of its expansion indices, provides infor-
mation important to the exploration for hydrocarbons. If time information is available, then at
least eight benefits derive from analyzing the expansion index plot of a growth fault.
Some disadvantages to using the expansion index method are that it has a low resolution
and may contain mathematical errors that can be significant (Bischke 1994b). In the next section
we describe the related, but high-resolution, ∆d/d technique, also referred to as the Multiple
Bischke Plot Analysis (MBPA). This technique minimizes errors, with maximum errors on the
order of 0.1 percent.
and is more accurate. The method allows for the recognition of faults, unconformities, and incor-
rect correlations in the form of discontinuities, or anomalies, on the plots.
Method. In this section we describe the high-resolution ∆d/d and MBPA methods and present a
number of practical examples of these methods. The methods have application from the produc-
ing-field scale, with excellent well control, to the regional basin scale with minimal well con-
trol. We present several examples of salt-related problems that are not resolvable using
conventional techniques, but are resolvable using growth plots. We first describe the ∆d/d
method, which is the foundation for MBPA.
To begin the method in an area of study, locate two wells in a general dip direction, so that
one well is structurally higher than the other. Consider first a stable, or nongrowth, tectonic envi-
ronment in which the correlative intervals between two wells have about the same thickness
(Fig. 13-4a left). Correlative horizons are represented by correlative markers in the well logs. For
simplicity in Fig. 13-4a, the top correlative marker is at the same depth in both wells. The differ-
ence in depth of the lower marker reflects growth. Plot this depth difference (∆d) of each marker
against its SSTVD depth d in the structurally higher well (Fig. 13-4a right). In a stable tectonic
environment, the slope on the ∆d/d plot is approximately flat.
In an unstable, or growth, tectonic environment, differences in the thickness of correlative
stratigraphic intervals are greater, and thus larger vertical distances (∆d) separate the horizons in
the wells (Fig. 13-4b left). Typically, only two wells establish this relationship, although the
wells need to be in the dip direction. The resulting ∆d/d plot (Fig. 13-4b right) has a higher, or
steeper, slope than the one for a stable environment.
Using a number of correlation markers generates more points for the ∆d/d plot. The slope of
the curve on the plot reflects the growth history of the interval plotted. In applying the technique,
use many correlation markers that represent time-stratigraphic horizons. Such horizons include
parasequence boundaries, which are valuable in MBPA. Determine the difference in depth of
each marker in the pair of wells, and then generate a plot of the data. If you enter the subsea
depths of the correlations in a spreadsheet, simply calculate the difference in depths of the corre-
lations in the structurally higher well and those in the lower well to generate ∆d values. Place the
subsea depths of the markers in the higher well on the x-axis, and then plot the ∆d value corre-
sponding to each marker on the y-axis. This is best done in a spreadsheet program that contains a
graphics package.
A pair of wells is always used in the analysis, but one of the wells can then be compared
with a third well, a fourth well, and so on. This comparison forms the basis of the MBPA de-
scribed in the next paragraph. A plot is generated for each pair of wells. To provide more analyti-
cal power to the method, when a given up-dip well is compared with a number of other wells,
place all the plots generated on a single graph (Fig. 13-17b). This type of plot provides signifi-
cant information on the overall area of study in terms of growth, faults, and unconformities. We
provide several examples of the ∆d/d analysis and MBPA later in this chapter.
When conducting an MBPA, compare a reference well to any number of adjacent wells by
using a set of ∆d/d plots of the reference well versus each of the other wells (Sanchez et al. 1997).
The reference well on the cross plots could be a type well that has a complete stratigraphic sec-
tion. Alternatively, the reference well can be any other well positioned on any structural level.
Unlike the ∆d/d method, plot the depths in the reference well on the x-axis, even though it may
not be the structurally higher well. Thereby the reference well is present on all the plots.
The advent of 3D statistical programs enhances MBPA by allowing the comparison of a ref-
erence well to any number of other wells. The procedure rapidly generates ∆d/d plots, printed on
a single page. A visual comparison of these plots rapidly locates tectonostratigraphic anomalies.
Multiple Bischke Plot Analysis and ∆d/d Methods 689
(a)
(b)
Figure 13-4 Sedimentary environments can be categorized into stable and unstable growth envi-
ronments. (a) In a stable, nongrowth environment, the vertical distance between correlative markers
is small. If the markers are plotted on ∆d/d plots, then the curves will have a gentle slope. (b) In an
unstable, growth environment, the vertical distance between correlative markers is large, and the re-
sulting ∆d/d curves are more steeply sloping. (Modified from Bischke 1994b; AAPG©1994, pub-
lished by permission of the AAPG whose permission is required for further use.)
If the anomaly is present in the reference well, then the anomaly exists on all plots. However, if
the anomaly is present in only one well, then the anomaly is in the well used for comparison,
rather than in the reference well. This process allows interpreters to rapidly identify problem
wells and to better understand the cause of the anomaly (Sanchez et al. 1997; Chatellier and Por-
ras 2001).
Structural and stratigraphic analysis, based on comparison of ∆d/d plots, can be enhanced if
stratigraphic data are posted on the plots, as in Figs. 13-14 and 13-17b. Then MBPA can be ap-
plied with respect to discrete stratigraphic intervals and thereby provide information directly rel-
evant to the history of the study area.
690 Chap. 13 / Growth Structures
Sanchez et al. (1997) describe their experience with previously correlated well logs. Should
interpreters use pre-existing correlations, and perhaps be prejudiced by these correlations? Often
this results in purely cosmetic changes to the correlations, adding little to the understanding of
the field. A second, more time-consuming but more objective approach is to recorrelate the logs
ignoring the existing correlations. A third, less time-consuming approach is to subject the exist-
ing correlations to MBPA in order to rapidly identify correlation problem areas.
Consider the purchase of an older, producing field. The task at hand, besides producing the
proven reserves, is to find any additional potential. How do we, as geoscientists and engineers,
find additional potential in previously worked areas? New potential is found by disagreeing with,
and changing, the previous correlations — in other words, finding error in the older work. We
must remember that oil and gas reserves are not found by making maps. It is correlation work
that identifies the potential, and then the mapping based on the correlations that graphically dis-
plays that potential.
If we take previous geoscience and engineering work, in the form of correlations of electric
logs or correlations on seismic sections, and submit these correlations to a test of validity
(MBPAs), we can rapidly identify problem correlations, previous correlation errors, faults and
even unconformities. We can define the growth history of the area, providing us guidance as to
the stratigraphic intervals that might hold the best potential for hydrocarbons. And most impor-
tantly, we can do this while reducing the cycle-time of the correlation and mapping process by
as much as 50 percent (Bischke and Tearpock 1999).
This methodology should be applied to all correlation data sets during the initial stages of
any project. Numerous field studies show that the MBPA method, when applied at the beginning
of a study, results in a higher quality product and can aid in the rapid identification of new hy-
drocarbon potential.
Figure 13-5 In extensional regimes, three patterns of discontinuities are observed on ∆d/d plots.
The discontinuities on the plots are caused by missing section that may be due to a fault or an un-
conformity. (a) If the missing section is in the structurally lower well, then the plot of ∆d/d contains
a negative, or downward, discontinuity. We use an example of a fault. (Published by permission of
R. Bischke and D. Tearpock.) (b) If the missing section is in the structurally higher well, then the plot
contains a positive, or upward discontinuity. We use an example of a fault. These patterns can help
distinguish faults from unconformities in areas of high bed dip. (Published by permission of R. Bischke
and D. Tearpock.) (c) A hiatus causes a change in slope on the plot. (From Bischke 1994b;
AAPG1994, published by permission of the AAPG whose permission is required for further use.)
692 Chap. 13 / Growth Structures
ever, if it covers a large area, appears at about the same stratigraphic level in the wells in the area
and typically dips at a lower angle. For example, a so-called “fault,” interpreted to be the cause
of the same stratigraphic section missing in many wells in a study area, actually would be an un-
conformity rather than a fault. The existence of missing section at the same stratigraphic level is
obvious on a set of ∆d/d plots. Accordingly, geoscientists employ the method to rapidly locate
unconformities and to rapidly distinguish faults from unconformities in well log data. If many
wells exist in an area of study, then the plots aid in keeping track of the missing section and its ap-
proximate value and stratigraphic level in each well (Fig. 13-17b). The plots represent a record of
the missing section in the form of a visual display that is readily apparent. Refer to Fig. 13-17b
and review Fault J and the recognized unconformity.
Often a more difficult problem develops for geoscientists where structure was growing very
slowly relative to changing sedimentation rates or where sedimentation rates were high relative
to deformation rates. In these cases, it is difficult to detect disconformities in well log data, and
downlap may not be obvious or present on seismic sections. If there were a hiatus in deposition
(a condensed section), then when sedimentation resumes, the rate of deposition would likely
change. Tobias (1990) stresses this point. Subtle unconformities, caused by gradual processes,
occur as changes in slope of the ∆d/d curves (Fig. 13-5c). Alternatively, if there is the slightest
change in the sedimentation rate, regardless of its cause (e.g., sea level fluctuations), then chang-
ing environmental conditions are likely to produce changes in the slope of a ∆d/d curve dis-
played on the plot. These plots are very sensitive to changes in the tectonically or eustatically
controlled sedimentation rate, so disconformities and/or subtle sequence boundaries commonly
occur as breaks in slope of the curves generated on the ∆d/d plots (Fig. 13-5c).
Unconformity Patterns
Two unconformity growth patterns exist, one for onlap and the other for downlap. These patterns
may help define systems tract boundaries.
If beds onlap an unconformity, then at the time of deposition the beds above the unconfor-
mity dip at a lower angle than the beds below the unconformity. Figure 13-6a shows the typical
occurrence of onlap above an unconformity, with the missing section being greater in the well
located at the structurally higher position at the time of deposition. Erosion typically removes
more section in the up-dip direction than in the down-dip direction. The missing section caused
by the unconformity occurs as a positive displacement on the ∆d/d plot (Fig. 13-6a), even if
strata were later structurally rotated. Thus, if a positive discontinuity is present on a ∆d/d plot
(Fig. 13-5b), the missing section results either from a fault in the structurally higher well or from
erosion, with subsequent onlap of sediments above the unconformity.
If beds downlap an unconformity, then at the time of deposition the beds above the uncon-
formity dipped at a higher angle than the beds below the unconformity. Downlap causes a nega-
tive discontinuity on ∆d/d plots (Fig. 13-6b), even if strata were structurally rotated. Again, more
section is missing in the area that was up-structure at the time of deposition. Thus, if a negative
Multiple Bischke Plot Analysis and ∆d/d Methods 693
(a)
(b)
Figure 13-6 (a) Onlap produces a positive discontinuity on the ∆d/d plot, whereas (b) downlap
creates a negative discontinuity. (From Bischke 1994b; AAPG1994, published by permission of the
AAPG whose permission is required for further use.)
discontinuity exists on a plot, then the missing section responsible for the discontinuity results
either from a fault in the structurally lower well or from erosion, with subsequent downlap of
sediments above the unconformity.
These patterns enable us to distinguish faults from unconformities on the steeply dipping
flanks of salt structures, where bed dips are commonly too steep to image on seismic data. For
example, on the flanks of salt diapirs the sedimentary units typically onlap, rather than downlap,
an unconformity. Therefore, unconformities on the flanks of salt domes produce positive, rather
than negative, discontinuity patterns (Fig. 13-6a). This concept can be used in comparing an up-
dip well to one or more structurally lower wells. If missing section causes a negative displace-
ment on the plots, as shown in Fig. 13-5a, then the missing section in the structurally lower well
must be due to a fault rather than to an unconformity.
The larger the missing section, the larger the discontinuity in the ∆d/d plots. However, con-
sider that an unconformity removes section in both of the wells used in a ∆d/d plot (Fig. 13-6).
694 Chap. 13 / Growth Structures
Then the amount of missing section, estimated from that plot, is the difference between the
amounts of the two missing sections, rather than the actual amount of section missing in one of
the wells.
ACCURACY OF METHOD
Well log data has a higher resolution than 3D seismic data. At a depth of –15,000 ft, geologists
can commonly determine the depths between two correlation markers to within a few feet (e.g.,
5 ft). A ∆d/d plot is a plot of change-in-depth versus depth. Therefore, if we define error as the
precision in correlation between two tops, divided by the subsea depths to the correlations, then
the error in well log data at –15,000 ft is 5/15,000, or 1 part in 3,000. This is a very high degree
of accuracy. Thus, ∆d/d plots constructed from well log data typically have errors on the order of
less than 0.1 percent.
Seismic data at a depth of –15,000 ft may have a velocity of about 10,000 ft/sec one-way
time, and a frequency of 30 Hz. These data would have a wavelength of about 500 ft and a reso-
lution of about 100 ft (Sheriff 1980). At a depth of –15,000 ft, seismic data has an error of
100/15,000 or about 1 part in 150, or about 1.5 percent. Well log data are at least one order of
magnitude more precise than seismic data. The intent of this discussion is not to degrade seismic
data that we use on a daily basis, but rather to point out the accuracy of the methods when using
well log versus seismic data.
The plots will contain errors in the slope of the growth curves if a well encounters changes
in bed dips or changes in wellbore deviation angle (Chatellier and Porras 2001). This error af-
fects only the slope of the growth curves, and the error increases as the tangent of the change in
bed dip angle or wellbore deviation angle (Bischke 1994b). The plots of (1) vertical wells
against the ramp portion of deviated wells or (2) two subparallel deviated wells, compared to
each other, contain fewer errors than plots based on highly deviated wells and horizontal wells,
and these latter types of plots should be avoided. Thus, during correlation work, if vertical wells,
the ramp portion of deviated wells, or two subparallel deviated wells are available for study, then
it is generally not necessary to correct the data. If, however, the intent of the study is to distin-
guish high-growth intervals from low-growth intervals, and the dipmeter data indicate rapidly
changing bed dips, then the data can be corrected for changing bed dips by using the correction
factors presented by Bischke (1994b). The errors present in uncorrected data accumulate as the
tangent of the change in bed dip angle; for example, if the bed dip changes by 10 degrees, then
the error introduced in the slope of the curves is 18 percent.
Figure 13-7 Stratigraphic interpretation of a generic delta based on well control. Parasequence
correlations define a nongrowth (pre-growth) section in the wells beneath a disconformity located at
correlative horizon 3. Growth section above the disconformity contains an expanded section be-
tween Wells No. 1 and 2, and a condensed section between Wells No. 2 and 3. Expanded and con-
densed sections are characteristic of wells located across a delta, and thus the method may define
sedimentary environments in areas that lack dense well control. A seismic line located between
Wells No. 1 and 3 would not detect the hiatus. (From Bischke 1994b; AAPG1994, published by
permission of the AAPG whose permission is required for further use.)
wells for each correlation marker. Horizontal lines are drawn between Wells No. 1 and 2 to illus-
trate the depth of correlation markers 1 through 6 in Well No. 1. Measure the vertical distance
(difference in depth) for each of the correlative markers, 1 through 6, in each pair of wells. Plot
these ∆d values on the y-axis against the subsea depth of the correlations in the structurally
higher well, plotted on the x-axis (Fig. 13-7). In practice, calculate the ∆d measurements on a
spreadsheet by entering the subsea depths as positive values and subtracting the depths in the
structurally higher well from the depths in the structurally lower well. The plot for Wells No. 1
and 2 are shown in Fig. 13-7. These data can be from anywhere in the world and from any
tectonostratigraphic environment, yet the following interpretation will be correct.
The interval between correlations 1 to 3 is an expanded, growth interval, and the interval be-
tween correlations 3 through 6 is a nongrowth interval. We can see that the interval 1 to 3 has a
positive slope on the plot, demonstrating expansion from Well No. 1 toward Well No. 2. The
slope is flat between correlations 3 through 6, indicating no growth. As no onlap is present, the
break in slope at correlation 3 is a possible disconformity that would not image on seismic pro-
files between Wells No. 1 and 2.
To continue the analysis, we conduct the same procedure between Wells No. 2 and 3. In this
case, a sequence boundary, labeled 0, is present in the two wells (Fig. 13-7). Again, a ∆d/d plot
shows the growth history relative to the two wells. Well No. 2 is the structurally higher well, so it
is plotted on the x-axis (Fig. 13-7).
696 Chap. 13 / Growth Structures
A unique interpretation of the plot of Wells No. 2 and 3 is made. The interval between cor-
relations 1 and 3 is a condensed section in Well No. 3, and the interval between correlations 3
through 6 is a nongrowth interval. The interval 1 to 3 has a negative slope, which means that the
section thins from Well No. 2 towards Well No. 3, indicating a condensed section in Well No. 3.
The slope is flat between markers 3 through 6, confirming the nongrowth interval. The break in
slope at correlation 3 is probably a time-transgressive disconformity that exists in all three wells.
We can now conclude that the growth section expands between Wells No. 1 and 2 and contracts
between Wells No. 2 and 3, which is characteristic of only a few sedimentary environments, in-
cluding a delta. When integrated with other geologic information, the technique can identify or
constrain the number of possible structural styles or sedimentary environments in the area of
study, even where data are limited.
(a)
(b)
Figure 13-8 (a) Seismic section in the Brazos Ridge area, northern Gulf of Mexico. Inclined
dashed lines are locations for correlation data in a ∆d/d plot. (b) A ∆d/d plot of data, derived from (a),
reflects changes in growth history of the structure. The breaks in slope, at data points 8, 12, 18, and
22, mark major sequence boundaries. (Bischke 1994; AAPG1994, reprinted by permission of the
AAPG whose permission is required for further use.)
698 Chap. 13 / Growth Structures
above –1000 ft. The maximum growth occurred in this area between correlations points 12 and
18. And finally, the breaks at correlation points 8, 12, 18, and 22 appear to mark major se-
quence boundaries.
This one ∆d/d plot, from one seismic section from a 3D seismic survey, provides significant
information about the stratigraphic and growth history of this area of the Brazos Ridge. Plots
from other seismic sections also exist. The combined use of multiple ∆d/d plots, using seismic
sections like the one shown in Fig. 13-8a, provide significant information on the tectonostrati-
graphic history of an area and begin to define its hydrocarbons potential.
(a)
(b)
Figure 13-9 (a) Growth plot of the type Well No. 5 and the difficult-to-correlate Well No. 1 in the
Eugene Island 208/215 field. Correlation marker No. 27 lies 230 ft off the general growth trend, sug-
gesting that this anomaly is due to miscorrelation or to faults in the two wells. See text for explana-
tion that marker No. 27 in Well No. 1 is miscorrelated to the type log by about 230 ft. (b) The 8600-ft
sand in Well No. 5 was initially incorrectly correlated to the 7800-ft sand in Well No. 1. The ∆d/d
method, when integrated with other geologic information, demonstrates that the 8600-ft sand in Well
No. 5 correlates to the sand below 8000 ft in Well No. 1. (From Bischke et al. 1999. Published by per-
mission of the Gulf Coast Association of Geological Societies.)
700 Chap. 13 / Growth Structures
growth trend is identifiable on the plot, then the method directs the interpreter to an alternative
correlation. The method uses a nonarbitrary process.
Figure 13-10 Growth plot illustrating carbonate ramp in Grayburg formation, Guadalupe Moun-
tains, New Mexico, USA. The change in slope at the top of Zone 2 partitions the Grayburg Formation
into high-growth and low-growth sequences, and indicates a probable tectonostratigraphic bound-
ary. A genetic relationship exists between Zone 1 in the Grayburg Formation and the overlying
Queen Formation. (Published by permission of R. Bischke.)
Examples of the ∆d/d Method 701
1 to 5 within the Grayburg Formation and tops of several parasequences in the Queen Formation.
Zone 1, at the top of the Grayburg Formation, is within a low-growth section (Fig. 13-10). This
linear low-growth pattern continues into the Queen Formation, and thus Zone 1, in the Grayburg
Formation, appears to be genetically related to the Queen Formation. Contrastingly, a high-
growth section is seen in Zones 2 through 5. Growth rate was highest during deposition of Zone
5, then gradually declined through Zone 2. Zones 2 to 5 in the Grayburg Formation define the
carbonate ramp.
A disconformity appears to exist at the top of Zone 2 within the Grayburg Formation (com-
pare Figs. 13-5c and 13-10). We interpret this hiatus to result from a change in sea level. The
growth plot, constructed from correlations based on well log and core data, is compatible with a
genetically related boundary at, or very near, the top of Zone 2 in the Grayburg Formation. Zone
1 in the Grayburg is genetically related to the Queen Formation. A ∆d/d plot of Wells EMSU 458
and 247, not shown in this book, yields similar results with the carbonate ramp represented be-
neath the top of Zone 2. If sedimentary sequences exist as genetically related units, regardless of
the cause, then these units are likely to be related to sediment supply. Changes in sediment sup-
ply should occur at or near system tract boundaries, where changes in sea level affect growth. We
conclude that changes in growth may occur at, or near, major sequence boundaries and may aid
in locating the position of sequence boundaries (Tobias 1990; Sanchez et al. 1997). The method
can help distinguish between lithostratigraphic and tectonostratigraphic boundaries.
Figure 13-11 Growth plot for Huntington Beach Field Anticline, Newport-Inglewood Trend, Califor-
nia, USA. The compressional fold grew during the Pliocene. The highest growth interval is between
the Top Repetto and Top Lower Pico section. Disconformities exist at discontinuities on the plot, indi-
cated by the circled numbers. The California Division of Oil and Gas mapped these disconformities
between fields in the Newport-Inglewood Trend.
were not subject to folding (Harding 1973). The conclusion is that the migration of hydrocar-
bons into the Huntington Beach Anticline was post-Miocene.
Lastly, the plot helps to interpret the structural style. The Newport-Inglewood Trend is sub-
ject to several interpretations. Harding (1973) proposes that strike-slip displacements caused,
and were coincident with, the compressional folding. Wright (1991) employs structural maps
and well log data from fields along the trend to conclude that the folding and faulting exhibit a
complex structural and stratigraphic history, not readily reconciled with a simple strike-slip ori-
gin. From a seismic study of the Los Angles Basin, Shaw and Suppe (1996) conclude that the
Los Angles Basin is subject to large-scale, blind-thrust faulting, and that the folding is not coeval
with recent strike-slip displacements on the Newport-Inglewood fault system. Thus different in-
terpretations are possible for the structural formation of the local folds. Which is correct? Can
the growth plot suggest the correct solution? The ∆d/d plot seems to support the interpretation of
Shaw and Suppe (1996). In the same Newport-Inglewood trend, a study of several growth plots
and a restoration of the Signal Hill restraining bend, at the Long Beach Anticline, provide evi-
dence that the strike-slip faulting post-dates the compressional folding. The Signal Hill analysis
is presented in Chapter 12.
A ∆d/d plot of the correlations presented in Harding (1973), when integrated with the local
geologic data (Wright 1991), suggests that the Huntington Beach field is a growth anticline (Fig.
13-11). A growth compressional style may be new to many geoscientists and engineers alike. A
Examples of the ∆d/d Method 703
lack of knowledge of growth compressional structures can cause misinterpretations that may af-
fect regional interpretations, the interpretation of the local petroleum system, and ultimately the
prospects generated. Any interpretation based on existing correlations must be consistent with
the correlations, regardless of the structural style under study.
Figure 13-12 Growth plot for a splay of the Zayante Fault, California. Growth began between the
deposition of parasequence boundary correlations 5 and 6 in the late Pliocene. Growth continues
into the Recent, as evidenced by offset terraces. (From Shaw et al. 1994. Published by permission of
the United States Geological Survey.)
between parasequence boundaries 5 and 6, and continued into the Recent, as evidenced by the
offset terraces (Fig. 13-12). Other ∆d/d plots from the area show similar growth histories on
different splays of the Zayante Fault. Therefore, in some cases ∆d/d plots can detect time of
movement on strike-slip faults. If hydrocarbon migration occurred within the fault zones during
periods of movement, the ∆d/d analysis of fault timing provides valuable information regarding
petroleum potential. The structures formed by movement on the fault(s) would necessarily pre-
date hydrocarbon migration in order for accumulation to occur.
ence well, can be grouped on a single plot to form a stacked Multiple Bischke Plot (Chatellier
and Porras 2001). Stacked MBPs allow geoscientists to compare thickness changes between
wells, or to locate an anomaly or missing section that exists in every well. As an example of a
stacked MBP, see Fig 13-17b.
Another type of analysis is the inverted MBP (Chatellier and Porras 2001). For this analy-
sis, the axes on the standard ∆d plot are inverted, with ∆d plotted on the x-axis and d plotted on
the y-axis. Inverted MBPs contain properties similar to conventional stratigraphic cross sections,
in that the y-axis values correspond to different depths in a reference well. These plots help to
identify decollement levels and stratigraphic intervals that are partially faulted out of wells
(Chatellier and Porras 2001).
In the format presented by Sanchez et al. (1997), MBPA is based on three or four ∆d/d plots
placed on a single page. Common to all these ∆d/d plots is the same reference well. On the left
side of each of the plots is the number of the reference well (top number), along with the number
of the well used for comparison (bottom number), as in (Fig. 13-13). The ∆d values are on the
y-axis, but in this case, the depths in the reference well are on the x-axis, regardless of its struc-
tural position relative to the comparison well. Thus, the reference well is present on all the plots.
Figure 13-13 Definitions for a Multiple Bischke Plot. The comparison of a reference well (top num-
ber on left), with depths plotted on the x-axis, to a comparison well (bottom number on left), forms
the basis for the MBPs. Unlike the ∆d/d method, the reference well can be on any structural level
and need not be the structurally higher well.
706 Chap. 13 / Growth Structures
Figure 13-14 A Multiple Bischke Plot of flooding surfaces from a field in Venezuela. Anomalous
point in the C6 unit, in all plots, suggests a possible stratigraphic miscorrelation in the reference well
(From Sanchez et al. 1997. Published by permission of Latinoamericano de Sedimentologia, Socie-
dad Venezolana de Geologos.)
become incoherent, and a rapidly thinning stratigraphic section on the flanks of the dome caused
reduced confidence in the well log correlations. The 3D seismic data were tied to off-structure
wells. The initial mapping, based on a computer-mapping program, indicated that two horizons
crossed each other in the up-dip direction, near the top of salt, which is impossible (Fig. 13-15).
How do you resolve a mapping and correlation problem if you cannot correlate the 3D seismic
data or the well logs with confidence?
The high bed dips and a rapidly thinning stratigraphic section along the flank of the diapir and
near Wells No. 1, 2, and 3 made the seismic and well log correlations inconclusive (Fig. 13-15). A
three-point problem, constructed from other wells that flank the diapir, results in an average bed
dip of 42 deg. Higher bed dips are likely to be encountered up-dip, as the strata conform to the salt
face. Best-guess correlations of the well log data resulted in an inaccurate pick for what is called
the 20 Horizon in the up-dip Well No. 1. Seismic picks were tied to the well control and maps
were generated. The computer-mapping program projected mapping horizons into the area of in-
coherent seismic data. The result was that the deeper 20 Horizon crosses the shallower 18 Hori-
zon up-dip of Wells No. 2 and 3. This is impossible. There is a significant distance between
where the seismic data deteriorate and where the log data for Wells No. 1, 2, and 3 are used to
generate ties to the seismic data, which were used in the final interpretation and maps.
Recognizing the problem, a solution must be determined. There are three questions to an-
swer: (1) Is the 20 Horizon correlated too high, or is the 18 Horizon correlated too low? (2)
Could there be unrecognized faulting that is not imaged in the 3D data set? (3) What does one do
The Multiple Bischke Plot Analysis 707
Figure 13-15 High bed dips on the flanks of diapirs may cause 3D seismic data to deteriorate, as
in this example. Inability to correlate the well logs may result in incorrectly correlated horizons in the
structurally higher wells. A projection, by a mapping program, of the 18 and 20 horizons up-dip into
the no-seismic-data zone caused the 18 and 20 horizons to cross, which is impossible. MBPA pro-
vides the means to resolve the structural and stratigraphic problems.
when both the seismic and well log data cannot be correlated with confidence? When confronted
with difficult problems of this kind, the MBPA may converge on a viable structural solution to
the problem (Bischke et al. 1999). Given their associated correlation problems, we want to com-
pare Well No. 1 with nearby Wells No. 2 and 3. These wells are down-dip of Well No. 1, have
thicker stratigraphic sections, are close to each other, and are easier to correlate than Well No. 1.
When using MBPA, experience has shown that comparing wells that have correlation prob-
lems to a type well can help resolve those problems. Wells No. 1, 2, and 3 are compared to a ref-
erence Well No. 4, which is an off-structure well. Figure 13-16a and b are two ∆d/d plots
constructed for Wells No. 2 and 3, using Well No. 4 as the reference well. These two plots indi-
cate that the growth rate gradually increases from the 1 Horizon down through the 20 Horizon in
both Wells No. 2 and 3. Gradual (monotonic) changes in growth on ∆d/d plots are a common
pattern observed in different tectonic environments (Bischke 1994b).
Figure 13-16c is a ∆d/d plot for Wells No. 1 and 4, using the original correlations. Well log
correlations are difficult below the 11 Horizon in Well No. 1. Notice where the correlation for the
20 Horizon appears in the ∆d/d plot. The plot is subject to one of three possible interpretations:
1. The growth increases dramatically between the 11 and 20 Horizons around Well No. 1;
2. Based on the discontinuity in the plot, a normal fault or unconformity, with a missing sec-
tion of about 600 ft, is present in Well No. 1 between the 11 and 20 Horizons; or
3. The 20 Horizon correlation in Well No. 1 is about 600 ft too high for a possible position of
the 20 Horizon. We obtain the value of about 600 ft by projecting the growth curve, defined
708 Chap. 13 / Growth Structures
(a)
(b)
(c)
Figure 13-16 MBPA using Well No. 4 in Fig. 13-15 as a reference well. (a) Growth plot using Well
No. 2 for comparison. Growth increases monotonically between the 1 and 20 Horizons. (b) Growth
plot using Well No. 3. Growth increases monotonically between the 1 and 20 Horizons. A compari-
son to Fig. 13-5 shows no significant apparent missing section in either Well No. 2 or 3. (c) Growth
plot for Well No. 1 located near the crest of the diapir. The plot suggests that about 600 ft of missing
section exists at the level of the 20 Horizon. As 600 ft of section is not missing in Wells No. 2 and 3,
or in other surrounding wells, the missing section in Well No. 1 is likely to result from a data miscor-
relation, rather than a fault. (From Bischke et al. 1999. Published by permission of the Gulf Coast As-
sociation of Geological Societies.)
The Multiple Bischke Plot Analysis 709
by the 1 to 11 data points, to beneath the 20 Horizon data point, and then determining the
difference in depths.
Using the data obtained from the Multiple Bischke Plots and knowledge of the local geol-
ogy, two of these interpretations can be eliminated and we can resolve the problem. We first con-
sider the increased growth interpretation. Although dramatic increases in growth are possible,
this increased growth is not observed in the nearby Wells No. 2 and 3 or in other nearby wells,
thus making the high growth interpretation very unlikely.
The interpretation based on a 600-ft fault or unconformity can be rejected for the following
reasons. A 600-ft unconformity should be observable on the down-dip portions of seismic pro-
files where bed dips are gentle or in growth plots constructed from other nearby, on-structure
wells. Furthermore, Wells No. 1, 2, and 3 are drilled from the same platform, and Wells No. 2
and 3 are deviated more to the south than Well No. 1. If a large 600-ft fault cuts Well No. 1, then
it could also cut Wells No. 2 and 3, or other nearby wells. Plots of the data shown in Fig. 13-16a
and b do not contain a large 600-ft discontinuity, nor do plots with other nearby wells. This
means that the ∆d/d plots do not indicate a large fault or unconformity below the 11 Horizon,
making interpretation No. 2 also highly unlikely. Also, the seismic data over the structure is of
reasonable quality. No large faults or unconformities are recognized on the seismic data near
Well No. 1.
Reexamination of the well logs suggests that the correlation of the 20 Horizon, near the bot-
tom of Well No. 1, is probably an incorrect correlation and that interpretation No. 3 is the likely
solution. Additional correlation work indicates that the miscorrelated horizon, which is about
550 ft below the 11 Horizon in Well No. 1, is not the 20 Horizon, but rather it is most likely the
14 Horizon. This change in correlation results in a completely new interpretation from the 14
Horizon through the 20 Horizon. The changes affect the fault interpretation, structure maps,
reservoir maps, and the upside prospect potential. The method results in a reasonable solution to
a difficult problem in an area where well log correlation and 3D seismic data alone could not re-
solve the problem.
(a)
(b)
Figure 13-17 (a) Map showing strike and dip on the Rob L horizon at locations on the flank of a
salt diapir in the northern Gulf of Mexico, USA. High bed dips on the flanks of salt structures cause
stratigraphic thinning and deterioration of the seismic data. In this environment, geoscientists have
difficulties attributing missing section to faults or to large unconformities. (b) The MBP for the struc-
turally higher reference Well No. 6 versus the five off-structure comparison wells. The missing sec-
tion above the Rob L sand in every well is interpreted to be due to a large unconformity. Fault J
produces about 250 and 340 ft of missing section in Wells No. 1 and 3 respectively, based on the
plot. (Published by permission of R. Bischke.)
Vertical Separation Versus Depth Method 711
Wells No. 1 and 3. The discontinuities represent missing section, either due to a fault in the
structurally lower well (compare to Fig. 13-5a) or due to downlap (compare to Fig. 13-6b). As
downlap is rare on the flanks of salt diapirs, it is more likely that these discontinuities result
from faulting in the structurally lower Wells No. 1 and 3. Correlation to other off-structure wells
confirms that the missing sections in Wells No. 1 and 3 on the MBP are due to Fault J, which is
noted on Fig. 13-17b.
On the flanks of salt diapirs, the amount of missing section typically increases up-dip as the
horizontal distance between the on-structure and off-structure wells increases (Chapter 4). The
∆d/d plots detect the amount of missing section caused by unconformities, as measured relative
to the two wells used for a given plot. If both wells penetrate the same unconformity, then the
amount of missing section determined from the plot will be the difference in the amounts of sec-
tions missing in the wells.
A positive discontinuity exists in every well plot in Fig. 13-17b. Furthermore, this missing
section occurs above the Rob L Horizon in every well plot. Thus, we attribute this missing sec-
tion above the Rob L to represent an unconformity. The results of the MBPA guided additional
correlation work and the interpretation when working with other structurally higher wells that
flank the diapir.
Therefore, the MBPA, using six wells on this structure, provides quantitative graphical evi-
dence for the following interpretation. The MBPA method (1) locates and defines the size (rela-
tive to certain wells) of the unconformity above the Rob L horizon, (2) defines the growth
history of the structure, (3) provides initial evidence for Fault J, (4) aids in distinguishing faults
from unconformities, and (5) helps keep the interpretation focused.
they cannot independently provide slip measurements (Chapter 7). Fault throw may be obtain-
able only after the construction of 3D maps, or from seismic sections that are perpendicular to
the strike of the fault surfaces (Chapter 8).
The VS/d plots measure the vertical component of fault displacement. We typically think of
vertical separation for normal and reverse faults. However, strike-slip faults also commonly exhibit
vertical displacements (Harding 1973; Wright 1991; Shaw et al. 1994), particularly at restraining
and releasing bends (Crowell 1974a and b) (Chapter 12). Also, because paleotopographic surfaces
are not planar and strike-slip faults exhibit vertical slip components, the method can apply to
growth sedimentation associated with strike-slip faults (Shaw et al. 1994).
Where subsurface data are sparse, the construction of subsurface maps and cross sections
typically involves the introduction of assumptions and interpretation. Fault slip components are
rarely obtainable in outcrop data, making fault slip determinations questionable even where
faults are subject to direct observation. However, VS is always obtainable directly from missing
or repeated section in well logs or from a single seismic profile (Chapter 7). Thus, we present a
method for studying fault displacements based on VS.
Method
Information on fault history and structural style is obtainable by plotting VS against the depth of
the faulted horizons at the footwall cutoff (Fig. 13-18). Every displaced horizon has a value of
VS (VS1, VS2, VS3, …VSn). For a VS/d plot, the VS is measured from the footwall horizon to
the hanging wall horizon. For a normal fault, the footwall horizon is projected across the fault
(Fig. 13-18). The measurement of VS in these cases is downward (increasing depth), and the
amount of VS is taken to be a positive value for purposes of the VS/d plot (Fig. 13-18). For a re-
Figure 13-18 A generic example of a post-depositional normal fault and its VS/d plot. A post-
depositional normal fault plots in the positive quadrant and has a flat slope. (Published by permis-
sion of R. Bischke.)
Vertical Separation Versus Depth Method 713
verse fault, the measurement of VS is upward (decreasing depth), and the amount of VS is taken
to be a negative value (Fig. 13-19).
If the fault exhibits growth, then sediments deposited coeval with fault motion will change
thickness across the fault surface, and the values of VS will change with depth. The slope of the
VS/d plots can be used to define fault growth, with higher slopes indicating higher growth rates.
A zero slope indicates a post-depositional (nongrowth) fault (Figs. 13-18 and 13-19). So, the
method readily defines the timing of fault movements.
The plots of VS versus depth document vertical displacements. If age data are available,
then it is possible to determine the displacement history within a time framework. However, the
plots cannot independently detect large rotational motions (see Chapter 12). Where applicable
with seismic data, use the workstation to convert the seismic data to depth before constructing
the plots. Typically, interval velocities are adequate, although better velocity–depth functions im-
prove the results.
Generic and Real Examples of Analysis of VS/d Plots
Vertical separation data taken along fault surfaces provides information on fault displacement his-
tory and structural style. Thus, in this section we concentrate on fault surfaces. The data required
are VS measurements and the depth to the footwall cutoff of each correlated horizon. We present a
number of generic and real examples of several common, and some less known, structural styles.
There are two basic types of faults, growth (expansion) faults and post-depositional (nongrowth)
faults (Thorsen 1963). The plots of VS versus depth document growth faulting or post-deposi-
tional faulting. Our discussion starts with simple examples, followed by more complex ones.
Figure 13-19 A generic example of a post-depositional reverse fault and its VS/d plot. A post-
depositional reverse fault plots in the negative quadrant and has a flat slope. (Published by permis-
sion of R. Bischke.)
714 Chap. 13 / Growth Structures
Growth Normal Faults. Sedimentary units change thickness across faults when the sedi-
mentation is concurrent with faults that have dip-slip components of displacement. Growth as-
sociated with faulting is common in areas that have rapid sedimentation rates (Suppe et al.
1992; Shaw et al. 1994), and thus normal, reverse, and strike-slip faults may exhibit growth.
One can distinguish growth faults from post-depositional faults in that the growth faults have
non-zero slopes on the VS/d plots. The VS/d method has a high resolution and can detect
growth where growth is not obvious in surface or subsurface data. The expansion faults com-
mon to deltas exhibit positive slopes on the VS versus depth plots, and the data are in the posi-
tive quadrant (Fig. 13-20).
Figure 13-21 is a VS/d plot for a growth fault in the northern Gulf of Mexico. Growth Faults
tend to seal through an interval of rapid growth, and any faults with large displacement tend to
seal. The larger the fault, the greater the probability that the fault will seal. Large displacements
on a fault can create a clay smear or barrier in the fault zone or can juxtapose permeable rocks
against shale (Brenneke 1995). In the United States, strata are typically dated by index fossil
Figure 13-20 A generic example of a growth normal fault and its VS/d plot. A growth normal fault
plots in the positive quadrant and has a positive slope. (Published by permission of R. Bischke.)
Vertical Separation Versus Depth Method 715
Figure 13-21 VS/d plot illustrating fault timing on a normal fault from the northern Gulf of Mexico,
offshore USA. Rapid growth on the fault, between 11,000 ft and 12,400 ft, enhances its trapping
capabilities within that depth range. (Published by permission of R. Bischke.)
taxa, such as Cibicides optima (abbreviated CIB.OP.). A review of Fig. 13-21 shows that the ear-
liest growth on the fault occurred just prior to CIB.OP.-10 time, with the older faulted interval
being pre-growth strata. The fault grew rapidly from CIB.OP.-10 time to about Bigenerina (3)-1
[BIG.(3)-1] time, and then became inactive. Just before Textularia (W)-2 [TEX.(W)-2] time, the
fault became active again and continued growing, at a decreasing rate, until Bigenerina (A)-1
[BIG.(A)-1] time. As the fault exhibits 400 ft to 900 ft of vertical separation and was growing
rapidly during CIB.OP.-10 time to BIG.(3)-1 time, the fault has a high probability of sealing
below the 11,000-ft level.
Growth Reverse Fault. Most geoscientists know that growth faulting is common to exten-
sional environments. Less well known is the case where deposition is concurrent with folding
and/or reverse faulting, and compressional growth structures form (Suppe et al. 1992; Tearpock
et al. 1994; Chapter 10 of this book). If, during compressional folding, the sediments flood across
and over the rising structure, then a compressional growth structure forms (Fig. 13-22). Plots of
VS versus depth for growth reverse faults are distinguished from those of growth normal faults
in that they exhibit negative slopes and plot in the negative quadrant (Fig. 13-22).
Complex Growth Structures. Complex growth structures exist where more than one fault,
or slip surface, is present, or where there was more than one period of deformation. Complex
structures occur in all tectonic environments. The hybrid inversion structures that occur in
Brunei (James 1984), the North Sea (Mitra 1993), Venezuela (Link et al. 1996), and elsewhere
are readily recognizable on the VS/d plots. Inversion structures originate as normal fault
structures that are later subject to reverse fault displacements. Growth across the faults may be
noticeable during both the extensional and compressional phases. On growth plots, the VS data
initially plot in the negative and then in the positive quadrant (Fig. 13-23). As these structures
716 Chap. 13 / Growth Structures
Figure 13-22 A generic example of a growth reverse fault and its VS/d plot. A growth reverse fault
plots in the negative quadrant, with a negative slope. (Published by permission of R. Bischke.)
Figure 13-23 A generic example of an extensional, growth inversion fault and its VS/d plot. An ex-
tensional growth inversion fault plots in both the positive and negative quadrants, with a positive
slope. (Published by permission of R. Bischke.)
Vertical Separation Versus Depth Method 717
plot in different quadrants and have positive slopes, inversion structures are distinct and easily
recognized from other fault styles.
Inversion structural styles have been easily confused with strike-slip fault structural styles.
In Chapter 12, we discuss methods and procedures for documenting horizontal displacements,
which are critical in determining strike-slip faulting. Plots of VS versus depth aid in recognizing
inversion structures. In the hanging wall portions of inversion structures, the shallow stratigraphy
is thin relative to its correlative footwall units, whereas the deeper stratigraphy is thick relative to
its correlative footwall units (Fig. 13-24 and Mitra 1993). Before the recognition of inversion
structures, and on older seismic data that poorly imaged the listric normal fault, geoscientists
commonly interpreted well log and seismic data incorrectly. We use a depiction of an inversion
structure in Venezuela as an example. First, at the shallow structural level of the stratigraphic
unit, geoscientists recognized that the thin, shallow-water environment, footwall stratigraphy
correlated to the thicker, deep-water environment strata in the hanging wall block (Fig. 13-24b).
Second, the thickness of the sediments and the environmental conditions reverse at the shallower
structural level, with the footwall section being thicker relative to the hanging wall section. The
deeper, and once downthrown, units were subsequently thrust back up, along the pre-existing
normal faults. This reverse motion thrusts the deeper water units to a structurally higher level
than the correlative thin shallow water units (Fig. 13-24b). Before the recognition of inverse
THIN
THICK
BEFORE INVERSION
(a)
THIN
THICK
AFTER INVERSION
(b)
structures, these complex and confusing relationships were interpreted to result from large
strike-slip displacements. Only strike-slip displacements were considered capable of creating
this confusing geometric style. Thus, inversion structures may be misidentified as “flower”
structures. To help distinguish inversion structures from flower structures, examine the correla-
tion markers on VS/d plots. If the markers plot in both the negative and positive quadrants and
have positive slopes, then the structure is most likely an inversion structure.
We can distinguish between an inversion structure and another structural style. Compres-
sional inversion structures are present where the displacements on reverse faults invert and be-
come normal. The positive and negative data points, combined with negative slopes, indicate
inversion on a VS/d plot (Fig. 13-25).
Downward-Dying Growth Faults. Another important but problematic structure is the down-
ward-dying growth fault, which commonly forms in the extensional environment (Bischke and
Suppe 1990b; Tearpock and Bischke 1991). Some geoscientists err in being unaware that a nor-
mal fault may die downward. An extension of a fault with depth is unreasonable if supporting
data does not exist. However, this fault style can be recognized because it plots on a VS/d dia-
gram in the positive quadrant and exhibits positive slopes at shallow depths and negative slopes
at greater depths (Fig. 13-26).
Some faults in deltaic areas may decrease in displacement in all directions away from a cen-
tral point. If the vertical separations on these faults surfaces are contoured into a map of fault
displacements, then these faults exhibit a more or less elliptical displacement pattern (Barnett et
al. 1987). The maximum displacements occur at the center of the ellipse and the fault dies out in
Figure 13-25 A generic example of a compressional, growth inversion fault and its VS/d plot. On a
compressional, growth inversion structure, the thin upthrown section is later downthrown, juxtapos-
ing thin sections in the hanging wall against thicker sedimentary sections in the footwall. A com-
pressional inversion fault plots in both the positive and negative quadrants and has a negative
slope. (Published by permission of R. Bischke.)
Vertical Separation Versus Depth Method 719
Figure 13-26 A generic example of a downward-dying, growth normal fault and its VS/d plot. The
displacements on downward-dying faults decrease with increasing depths. A downward-dying
growth fault plots in the positive quadrant. It has a positive slope at shallower depths and a negative
slope at greater depths. (Published by permission of R. Bischke.)
every direction, upward, downward, and laterally. In Chapter 11 we present a model to explain
these unusual relationships in a syntectonic depositional environment.
Figure 13-27 shows a map of a downward-dying growth fault, with accompanying seismic
section, and a VS/d plot (Rabbit Island, coastal Gulf of Mexico, USA). The Rabbit Island fault
exhibits 450 ft of missing section in a well at the 5500-ft level and about 70 ft of missing section
in a well at about the 11,000-ft level. The structural question is, what happens to the fault below
11,000 ft? In order to determine the downward extent of the fault, based only on well control,
there must be wells that go deeper than 11,000 ft in the area where the fault could reach this
depth. With seismic data, one must approach the interpretation with caution. At times, inter-
preters do not pay attention to the displacements when picking faults, but rather they concentrate
on interpreting the fault location on the seismic sections. If an interpreter does not realize that he
or she is dealing with a downward-dying fault, an interpretation can be generated of a through-
going fault extending to depth. This apparent throughgoing fault may set up prospects or perhaps
divide producing reservoirs. If the fault in our example is absent beneath the 11,000-ft level, then
no fault exists to trap hydrocarbons beneath that level. Or, if a reservoir exists just beneath
11,000 ft, then the partitioning of the reservoir is incorrect. In either case the interpretation is
wrong and can have serious implications to any exploration or development activities. Geoscien-
tists can construct VS/d plots to predict fault displacement and to evaluate whether they are deal-
ing with a downward-dying fault, and to determine at which level the downward-dying fault
loses its displacement (Tearpock et al. 1994). The Rabbit Island fault apparently dies out near the
11,000-ft level, as shown on the seismic line in Fig. 13-27c, and it cannot trap hydrocarbons be-
neath that depth.
720 Chap. 13 / Growth Structures
(a)
(b)
Figure 13-27 (a) Map, (b) VS/d plot, and (c) seismic line showing a downward-dying growth fault
in Rabbit Island Field, coastal Gulf of Mexico, USA. Vertical separation decreases with increasing
depth. The trapping capabilities along the fault surface diminish at about the 11,000-ft level. (Maps
and seismic line published by permission of Texaco, Inc.; VS/d plot published by permission of R.
Bischke.)
Conclusions 721
(c)
Figure 13-27 (continued )
Downward-dying growth faults are common in many areas, and seismic displays of them
from the northern Gulf of Mexico and Brunei are in Chapter 11. In other areas, some faults may
extend to depth, whereas other faults die downward. When dealing with potential hydrocarbon
traps on downward-dying faults, care must be exercised to be sure that the potentially trapping
fault will offset the objective horizon (Tearpock et al. 1994). The VS/d method or MBPA can aid
in the recognition and documentation of the trapping capability of downward-dying growth
faults.
CONCLUSIONS
Reliable correlation data are obtainable in many areas, but the data may deteriorate in key areas.
The ∆d/d, MBPA, or VS/d methods, when integrated with geologic and geophysical data, can re-
solve difficult structural and stratigraphic problems where 3D seismic and well log correlation
data are ambiguous (Sanchez et al. 1997). The methods are easy to use, require little additional
effort, and can eliminate the guesswork that could lead to incorrect structural interpretations and
722 Chap. 13 / Growth Structures
maps that could result in dry holes. Where correlation problems result from complicated struc-
tural features, the Multiple Bischke Plot Analysis can resolve these problems during the initial
stages of a project, prior to constructing subsurface maps and generating prospects. This can
help eliminate the need to remap the structure and/or reinterpret the seismic and well log data at
a later stage in a project.
Additionally, as the methods are easy to use and rapid to apply, they serve as excellent
checks on any interpretation involving structures or correlations. Often, after finishing a project,
interpreters assume that their interpretations and maps are correct and do not subject their inter-
pretations to consistency checks (Tearpock et al. 1994). This can result in interpretations that are
inconsistent with the interpreter’s own correlation data, maps, or cross sections. Growth tech-
niques are an excellent consistency check and introduce a relative age factor into interpretations.
Managers may find these techniques useful during prospect reviews and when analyzing
prospects and the petroleum system.
The ∆d/d and MBPA methods have a resolution of about one part in a thousand when using
well log data, so the methods can often identify disconformities, subtle systems tract boundaries,
or small growth faults (Bischke 1994b). The techniques can distinguish faults from unconformi-
ties where bed dips are high and where the seismic data is incoherent. If a well log correlation is
the probable cause of the problem, then the method can direct geoscientists to the correct corre-
lation, thus eliminating guesswork. Analysis of growth history often leads to viable structural so-
lutions to complex problems.
Syntectonic sedimentation can occur in any tectonostratigraphic regime, not solely in the
extensional, normal fault regime. In areas where the sedimentation rates keep pace with the tec-
tonic uplift rates, structures formed by compressional and strike-slip faults can exhibit growth
(Suppe et al. 1992; Shaw et al. 1994). These growth environments allow for the determination of
fault timing that influences migration and petroleum system studies. Lastly, in areas where struc-
ture is complex or ambiguous, growth plots provide the results to determine structural style and
resolve interpretation problems. Any interpretation of a stratigraphic sequence or structure must
be consistent with the correlations.
Chapter 14
INTRODUCTION
Two key terms, isochore and isopach, are often used synonymously in the petroleum industry as
measures of thickness, but they are different. An isochore map (Fig. 14-1a) delineates the true
vertical thickness of a stratigraphic interval, whereas an isopach map (Fig. 14-1b) illustrates
the true stratigraphic thickness of a stratigraphic interval. These two terms are often confused
with respect to their geologic meaning. It is vital for both exploration and development work that
the correct meaning and, more importantly, the correct application of these two thicknesses be
understood.
An isochored or isopached unit may be as small as an individual sand only a few feet thick,
or as large as several thousand feet thick and encompassing a number of stratigraphic units. An
isopach map is extremely useful in determining the stratigraphic framework or the structural re-
lationship responsible for a given type of sedimentation, and for recognizing paleohigh areas.
The shape of a basin, the position of the shoreline, areas of uplift, and under some circumstances
the amount of vertical uplift and erosion, can be recognized by mapping the variations in thick-
ness of a given stratigraphic interval (Bishop 1960).
Isochore and isopach maps are used for a number of purposes by the petroleum geologist,
including (1) depositional environment studies, (2) genetic sand studies, (3) growth history
analyses, (4) depositional fairway studies, (5) derivative mapping, (6) the history of fault move-
ment, and (7) calculation of hydrocarbon volumes.
In this chapter, we discuss several different types of isochore and isopach maps important to
the evaluation of petroleum potential. These include interval isopach maps and net sand and net
pay isochore maps. An interval isopach map delineates the true stratigraphic thickness of a
specific unit or units.
723
724 Chap. 14 / Isochore and Isopach Maps
(a)
(b)
Figure 14-1 (a) An isochore map delineates the true vertical thickness of a stratigraphic interval,
such as a rock unit containing a reservoir. (b) An isopach map delineates the true stratigraphic thick-
ness of a stratigraphic interval. The same dipping stratigraphic unit is used in both (a) and (b), with
the same edge-of-map boundaries. Note different thickness values assigned to the isochore map
versus the isopach map of the same unit.
Introduction 725
An isochore map may be based on the total vertical thickness of a specific unit or based on
an aggregate vertical thickness of a particular rock type within a stratigraphic unit. For example,
a given unit may consist of interbedded permeable and impermeable strata. An isochore map can
be made for the total unit consisting of both permeable and impermeable strata, or an isochore
map can represent the aggregate vertical thickness of only the reservoir-quality rock within that
stratigraphic unit. For simplicity and brevity in this chapter, we use the term net sand isochore
map to refer to a map of only the reservoir-quality rock. Therefore, a net sand isochore map
represents the aggregate vertical thickness of reservoir-quality rock present in a particular strati-
graphic interval, which is illustrated in Fig. 14-2. The techniques and calculations to derive verti-
cal thickness are explained in detail in this chapter. The fluid contained in a stratigraphic unit
may be hydrocarbons or water, or any combination of the two. Figure 14-3 shows a net sand iso-
chore map of the 10,500-ft Sand in Golden Meadow Field, Lafourche Parish, Louisiana, USA. A
net pay isochore map delineates the aggregate vertical thickness of reservoir-quality rock that
contains hydrocarbons (gas, oil, or both). An example is shown in Fig. 14-5.
Net sand and net pay isochore maps of subsurface units are usually prepared from well log
data, whereas interval isopach maps may be constructed from well log data and seismic data,
where coverage is adequate. As with structure maps, the completeness and accuracy of an iso-
chore or isopach map depends upon the amount and accuracy of data available. Even in isochore
and isopach mapping, we cannot get away from the importance of log correlation work. Well log
data, particularly correlations, should be studied very carefully in order to prepare an accurate
and precise isochore map.
For volumetric reserve calculations, we are interested in obtaining the volume of a reservoir
(solid material plus fluid-filled pore space). In this book, we use the volume unit acre-foot,
which is standard practice within the United States. One acre-foot is equivalent to 1233.48 cubic
meters, and to 0.123348 hectare-meters. To many people, an acre-foot is an abstract measure-
ment, but the concept is relatively simple. One acre-foot can be defined as that volume of rock
plus fluid contained in an area one acre in size, with a thickness of one foot. How big is an acre?
There is a very easy way for some people to visualize the size of an acre. One acre contains just
Figure 14-2 Net sand consists of porous reservoir-quality rock. All nonreservoir-quality rock is ig-
nored. (From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
726 Chap. 14 / Isochore and Isopach Maps
Figure 14-3 Portion of the net sand isochore map of the 10,500-ft Sand in Golden Meadow Field,
Lafourche Parish, Louisiana, USA. (Published by permission of Texaco, USA.)
Introduction 727
728 Chap. 14 / Isochore and Isopach Maps
about the same area as an American football field from goal line to goal line. A football field 300
ft long and 160 ft wide is equal to 48,000 sq ft, whereas one acre is equal to 43,560 sq ft. If we
fill, with oil, a box that is one foot deep and the size of an American football field, the total vol-
ume of oil is just about equal to one acre-foot. In terms of barrels of oil, there are 7758 barrels
(1233 cubic meters) of oil in one acre-foot. However, a reservoir volume of one acre-foot is actu-
ally comprised of solid material plus fluids that occupy the pore space. So a calculation of oil in
place in a reservoir must take into account pore space and water saturation, as well as other fac-
tors. In creating a net pay map, we assume that within a net sand interval that contains hydro-
carbons, all pore spaces are filled with hydrocarbons. The later calculation of hydrocarbons-in-
place takes account of water saturation. In most cases, the structurally lower limit of a reservoir
is a hydrocarbon/water contact, typically defined on the basis of parameters related to economic
productivity, such as porosity and water saturation.
SAND–SHALE DISTRIBUTION
Most individual rock bodies do not consist exclusively of permeable rock; shale and other imper-
meable rock material are commonly distributed throughout the rock body as interbedded shale
or impervious (tight) zones. The percentage and distribution of shale members or impervious
zones varies widely within rock units. Net sand and net pay isochore maps are drawn on net ef-
fective sand (porous and permeable sand) only; therefore, shale and other nonreservoir-quality
rock must be subtracted from the total interval to determine the net effective sand for isochore
mapping.
In a given oil or gas well, the net effective sand to be used for isochore mapping is normally
determined by detailed analysis of 5-in. well logs, supplemented by available core analysis. In
Chapter 4, in the section Annotation and Documentation, we outline a method for annotating the
percentage and distribution of net sand and impermeable layers that are present within a particu-
lar productive unit (Fig. 4-47). Once the net sand is determined for each well, a net sand map can
be prepared for that sand. The aggregate of net sand for any particular well may contain water or
hydrocarbons; net pay is that portion of the net sand that contains hydrocarbons.
For a net sand or net pay map, the gross, or overall, interval to be mapped extends from the
top of porosity to the base of porosity within the productive unit. Within that interval, the
amounts of net sand and net pay vary with location. The net/gross ratio is the amount of net
sand divided by the thickness of the gross interval, as determined from well logs and core analy-
ses. The net/gross ratio within a productive unit can be mapped, based on log and core data, and
used to estimate net sand and net pay at selected locations within the unit, as described in a later
section in this chapter.
GAS
OIL
WAT
ER
Net Pay Isochore Map Construction. The construction of a net pay isochore map for a bot-
tom water reservoir requires a structure map on the top of porosity for a given productive unit,
net pay values for each well in the reservoir, and the depth of the hydrocarbon/water contact. The
following procedure is used to construct the net pay isochore map for a bottom water reservoir
(Fig. 14-5).
1. Post the net pay values for each well on a base map. If deviated wells are included, the net
pay values must be corrected to true vertical thickness.
2. Overlay the isochore base map on the top-of-porosity structure map for the reservoir, and
draw the outer limit of the hydrocarbon-bearing reservoir. The outer limit may be any
boundary, or combination of boundaries, such as an oil/water contact, fault, pinchout, or
permeability barrier. Where this outer limit of the productive reservoir area is a hydrocar-
bon/water contact, it becomes the zero contour line on the net pay isochore map, as shown
in Fig. 14-5. In this case, the outer limit of the reservoir is an oil/water contact at a depth of
–10,250 ft. However, if a boundary of part of a reservoir is a more or less vertical surface,
such as a vertical fault or a vertical salt face, then the amount of net pay adjacent to this sur-
face will exceed zero.
3. Contour the net pay isochore map, which is contained within the area outlined by the zero
line on the base map. Be sure to honor all posted net pay values. The net pay contours may
generally conform to the structure contours. Net pay contours are commonly drawn as being
equally spaced. However, due to variations in net sand thickness within the reservoir, the iso-
chore contours need not be equally spaced. Figure 14-5 is a net pay map of a reservoir in
which the net oil contours reflect the structure in a general way, yet the contours are propor-
tionally spaced due to variation in net sand within the productive unit. If the well control is
limited, additional points of contour control may be obtained by using a method called walk-
ing wells or by using a net/gross ratio map, each of which allows us to estimate net pay at
chosen points in a reservoir. These methods are explained in detail later in this chapter.
Always indicate the maximum or minimum thicknesses of pay within the area bounded by a
maximum or closed minimum net pay contour (Fig. 14-5), as this information is necessary for
volumetric calculation after planimetering of the net pay map.
GAS
OIL
WAT
ER
Figure 14-7 Structure maps, cross section, and net oil isochore map for an edge water reservoir.
732 Chap. 14 / Isochore and Isopach Maps
Edge water reservoirs are obviously more complex than bottom water reservoirs. An edge
water reservoir can be extremely complex if it contains both oil and gas and is cut by one or
more faults. When mapping a reservoir cut by faults, consideration may be given to mapping one
or more fault wedges in addition to mapping the hydrocarbon wedges above water. The result
may be a complex isochore map. Fault wedges are discussed in detail later in this chapter.
An understanding of the reservoir type and configuration is very important in such decisions
as the location of development wells, completion practices, and production plans. Take a few
minutes and review Fig. 14-6, especially the areas of multiple wedge zones.
The generally accepted method for construction of a net hydrocarbon isochore map for an
edge water reservoir is called the Wharton Method, after J. B. Wharton (1948). The data
needed to construct a net hydrocarbon isochore map for this type reservoir are
Net Pay Isochore Map Construction for a Single-Phase Reservoir. We first outline the
procedure for construction of a net hydrocarbon isochore map for an edge water reservoir con-
taining a single type of hydrocarbon. For this example, we consider the rock type to be sandstone
and the hydrocarbon to be oil. The procedure is illustrated in Fig. 14-8.
1. Start with a base map with all the well control spotted.
2. Place the base map over the structure map on the top of the unit (top of porosity) (Fig. 14-8a),
and trace the outer limit of the productive reservoir. In this example, the limit is the
oil/water contact on the top of the unit, and it becomes the zero line on the net oil isochore
map, as in the previous example. The zero line is shown in Fig. 14-8b. From this point for-
ward, we refer to the base map as the net oil isochore map.
3. Place the net oil isochore map over the structure map on the base of the unit (base of poros-
ity) (Fig. 14-8c), and trace the oil/water contact on the isochore map using a dashed line.
This dashed line represents the inner limit of water for the reservoir. Within the area inside
this dashed line, oil fills the productive unit (Fig. 14-8d), and this area is referred to as the
full thickness area. The intersection of the oil/water contact with the top and base of the
unit outline the wedge zone.
4. Post net pay values for all wells within the reservoir, corrected to true vertical thickness, on
the net oil isochore map.
5. In the full thickness area, the entire net sand is filled with oil, so the net oil in this area
equals the amount of net sand, as interpreted on the net sand isochore map. Notice that the
net sand contours are based on all wells, within and outside the reservoir. Therefore, to con-
tour the full thickness area, place the net oil isochore map over the net sand isochore map
as shown in Fig. 14-8e, and trace the contours within the dashed-line area onto the net oil
isochore map. Label the maximum thickness of oil (63 ft) within the area enclosed by the
maximum net pay contour (60 ft). The full thickness area of the net oil isochore map is now
finished, as illustrated in Fig. 14-8f.
6. The next step is to contour the oil wedge. The wedge zone is the area between the oil/water
contact on the top of the unit and the oil/water contact on the base of the unit, as shown in
Basic Construction of Isochore Maps 733
(a)
(b)
Figure 14-8 (a) Overlay a net oil base map on the structure map of the top of sand (top of poros-
ity). The oil/water contact is traced on the overlay. (b) The oil/water contact becomes the zero con-
tour line on the net oil isochore map. (c) Overlay the net oil isochore map on the structure map of the
base of sand (base of porosity), and trace the oil/water contact as a dashed line. (d) Isochore base
map delineating the two major areas comprising the net oil isochore map: (1) oil wedge from the
zero line to the inner limit of water (oil/water contact on the base of the sand), and (2) the area of full
hydrocarbon thickness. (e) To contour the full thickness area, superimpose the net oil isochore map
on a net sand isochore map and trace the net sand contours inside the inner limit of water (dashed
line). (f) Net oil isochore map with contours drawn for the oil-filled area. (g) All full thickness area
contours that intersect the inner limit of water must connect through the wedge with contours of
equal value (see text for procedure). (h) Completed net oil isochore map with important points of iso-
chore construction listed. (From Tearpock and Harris 1987. Published by permission of Tenneco Oil
Company.)
734 Chap. 14 / Isochore and Isopach Maps
(c)
(d)
(e)
Figure 14-8 (continued)
Basic Construction of Isochore Maps 735
(f)
(g)
(h)
Figure 14-8 (continued )
736 Chap. 14 / Isochore and Isopach Maps
Fig. 14-8d and f. The oil wedge overlies a water wedge (see cross section in Fig. 14-10).
All well data in the wedge must be honored. In the full thickness area, net sand distribution
controls the net pay contours. However, in the oil wedge, the structural attitude of the pro-
ductive unit and the distribution of impermeable rock within the unit, influence the position
of net pay contours. The contours conform in general to the structure contours, but they are
not necessarily equally spaced (variations in contour spacing are discussed below). As the
first step in contouring the wedge, draw the net pay contours from the full thickness area
into the wedge. Notice in Fig. 14-8g that the full thickness contour lines turn abruptly at the
inner limit of water, in the direction of increasing sand thickness (in the direction of the
next net sand contour of higher value). They connect through the wedge with contours of
the same value elsewhere in the full thickness area. Lastly, draw the net pay contours in the
oil wedge that do not correspond to contours within the full thickness area. Honor the well
data within the wedge, and more or less equally space contours where data is lacking (Fig.
14-8h). If the dip of the productive unit is more irregular than in this example, then that
variation in structure may be considered in contouring the wedge.
The completed net oil isochore map is shown in Fig. 14-8h. This figure highlights five im-
portant points of the net pay isochore map construction. We recommend that you leave the dashed
inner limit of water on the net pay isochore map. It allows others to review and verify that you
have used the water contact on the base of porosity correctly in creating the net pay map.
The method of connecting the full thickness contours to those in the wedge zone is
extremely important and deserves special attention. Why can we not extend the net pay contours
in the full thickness area on trend, past the inner limit of water and into the wedge? Look at
Fig. 14-9, which is similar to Fig. 14-8g. Let’s discuss the construction of the easternmost 50-ft
contour in the figure. Why must this contour line sharply change direction at the inner limit
of water on the net oil isochore map, rather than continue straight into the wedge zone? Figure
14-10 is a diagrammatic cross section along the 50-ft net sand contour. Everywhere along the
cross section are exactly 50 ft of net sand. In the portion of the reservoir that is up-dip to the
Figure 14-9 Full thickness net pay contours make an abrupt turn at the inner limit of water, toward
the next net sand contour of higher value. (Modified from Tearpock and Harris 1987. Published by
permission of Tenneco Oil Company.)
Basic Construction of Isochore Maps 737
Figure 14-10 Cross section along the 50-ft net sand contour line in Fig. 14-9.
oil/water contact on the base of porosity (inner limit of water), oil fills the entire 50 ft of net
sand. However, a point one foot down-dip of the inner limit of water is within the wedge zone,
where the sand contains both oil and water. Therefore, anywhere outside the inner limit of water,
both oil and water are present and, since the total net sand is still 50 ft thick, there must be less
than 50 ft of oil. The 50-ft net pay contour, therefore, cannot continue along the 50-ft net sand
contour down-dip of the inner limit of water.
Where must the 50-ft contour from the full thickness area be drawn in the wedge zone? It
must be drawn through an area of 50 ft of net oil in the wedge zone. This area exists only where
the net sand is greater than 50 ft. In Fig. 14-8e, notice that the net sand increases in thickness
west of the 50-ft contour to a maximum of 63 ft. Therefore, the 50-ft contour, at its intersection
with the inner limit of water, must turn sharply toward the area of thicker sand. Since contour
lines must close, the contour connects to the west to connect with the other 50-ft contour in the
full thickness area (Fig. 14-8g).
This procedure must be undertaken for all contour lines contained within the full thickness
area of the net pay isochore map. The correct application of this technique is most important. If
the full thickness area net pay contours are carried incorrectly into the wedge zone, the volume
of hydrocarbons determined for the reservoir will be overestimated.
Figure 14-11a and b present a summary of the foregoing method for preparing a net pay iso-
chore map for an edge water reservoir containing one type of hydrocarbon.
Comprehension of, and commitment to, the foregoing procedure will help you avoid the
most common pitfall in net pay mapping, which is the mapping of more net pay than net sand
that is filled with hydrocarbons. In the full thickness area, a net pay contour corresponds exactly
to a net sand contour. In the hydrocarbon wedge, a net pay contour corresponds to that same
amount of net sand above the water level. Therefore, in the wedge, a net pay contour of a given
value must be within an area where the total net sand is of greater value. Always overlay and
compare the net pay contour map to the net sand contour map in order to check that the positions
of the net pay contours are reasonable.
Net Pay Isochore Map Construction for a Reservoir Containing Oil and Gas. Two
methods may be used to estimate the volumes of oil and gas in a reservoir containing both types
738 Chap. 14 / Isochore and Isopach Maps
(a)
(b)
Figure 14-11 (a) - (b) Summary of method for constructing a net hydrocarbon isochore map for an
edge water reservoir containing one hydrocarbon (oil or gas). (Modified from Tearpock and Harris
1987. Published by permission of Tenneco Oil Company.)
of hydrocarbons. The simplest and quickest method is to construct a total net hydrocarbon iso-
chore map and a net gas isochore map, calculate the volumes of each, and subtract the gas vol-
ume from the total hydrocarbon volume to determine the oil volume. This method is appropriate
when only one lease owner is involved or when only the total estimated volumes of oil and gas
are required (there being no interest in the actual distribution of oil and gas within the reservoir).
Where a reservoir underlies two or more separate ownerships, it is very important to know the
estimated volumes of gas and oil under each lease. In this case, net gas and net oil isochore maps
must be constructed, preferably using the procedure outlined in this section.
Construct the net gas isochore map first. Draw the basic maps used in the Wharton method,
which are the structure map on top of porosity, structure map on base of porosity, and net sand
isochore map. Using these maps, construct the net gas isochore map as shown in Fig. 14-12a and
b. The net gas isochore map is constructed using the same procedure explained in the previous
Basic Construction of Isochore Maps 739
(a)
(b)
Figure 14-12 (a) - (b) Summary of procedure to construct a net gas isochore map for a reservoir
containing both oil and gas. (Modified from Tearpock and Harris 1987. Published by permission of
Tenneco Oil Company.)
section on edge water reservoirs containing one type of hydrocarbon. The only difference in this
case is that the gas/oil contact defines the down-dip, outer limit of the gas reservoir, whereas in
the previous cases the limit was determined by a hydrocarbon/water contact.
The last map to be constructed is the net oil isochore map. This map differs from the previ-
ous maps in that it has two wedge zones, an inner wedge zone (gas/oil) and an outer wedge zone
(oil/water), as shown on the cross section in Fig. 14-12a. The outer oil wedge and any full thick-
ness areas are constructed using the Wharton method, as previously discussed; the inner oil
wedge requires additional steps.
Figure 14-13a shows a partitioned map (without wells) for the oil reservoir, with an inner
wedge and an outer wedge of oil, and an area of full thickness in between. The fluid contacts on
the structure maps are used to determine the zero net pay limits and dashed wedge limits by trac-
ing them onto the net oil map.
First, contour the area containing a full thickness of oil. Post the net oil values next to each
well and overlay the isochore base map onto the net sand isochore map, as shown in Fig. 14-13b.
Trace the net sand contours on the net oil map only within the full thickness area of oil, as shown
740 Chap. 14 / Isochore and Isopach Maps
(a)
(b)
(c)
Figure 14-13 (a) Outline of oil isochore map showing the inner and outer wedge zones and the full
thickness area. (b) Overlay of net oil isochore map on the net sand isochore map. The contours in
the area of full oil thickness are equal to the net sand contours. (c) Full thickness area is contoured.
(d) Overlay of net gas and net sand isochore maps used to aid in the construction of the inner oil
wedge contours. (e) Completed net oil isochore map. (From Tearpock and Harris 1987. Published by
permission of Tenneco Oil Company.)
Basic Construction of Isochore Maps 741
(d)
(e)
Figure 14-13 (continued)
in Fig. 14-13c. Indicate the maximum thickness of oil in each of the areas enclosed by the 60-ft
net pay contours. This portion of the isochore map is now complete.
We now contour the inner oil wedge on the map, to be followed by contouring of the outer
oil wedge. By referring to the cross section in Fig. 14-12b, we see that within the inner oil
wedge, the sum of net oil-filled and net gas-filled sand equals the total net sand. Therefore, in
order to determine the amount of net oil in the inner wedge zone, the following procedure is
used. Overlay the net gas isochore map on the net sand isochore map, and note each location
where the contours on the two separate maps cross. The net oil sand value at each contour inter-
section is equal to the difference in values of the two contours. For example, at point C in Fig.
14-13d, the 30-ft contour line on the net gas isochore map crosses the 50-ft contour line on the
net sand isochore map. By subtracting the 30 ft of net gas from the 50 ft of net sand, a value of
742 Chap. 14 / Isochore and Isopach Maps
20 ft of net oil is obtained for this point. Take a minute and review the data at point D. As indi-
cated, a known value is established wherever a contour line on the net gas isochore crosses a
contour line on the net sand isochore. The net oil sand value at each intersection is the difference
in values of the two contours. Figure 14-13d shows 48 points of control, plus data from four
wells, to aid in contouring the inner wedge zone of the oil isochore map. Overlay the net oil map
on the net gas and net sand maps, mark selected points with the corresponding value of oil, and
then contour the inner oil wedge. The net pay contours at the edge of the full thickness area turn
abruptly towards the thicker net sand values (Fig. 14-13e), just as described in the contouring of
a single-phase reservoir.
When contouring the inner oil wedge, the net oil isochore map should be overlain on both
the net gas and net sand isochore maps. This allows mapping the inner oil wedge with the con-
straint, and thus the assurance, that the sum of the net gas and net oil does not exceed the total
net sand. This is one of the most complex, and therefore most difficult, areas to construct within
an isochore map.
Finally, contour the outer wedge precisely as described in the previous section on contour-
ing the wedge in a single-phase reservoir (Fig. 14-8h). We now have constructed the inner and
outer wedges and the full thickness area for the oil isochore. The completed net oil isochore map
is shown in Fig. 14-13e.
(a)
(b)
(c)
Figure 14-14 (a) Net/gross ratio map based on well control. (b) Overlay of partially completed net
gas map on the net/gross ratio map. (c) Completed net gas isochore map. (Published by permission
of D. Tearpock.)
744 Chap. 14 / Isochore and Isopach Maps
gross interval thickness above the water level as the difference in depth between the top of
porosity and the water level. Multiply this calculated gross interval by the net/gross ratio to esti-
mate the net sand above the water level, which is equal to net pay at the selected point.
We use the maps in Fig. 14-14 as an example of the technique, applied to an edge water gas
reservoir with a gas/water contact at –8350 ft. The net/gross mapping technique begins after con-
struction of the outline of the net pay map and contouring of net pay within the full thickness
area (Fig. 14-14b). As the first step in constructing a net/gross ratio map, calculate the net/gross
ratio for wells within and outside the reservoir. Figure 14-14a shows the net/gross ratio calcula-
tion beside each well in the area of study. Contour the values using a contour interval of 5 per-
cent, or 0.05. Then estimate net pay at selected points within the wedge. Figure 14-14b is a
partially completed net pay isochore map of the reservoir, shown as an overlay on some contours
in the net/gross map. For efficiency, select estimated net pay points to lie directly on the
net/gross contours, such as points (1) through (9). As an example, use point (3) on the 50 percent
contour. Calculate the gross interval thickness above the water level at point (3) by taking the
difference in depth between the gas/water contact at –8350 ft and the top of porosity at point (3),
which is –8310 ft on the structure map (not shown). The gross interval thus equals 40 ft. Then
estimate the net pay by using the following calculation.
Let’s say we wish to construct a net gas isochore map, with a 10-ft contour interval, for a
sandstone reservoir with limited well control. We want additional control in the gas wedge, so
we decide to walk a well through the wedge and estimate the amount of net gas at selected
points. Any well to be walked can be located in the reservoir itself, or even down-dip of the
hydrocarbon/water contact and thus outside the reservoir. The key point in walking a well is to
choose a well that can be walked parallel to the nearest contour line on the net sand isochore
map. This is a critical point because when walking a well through the wedge, the assumption is
made that the amount and distribution of net sand, as seen in the well log, are the same within
the reservoir in a direction parallel to the net sand contours. For example, if a well has 50 ft of
net sand, it will fall on the 50-ft total net sand contour line. Construction of this 50-ft contour
line assumes that exactly 50 ft of net sand exists all along this contour line, not just at the well
location (Fig. 14-15). If a series of wells were drilled along this contour line, each well should
encounter exactly 50 ft of net sand. We further make the assumption that along the net sand con-
tour line (at least for any limited distance), the distribution of net sand and impermeable rock, as
seen in the well, remains constant within the unit being mapped. This assumption regarding the
rock distribution may not be true along the 50-ft contour over long distances, or on opposite
limbs of a fold, because of changes in depositional environments, variable structural growth his-
tories, and other factors. However, it is a reasonable assumption to make for a limited distance
from the well, parallel to the nearest net sand contour line.
Procedure for Walking a Well. Use three maps in walking a well through the wedge zone.
First, lay the structure map of the top of porosity over the net sand isochore map. Then overlay the
net gas base map on the two maps, as shown in Fig. 14-16a. We wish to walk Well No. 2, which is
located near the crest of the structure, through the wedge zone in order to estimate net pay in part
of the reservoir. We walk the well along the dashed line, parallel to the 50-ft net sand contour.
Notice that the inner limit of water (ILW) is shown on the structure map of the top of poros-
ity as a means to delimit the full thickness area on this map. From its position on the base-of-
porosity structure map, the ILW was simply traced on the top-of-porosity structure map. Now
we have the data necessary to make a net pay isochore map after walking the well, as described
in the following procedure.
1. Preparatory to walking a well, use a detailed 5-in. electric log to determine the net sand in
the well and the distribution of sand and impermeable rock. The 5-in detailed electric log
for Well No. 2 is shown in Fig. 14-l6b. The well contains 48 ft of net sand, and the log
shows increments of 10 ft of net sand per gross feet of interval, so chosen because we use a
10-ft contour interval in the net pay map. The productive unit is full of gas in Well No. 2, at
its position in the full thickness area.
2. Overlay the top-of-porosity structure map on the net sand map, then overlay the net gas
base map. Trace the gas/water contact as a zero net gas contour on the base map, and trace
the limit of the full thickness area as a dashed line.
3. On the net gas map, lightly mark a dashed line for the well path, parallel to the nearest net
sand contour. The net sand contour map may be removed for now; use it later in net pay
mapping of the full thickness area and for checking positions of net pay contours in the
wedge.
4. To begin walking Well No. 2, move the well, from its actual structural position, along the
well path and place it such that the top of porosity in the productive unit is at the gas/water
contact at –9298 ft, which corresponds to the zero contour line on the net gas isochore map
(Point A in Fig. 14-16a). A well drilled at this location would (1) encounter the top of
746 Chap. 14 / Isochore and Isopach Maps
Figure 14-15 Net sand contour map. Each net sand contour line is constructed with the assump-
tion that the amount of net sand is constant along the contour line.
porosity of the productive unit at the gas/water contact, (2) contain 48 ft of net sand, and (3)
have no pay.
5. Starting at the top of the sand on the electric log, determine the number of vertical gross
feet of section containing 10 ft of net sand. In Fig. 14-16b, a total of 16 ft of gross section
includes the 10 ft of net sand. From Point A, move the well up-structure 16 ft, along the
dashed line to locate Point B at –9282 ft. This Point B becomes a 10-ft net gas data point
for contouring the gas wedge, because a well drilled at Point B would encounter the top of
the sand 16 ft above the gas/water contact, and it would contain 10 ft of net gas sand.
6. To determine the location of the 20-ft net gas point, start at the base of the previous 10-ft
net sand section in the log and repeat the procedure. In this example (Fig. 14-16b), it re-
quires 21 vertical gross ft to obtain the next 10 ft of reservoir-quality sand. Move the well
up-structure 21 ft from Point B, along the dashed line, to locate Point C at –9261 ft. Point C
is the 20-ft net gas data point at 37 ft (16 ft + 21 ft) above the water level.
7. Continue the same procedure until the well is back at its original structural position. Erase
the dashed well path on the net gas map.
Using this method, the well may be walked completely across the wedge, resulting in a
more accurate contour spacing than using the method incorporating equal spaced and propor-
tionally spaced contours. We make no assumption about the vertical distribution of net sand
being uniform. We use the real distribution of net sand in an actual well to estimate net sand dis-
tribution in a part of the reservoir. Figure 14-16c shows the completed net gas isochore map con-
structed for this reservoir, incorporating the data obtained from walking Well No. 2.
We can walk as many wells as we deem necessary. However, we emphasize strongly that a
well must be walked parallel to the nearest contour line on the net sand isochore map. Signifi-
cant net pay contouring errors can occur if this procedure is not followed. Consider Well No. 5 in
the western portion of the reservoir (Fig. 14-17). If we wish to develop a more accurate contour
Methods of Contouring the Hydrocarbon Wedge 747
(a)
(b)
(c)
Figure 14-16 (a) Net gas base map overlain on top-of-porosity structure map and net sand iso-
chore map. Walk Well No. 2 through the wedge zone along the dashed line, parallel to the nearest
net sand contour line (50 ft). (b) Five-inch detailed log for the 9200-ft Reservoir, showing increments
of 10 ft of net sand per gross feet of interval. (c) Completed gas isochore map for the 9200-ft Reser-
voir. The contour spacing in the southeastern portion of the wedge zone was improved by walking
Well No. 2. (From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
748 Chap. 14 / Isochore and Isopach Maps
Figure 14-17 Structure map superimposed onto the net sand isochore map. Well No. 5 in the
western portion of the reservoir cannot be walked through the wedge zone to improve the spacing of
net pay contours. (Modified from Tearpock and Harris 1987. Published by permission of Tenneco Oil
Company.)
spacing in this area of the reservoir, can Well No. 5 be walked from the water level up-dip to the
inner limit of water, along the dashed line? The answer is no. Well No. 5 has 28 ft of net sand. If
the well is walked from the gas/water contact up-dip, it will be walked into an area of greater net
sand than is actually present in the well, as seen on the net sand map. Therefore, Well No. 5 can-
not be walked to improve the contour spacing in this area of the wedge. Caution must be taken
when walking wells to be sure that the assumptions made in choosing a well to walk are geologi-
cally reasonable and can be supported by a net sand isochore map or, if necessary, additional
study of sand distribution.
(a) (b)
Figure 14-18 (a) Five-inch detailed electric log for Well No. 2. The sand and impermeable rock are
not evenly distributed throughout this productive unit. (b) Top-of-porosity structure map superim-
posed on the net sand isochore map. The dashed lines indicate paths along which Wells No. 2 and 3
were walked through the wedge zone to improve the net gas contour spacing. (From Tearpock and
Harris 1987. Published by permission of Tenneco Oil Company.)
(a)
Figure 14-19 (a) Net gas isochore map based on equally and proportionally spaced contours. (b)
Net gas isochore map with the contour spacing based on walking Wells No. 2 and 3 through the
wedge. Compare this map to that shown in Fig. 14-19a. (From Tearpock and Harris 1987. Published
by permission of Tenneco Oil Company.)
750 Chap. 14 / Isochore and Isopach Maps
(b)
Figure 14-19 (continued)
No. 2 and 3 through the wedge. Observe the significant difference between the two net gas iso-
chore maps. If there were several leases involved in this gas reservoir, the net gas isochore map
prepared by walking the wells would provide a more accurate map for assigning equities to each
lease.
Several other methods exist for constructing accurate net pay isochore maps. For example,
there is a more accurate method of estimating pay when we walk wells. Also, using the method
that employs the construction of a net-to-gross ratio map for the entire productive unit provides
estimates of net pay at points within the wedge. In this section of the chapter, we review the
more detailed method of walking wells.
Using the same reservoir as in Fig. 14-18, we illustrate the more detailed technique of walk-
ing wells. Figure 14-20 shows a north-south diagrammatic cross section along the path used to
walk Well No. 2. On the right side of the figure, we position Well No. 2 so that the top of the
sand is at the gas/water contact (–9298 ft). On the left side of the figure, we position the well at
the inner limit of water (–9218 ft) where the base of porosity is at the gas/water contact. From
the right, Well No. 2 must be walked 16 ft up-structure from the gas/water contact to –9282 ft, to
obtain 10 ft of net pay sand. From this point up-dip to –9248 ft, the well gains no additional pay
because the section being raised above the water level contains nothing but shale. So at –9248 ft,
the net pay remains 10 ft. At this point, the gas/water contact intersects the top of the lower sand
member; therefore, moving up-structure from this position, the well gains additional pay. Con-
tinuing to walk the well up-dip, it must reach –9230 ft before an additional 10 ft of net sand, in
the lower member, lies above the water contact. So we estimate 20 ft of net pay at –9230 ft. At
the edge of the full thickness area, with the gas/water contact on the base of sand, all the net sand
(28 ft) lies above the water contact. From this point to the actual well location, shown on the far
left in the figure, the reservoir contains 28 ft of estimated pay.
On the cross section, two locations have 10 ft of net gas sand (–9282 ft and –9248 ft), and
the area between these points has a constant 10 ft of net gas sand. The accuracy of the net gas
isochore map can be improved by constructing a map honoring the two 10-ft net pay values.
Well No. 3 was also walked through the wedge zone, as indicated in Fig. 14-18b, to aid in the
construction of the net pay isochore map.
Methods of Contouring the Hydrocarbon Wedge 751
Figure 14-20 Diagrammatic cross section along the path taken to walk Well No. 2.
Figure 14-21 shows the resulting net gas isochore map. At first glance, it may appear as if
an important contouring rule was broken in the construction of the isochore map: contours can-
not merge or split. However, no rules are broken. The two 10-ft contour lines that appear to
merge represent the limits of a very wide 10-ft contour line. Everywhere within the area of the
wide contour, the net gas has a constant value of 10 ft.
One may ask, since the sands are so far apart and separated by such a thick shale break, why
we would not map each sand separately and construct two isochore maps. In the western part of
the reservoir, some of the sands within the two sand members shale-out, and sands develop
within the thick shale interval. The result is a reservoir consisting of interfingering shales and
sands in which there is productive communication throughout the reservoir, with the sands act-
ing as a single reservoir. Therefore, the thick shale wedge is localized in the eastern section of
the reservoir. The rapid decrease in width of the 10-ft contour line reflects the fact that the shale
interval decreases to the west. If such a shale interval were known to be continuous over the en-
tire reservoir, it would be necessary to prepare a structure map for each sand member and con-
struct a separate net pay isochore map for each sand member.
The procedure outlined in this section is more involved than the two previous methods
shown, but the technique provides further accuracy in the construction of a net hydrocarbon iso-
chore map. The method chosen to prepare a net hydrocarbon isochore map depends upon a num-
ber of factors, including the available time, detail, and accuracy required.
752 Chap. 14 / Isochore and Isopach Maps
Figure 14-21 Net gas isochore map using a more accurate method of contouring the wedge
based on the results of walking Wells No. 2 and 3.
Figure 14-22 Cross-sectional area of two reservoirs of equal volume and a constant TST of 100 ft.
One reservoir is horizontal, the other is dipping at 45 deg.
order to maintain the same cross-sectional area or volume of the reservoir, the shortened length
must be multiplied by the TVT.
For directionally drilled wells, the log thickness of a given stratigraphic interval can either
be thicker, equal to, or thinner than that seen in a vertical well drilled through the same strati-
graphic section. A correction factor must be applied to the MLT in most deviated wells to con-
vert the borehole thickness to TVT. The correction factor consists of two parts: (1) the correction
for wellbore deviation angle within the interval of interest, and (2) the correction for bed dip.
Any one of equations (4-3), (4-4), (4-5), or (4-6) in Chapter 4 may be used to calculate this cor-
rection factor. In Chapter 4, we used the equations to estimate the TVT of missing or repeated
section in a well as the result of a fault. Remember, the vertical separation of a fault at a well
equals the TVT of the stratigraphic section missing or repeated in the wellbore. In this chapter,
we look at the same correction factor equations in order to convert deviated wellbore thickness
to TVT, for use in net sand and net pay isochore mapping.
For convenience, we repeat the correction factor equation (4-6). Equation (4-6), which is a
3D equation, is the preferred equation to use for correction factors because this one equation can
be used to calculate the thickness correction factor regardless of the direction of wellbore devia-
tion, and the true dip of the beds is used instead of the apparent dip required in the two-dimen-
sional equations. Recall from Chapter 4 that we refer to this equation as Setchell’s equation.
where
If the beds are horizontal, then Setchell’s equation reduces to the simple correction factor
equation
Figure 14-23 illustrates the measurement of azimuth and ∆ azimuth for use in Eq. (4-6). For
convenience, the ∆ azimuth used in the equation is typically the acute angle between the azimuth
of the wellbore and the azimuth of the true bed dip.
In order to more closely examine the two directionally drilled wells shown in Fig. 14-24,
look first at the well drilled to the east in a down-dip direction (Fig. 14-24a). Consider the inter-
val to be a reservoir filled with gas or oil. The reservoir has a MLT of 476 ft. We apply the cor-
rection factor for wellbore deviation only, using Eq. (4-3). The MLT reduces to 357 ft, shown in
the figure as the TVD thickness, or the true vertical depth thickness (TVDT). This thickness
also exceeds the TVT of the interval, because the correction for only wellbore deviation does not
take into account the dip of the beds. The TVDT is that thickness of an interval obtained from a
true vertical depth (TVD) log and, for dipping beds, TVDT does not equal TVT (see Chapter 4).
With the final correction for bed dip, the MLT converts to the TVT of 150 ft, shown in Fig. 14-24a
at the penetration point of the wellbore in the top of the reservoir. Note that the TST is 123 ft.
(a) (b)
Figure 14-24 (a) TVT correction for a well drilled in a down-dip direction. (b) TVT correction for a
well drilled in an up-dip direction.
The TST is calculated by multiplying the TVT by the cosine of the angle of bed dip (35 deg in
this example).
The well in Fig. 14-24b deviates up-dip, to the west. The MLT of 127 ft measures less than
the TVT. Applying a correction factor for the well deviation angle alone, which is equivalent to
the correction to TVDT, provides an even smaller thickness of 82 ft. When Eq. (4-6), the correc-
tion factor equation for both bed dip and wellbore deviation, is applied, the MLT converts to the
TVT of 150 ft, the value needed for net sand and net pay mapping. As an exercise, use Eq. (4-6)
to calculate the TVT for the two wells in Fig. 14-24 and to confirm the results shown.
Various computer programs can be used to create TVD, TVT, and TST logs from measured
depth (MD) logs for use in mapping. The deviated well log data, the directional survey for the
well, and bed dip information are necessary as input data. The log data are obtained from the
logging company tapes or digitized from the actual log. The directional survey data are furnished
by the directional company that worked the well. The bed dip information can be obtained either
from completed structure maps or from a dipmeter. The output logs can be in standard presenta-
tion or at any scale desired. The logs in Fig. 14-25 were created by using IEPS (Integrated Ex-
ploration and Production System). The log sections for Well No. MP-D5, shown from left to
right in the figure, represent the (1) measured depth log, (2) true vertical depth log, (3) true verti-
cal thickness log, and (4) true stratigraphic thickness log. This well is in an area of significant
bed dip. Notice the similarity of the MD log and the TVD log. This is so because the TVD log
thickness is equivalent to a correction for wellbore deviation only, and not bed dip. However, the
TVT log shows a considerable reduction in thickness from the MD log because the TVT log re-
flects corrections for both wellbore deviation and bed dip.
We caution here that TVD logs, which are a standard part of the log suite for a deviated
well, are used too often for purposes that are not applicable with this log. A widespread misun-
derstanding exists that a TVD log prepared from a MD log can be used to (1) correlate with
other well logs, (2) determine the vertical separation for a fault, and (3) count net sand and net
pay for isochore mapping. Remember, a TVD log is corrected only for wellbore deviation, and
not bed dip. In areas of flat-lying beds, a TVD log is equivalent to a TVT log because the only
756
Chap. 14 / Isochore and Isopach Maps
Figure 14-25 Computer-generated electric logs illustrating the difference in thickness between measured depth, true vertical depth,
true vertical thickness, and true stratigraphic thickness logs, generated for the same well. (From Tearpock and Harris 1987. Published by
permission of Tenneco Oil Company.)
Vertical Thickness Determinations 757
correction factor required is for wellbore deviation (Fig. 14-26). However, if the beds are dipping
(particularly over 10 deg), a TVD log typically does not represent the log thickness required to
aid in correlation work, to determine the vertical separation for a fault, or for use in net sand and
net pay counting. For these purposes, we must correct a deviated well log so that the log thick-
ness represents the TVT. Look again at Fig. 14-25 and observe the significant difference in
thickness between the TVD and the TVT logs. To determine the vertical separation of a fault by
correlation of a faulted well with a deviated well, and to count all net sand and net pay from a
deviated well log, we must use a TVT log or its equivalent. By the equivalent of a TVT log, we
mean calculating and using correction factors if a TVT log is unavailable, which is commonly
the case. Therefore, for each interval on the deviated well log requiring the conversion of MLT to
TVT, determine correction factors and apply them to the MLTs of the intervals of interest.
Figure 14-27 (a) Net pay isochore map for the T-1 Sand, Reservoir A. The net pay values for the
deviated wells from Platforms A and D were corrected only for wellbore deviation. (b) Net pay iso-
chore map for the T-1 Sand, Reservoir A. The net pay values for the deviated wells from Platform A
and D were corrected for wellbore deviation and bed dip. Compare the net pay value for each well
with those shown in Fig. 14-27a. (From Tearpock and Harris 1987. Published by permission of Ten-
neco Oil Company.)
taking into account the correction factor for bed dip. So, based on the incorrect map, the volu-
metric calculation overstates the reserves by 18 percent.
In a situation like this, we would expect the error factor to be larger than 18 percent, and it
would be in most cases. However, look at Wells No. A-2 and D-5. Well No. A-2 was corrected
upward from 17 ft net pay to 24 ft net pay, whereas Well No. D-5 was corrected downward from
44 ft net pay to 30 ft net pay. It so happens that the D Platform wells, drilled down-dip, result in
a reduction in net pay values when the correction factor for bed dip is considered, whereas the
A Platform wells, drilled up-dip, result in an increase in net pay. Therefore, a significant part of
the error is negated because of the directions in which the wells were drilled.
This reservoir is only one of a number of oil and gas reservoirs that are productive within
this field. A complete remapping of the field was undertaken when several major mapping er-
rors, such as the one shown here, were identified. The remapping of the field resulted in a signif-
icant write-down of hydrocarbon reserves that were overestimated because of mapping errors,
such as the failure to incorporate the proper correction factors in determining net pay values.
Several development wells planned for the field based on the overestimated reserves were not re-
quired, saving the company millions of dollars in unnecessary wells. This example illustrates the
impact that correction factors can have on estimated hydrocarbon volumes determined from vol-
umetric calculations using net pay isochore maps.
We have seen error factors of several hundred percent. In one example, the net pay used for
the reserve estimate from a deviated well was 750 ft, when in fact the actual TVT value was
about 150 ft. This is a major error, resulting in a significant overestimation of reserves.
Vertical Thickness and Fluid Contacts in Deviated Wells 759
Figure 14-28 A deviated well penetrates the top and base of a stratigraphic unit at different posi-
tions relative to the surface location of the well.
760 Chap. 14 / Isochore and Isopach Maps
150 ft TVT. Observe an oil/water contact just down-dip from the well’s penetration of the base
of the reservoir. The entire wellbore penetration is therefore confined to the full thickness area of
the reservoir, with no fluid contacts present in the well.
The well on the left in Fig. 14-29 penetrates an oil/water contact. The MLT of the entire
stratigraphic interval is 282 ft. What correction factor equation would you use for this MLT to
calculate TVT of the oil column, at the penetration point of the top of the reservoir? In this par-
ticular situation where the penetration point is within the full thickness area, Setchell’s equation,
applied to the MLT of 282 ft, will provide the TVT of the oil at the penetration point in the top of
the reservoir.
Application of Setchell’s equation to the MLT of the net oil (less than 282 ft) would result in
an incorrect TVT at the penetration point in the top of the reservoir. However, if you wish to ob-
tain the value of the vertical feet of oil directly above the penetration point of the wellbore at the
oil/water contact, the use of the Setchell equation provides this value of net pay in the wedge.
We will discuss this situation next.
Now consider the following situation. A well penetrates the top of an oil reservoir within the oil
wedge; i.e., directly above the oil/water contact. The following data apply to the well and well log.
What is the vertical thickness of the oil column under the penetration point of the well at the
top of the reservoir? Draw a cross section showing the relationship of the well to the productive
unit and, using Eq. (4-6), calculate the TVT of the oil under the penetration point at the top of the
reservoir. Is the correct answer 150 ft, 100 ft, or 79 ft? If you calculated 150 ft, this thickness is
equal to the total true vertical thickness of net sand directly under the penetration point of the well
Vertical Thickness and Fluid Contacts in Deviated Wells 761
at the top of the sand (Fig. 14-30). If you calculated 79 ft, this is the vertical thickness of net oil pay
directly above the point where the well penetrates the oil/water contact. A review of Fig. 14-30
shows that the oil/water contact is a horizontal surface. Therefore, we do not have to consider any
bed dip correction factor to calculate the net pay directly under the penetration point. No bed dip
affects the TVT of the net oil sand, since the oil/water contact is a horizontal surface, and the pene-
tration point at the top of the sand is a point directly over the oil/water contact. Therefore, the Eq.
(4-6), as applied to calculation of net oil sand directly beneath the penetration point of the well at
the top of the sand, reduces to the correction factor for wellbore deviation multiplied by the MLT
from the top of the sand to the oil/water contact, which is 115 ft. The net oil pay calculates as
Therefore
CF = [cos 30° – (sin 30° cos 0° tan 20°)]
CF = 0.866 – (0.5)(1)(0.364)
CF = 0.684
1. The true vertical thickness of the net sand at the penetration point of the well at the top of
the sand is
TVT = MLT × CF
TVT = (219 ft)(0.684)
TVT = 150 ft
2. The true vertical thickness of the net oil pay sand above the penetration point of the well at
the oil/water contact is
by a deviated well. In this case, the vertical thicknesses of the oil and gas columns must be deter-
mined for isochore mapping. Using the data provided in the figure, verify the TVT for the oil and
for the gas at two separate locations: (1) directly beneath the penetration point of the well at the
top of the reservoir, and (2) the penetration point of the gas/oil contact. Finally, calculate the TVT
of water, oil, and gas at the point where the well penetrates the oil/water contact (dashed line).
Figure 14-32 illustrates a well deviated in an up-dip direction, penetrating an oil reservoir in
the oil wedge. The same procedures as previously discussed are used to calculate the TVT of oil
at the locations where the well penetrates the top of the reservoir and the oil/water contact. If de-
sired, the TVT of oil can also be calculated at the position where the well penetrates the base of
the productive unit. First, calculate the TVT of the water column at this location, and then sub-
tract this value from the TVT of the total net sand to arrive at the vertical thickness for oil.
By having a good understanding of the geometric relationship of the productive unit, well-
bore, and fluid contacts, we can use this knowledge to our advantage. The calculation of the net
gas or net oil at various points in a well, such as the well’s penetration point at the top of the pro-
ductive unit, base of the unit, or at fluid contacts, can provide additional net pay values. These
values can be used in the preparation of the net gas or net oil isochore maps, providing additional
control in the wedge zones.
The detailed calculations shown in this section are not always required or justified. How-
ever, the use of these techniques may prove to be very important where detailed, accurate map-
ping is needed for some specific reserve estimate, development plan, enhanced recovery
program, unitization, equity determination, or litigation.
Once the structure mapping is complete, the question arises as to whether a separate map on
the top of porosity is required for accurate delineation of the reservoir and for use in the con-
struction of net hydrocarbon isochore maps, from which estimates of reserves are made. Two pa-
rameters are considered in evaluating the importance of the nonreservoir-quality rock: (1) the
thickness of the tight zone, and (2) the relief of the structure. A thick tight zone has a greater ef-
fect on ultimate reserve estimates than one that is thin. Low-relief structures introduce greater
error in delineating the limits of a reservoir than steeply dipping structures, particularly if the
low-relief structure contains a reservoir with bottom water. This is true because a steeply dipping
reservoir is associated with a relatively small wedge zone when compared to the total area of the
reservoir.
On a low-relief structure, the wedge zone of a reservoir can represent a significant portion
of the total reservoir area (Fig. 14-33). Figure 14-33a shows in map and cross-sectional views a
low-relief bottom water reservoir mapped on the top of a productive unit, which consists of non-
reservoir-quality rock in the upper 75 ft. The same reservoir is mapped on the top of porosity in
Fig. 14-33b. The net oil isochore map prepared from each structure map is shown in Fig. 14-33c.
The same net pay values are assigned to each well in both isochore maps. In this case, because
the reservoir is on a low-relief structure, the difference in reservoir volume between the isochore
map based on the top of structure and the isochore map based on top of porosity is 637 acre-feet,
(a)
Figure 14-33 (a) The structure map on the top of the stratigraphic unit containing the 6000-ft Reser-
voir, and the cross section A-A′. Upper 75 ft of the unit contains nonreservoir-quality rock. (b) The
structure map on top of porosity (6000-ft Reservoir) and the cross section A-A′. (c) Two separate net
pay isochore maps: (1) the upper isochore map is based on the structure map on the top of the strati-
graphic unit, and (2) the lower isochore map is based on the top-of-porosity structure map. Net pay val-
ues for all the wells are the same for each map. (d) There is a 32 percent reduction in reservoir volume
for the net pay isochore map constructed from the top of porosity map versus the net pay isochore
map constructed from the top of the stratigraphic unit. This is a significant reduction in volume.
Mapping the Top of Structure Versus the Top of Porosity 765
(b)
(c)
Figure 14-33 (continued)
766 Chap. 14 / Isochore and Isopach Maps
(d)
Figure 14-33 (continued)
FAULT WEDGES
A fault wedge within a reservoir is defined as a wedge-shaped section of strata bounded by a
fault. It is commonly just as important to map net pay within the fault wedge of a productive
reservoir as it is to map net pay within the hydrocarbon wedge above water. If a productive unit
is thin or the dip of the bounding fault is a high angle, the reservoir volume affected by the fault
wedge may be insignificant and can be ignored for all practical purposes. In cases where the
reservoir is thick or the fault dips at a low angle, the reservoir volume within the fault wedge
may be significant and must therefore be included when constructing the net pay isochore map.
There are several ways to handle the mapping of a fault wedge. It may be contoured in the con-
ventional way, or a mid-trace (mid-point) method can be used.
Figure 14-34 Structure map on top of a reservoir shows the intersections of a fault with the top
and base of the unit. These two intersections are required in mapping the fault wedge.
bounded by an oil/water contact at –8000 ft. The structure map shows the intersection of the top
of the productive unit with the fault, as well as the intersection of the fault with the base of the
unit (shown as a dashed line). The area between these two intersections is the area of fault
wedge. In this case, the fault dips to the west at 45 deg and the productive unit dips at 30 deg. It
is readily apparent from the fault intersections with the top and base of the unit that the fault
wedge comprises a large portion of this reservoir. For simplicity, we assume the reservoir to have
50 percent net sand and 50 percent shale, evenly distributed throughout the gross interval. Be-
cause of this even shale distribution, we can evenly space the contours in both the hydrocarbon
and fault wedges. The reservoir is in the downthrown fault block, so the control required to map
the fault wedge in this example includes the intersections of the top and base of the productive
unit with the fault. The upthrown trace of the fault, and therefore the fault gap as seen on a struc-
ture map, plays no part in mapping the fault wedge.
Figure 14-35a illustrates the net oil isochore map for this reservoir with both the water and
fault wedges conventionally contoured. The cross section A-A′, drawn below the net oil isochore
map, depicts certain key control points in the reservoir, including the oil/water contact on the top
of the productive unit, the inner limit of water (the oil/water contact on the base of the unit), the
intersection of the base of the reservoir with the fault, and the intersection of the top of the reser-
voir with the fault. These key control points play an important part in the construction of the
wedges for this net oil isochore map.
Mid-Trace Method
The use of the conventional method of contouring the fault wedge can be tedious and at times
unjustified. In such cases, we can use a shortcut method for contouring the fault wedge, referred
to as the Mid-Trace Method, or Mid-Point Method (Fig. 14-35b). To construct an isochore
768 Chap. 14 / Isochore and Isopach Maps
(a)
Figure 14-35 (a) The fault wedge is mapped using the conventional method. See cross section
below map. (b) Mid-trace method for mapping the fault wedge. See cross section.
map using this method, first draw a line that is one-half way between the intersections of the top
and base of the reservoir with the fault. This line is the mid-trace. Draw the line to stop at the
inner limit of water, shown as points a and a′ in Fig. 14-35b. Next, extend this line from the inner
limit of water straight to the intersection of the fault with the oil/water contact on the top of the
reservoir, as indicated by a-b and a′-b′ on the figure. This line b-a-a′-b′ becomes a boundary of
the reservoir. Any part of the reservoir in the fault wedge, but outside this boundary line, is con-
sidered as being folded under to convert the wedge zone inside this line to an area of full thick-
ness. This is illustrated in the cross section below the net oil isochore map in Fig. 14-35b. Notice
that this procedure extends the full thickness area to the mid-trace. Next, extend all contours in
the hydrocarbon wedge to intersect with line segments a-b and a′-b′, as illustrated in the figure.
The finished map includes an enlarged full thickness area (125 ft of net pay) plus a wedge.
Figure 14-36 illustrates the use of the mid-trace method of contouring a fault wedge with a
reservoir bounded by two intersecting faults. Note that the mid-trace extension, from the inner
limit of water to the intersection of the fault with the oil/water contact on the top of the unit, is in
opposite directions on the individual fault wedges. Remember, always draw a line from the mid-
trace to the intersection of the zero net pay contour with the fault.
Nonsealing Faults 769
(b)
Figure 14-35 (continued)
NONSEALING FAULTS
Although faults play a very important role in the trapping of hydrocarbons, studies have shown
that some faults are nonsealing, thereby permitting the migration of hydrocarbons from one fault
block to the next. One of the most common situations resulting in a nonsealing fault is the juxta-
position of permeable stratigraphic units across a fault, with capillary properties of the fault zone
being insufficient to create a seal. This can occur where a fault does not have sufficient displace-
ment to separate an entire permeable unit from one fault block to the next, or where juxtaposi-
tion of different permeable units exists.
Hydrocarbon accumulations can occur on both sides of a nonsealing fault if a trap exists in
the up-structure fault block. The accumulations would have a common water level. With non-
sealing faults, it is important to map fault wedges because they can contain significant amounts
of hydrocarbons. We contour the fault wedge basically the same as presented in the previous sec-
tion, with one exception. Since there is hydrocarbon pay upthrown and downthrown to the fault,
two reservoirs and two fault wedges must be mapped to account for all the hydrocarbon volume.
The isochore maps for the two reservoirs can be constructed individually or contoured as one
map.
770 Chap. 14 / Isochore and Isopach Maps
Figures 14-37 and 14-38 illustrate an example of reservoirs juxtaposed across a nonsealing
fault. A productive sandstone unit contains two fault-separated reservoirs on the southern flank
of a salt structure. Figure 14-37a shows the structure map on the top of porosity in Reservoirs A
and B. The reservoirs are limited to the north by a vertical salt face, to the east and west by
faults, and in the down-dip direction by a common gas/water contact at –10,550 ft. The interval
from top to base is in excess of 200 ft thick and it is all reservoir-quality sand. Fault B, with a
vertical separation of 200 ft, has insufficient displacement to completely separate the sand, so
part of the sand is juxtaposed across the fault. The same gas/water contact in both Reservoirs A
and B indicates that the fault is nonsealing and that the two reservoirs are in communication.
Figure 14-37b through d are the structure map on the base of the 9500-ft Sand, the fault surface
map, and the net sand map respectively. These maps are required for isochore construction.
The net gas isochore maps for the Reservoirs A and B can be constructed separately or as
one single net gas isochore map. Figure 14-38a and b show the individually constructed net gas
Nonsealing Faults 771
(a)
(b)
Figure 14-37 (a) Structure map on top of Reservoirs A and B. Salt face is vertical at this depth. (b)
Structure map on base of Reservoirs A and B. (c) Fault surface map of Faults A and B. (d) Net sand
isochore map for 9500-ft Sand.
772 Chap. 14 / Isochore and Isopach Maps
(c)
(d)
Figure 14-37 (continued)
Nonsealing Faults 773
(a)
(b)
Figure 14-38 (a) Net gas isochore map for Reservoir A. The map includes two fault wedges and a
water wedge. (b) Net gas isochore map for Reservoir B. Fault wedge and water wedge are signifi-
cant portions of the reservoir volume. (c) Composite net gas isochore map for Reservoirs A and B.
Notice the complexity in the mapping of the two fault wedges created by Fault B when both reser-
voirs are mapped together.
774 Chap. 14 / Isochore and Isopach Maps
(c)
isochore maps for Reservoirs A and B respectively. Notice that the fault wedges for Faults A and
B comprise a significant portion of each of the reservoirs. Figure 14-38c illustrates how the two
reservoirs can be mapped together by combining the wedges for Fault B. Cross section A-A′
shows about 75 ft of sand juxtaposed across the fault at this position in the reservoir.
Isochore maps in which the fault wedges are combined, such as the one shown in Fig.
14-38c, are difficult to construct and can easily result in errors. We recommend for simplicity of
construction and planimetering that, even with nonsealing faults, each reservoir and fault wedge
be mapped separately, as shown in Fig. 14-38a and b.
Figure 14-39 In the preparation of a net hydrocarbon isochore map, the configuration of the reser-
voir is completely rearranged and the base is artificially flattened. Compare the configuration of the
net gas and net oil cross sections in the lower portion of the figure to that of the structure cross sec-
tion in the center.
This equation is used to determine the volume of a layer between successive slices, which
are based on vertical thickness and represented on the map by net pay contour lines (Fig. 14-40).
The total volume of the reservoir is the sum of these separate volumes.
The second equation used in the horizontal slice method determines the volume of a trapezoid.
1
Volume = h( An + An + 1 )
2
where
The pyramidal equation usually provides the most accurate results; however, because of its
simplicity, the trapezoidal equation is commonly used. Since the trapezoidal equation introduces
an error of about 2 percent where the ratio of successive areas is 0.5, a common convention is
used to employ both equations. Wherever the ratio of the areas within any two successive isochore
lines is smaller than 0.5, the pyramidal equation is applied. Wherever the ratio of the areas within
any two successive isochore lines is larger than 0.5, the trapezoidal equation is used. Computer
programs, for calculating reservoir volumes from net pay maps, are capable of combining the
pyramidal and trapezoidal equations in the manner described. However, the programs may vary in
the cutoff ratio that is used, so that ratio for a given program should be determined by the user.
Figure 14-40 and Table 14-1 outline the volume determination using the horizontal slice
method. Take a few minutes and review this example to obtain a good understanding of the
procedure.
Reservoir Volume Determinations from Isochore Maps 777
Figure 14-40 Cross section and net pay isochore map of an idealized reservoir. The cross section
shows the depicted solid divided into 5-ft-thick horizontal slices. (Craft and Hawkins 1959. Reprinted
by permission of Prentice-Hall, Inc.)
Table 14-1
Planimeter Area Ratio Interval ∆V
Area Sq. In. Area Acres of Areas H Feet Equation Ac. Ft.
A0 19.64 450 – – – –
A1 16.34 375 0.83 5 Trapezoid 2063
A2 13.19 303 0.80 5 Trapezoid 1695
A3 10.05 231 0.76 5 Trapezoid 1335
A4 6.69 154 0.67 5 Trapezoid 963
A5 3.22 74 0.48 5 Pyramid 558
A6 0.00 0 0.00 4 Pyramid 99
Total ac. ft. 6713*
*The percentage difference in acre-feet between the horizontal and vertical slice methods is less than 1%.
where
h = Average thickness between successive contour lines
A0 = Zero contour line
A1 = Next higher value, or next successive, contour line
An = Highest value contour line
havg = Average thickness within An
Figure 14-41 and Table 14-2 illustrate the procedure for volume determinations using the
vertical slice method. The reservoir used for this example is the same one used for the horizontal
slice method (Fig. 14-40), so the results can be compared. The difference in calculated volume
between the horizontal and vertical slice methods, for the example in Figs. 14-40 and 14-41, is
less than 1 percent.
Figure 14-41 Cross section and net pay isochore of an idealized reservoir. The cross section
shows the reservoir divided into vertical slices. (Craft and Hawkins 1959. Reprinted by permission of
Prentice-Hall, Inc.)
Table 14-2
Planimeter Area Difference in Average V
Area Sq. In. Area Acres Areas An-1 – An Thickness Feet Ac. Ft.
A0 19.64 450 – – –
A1 16.34 375 75 2.5 187
A2 13.19 303 72 7.5 540
A3 10.05 231 72 12.5 900
A4 6.69 154 77 17.5 1347
A5 3.22 74 80 22.5 1800
A6 0.00 0 74 27.0 1998
Total ac. ft. 6772*
*The percentage difference in acre-feet between the horizontal and vertical slice methods is less than 1%.
Introductory Reservoir Engineering 779
Choice of Method
The choice of using the horizontal or vertical slice methods is usually based on individual prefer-
ence, since both methods are reasonably accurate. The choice is less important than the assurance
that the method is used correctly, avoiding planimetering pitfalls. It is therefore of utmost impor-
tance that anyone doing actual planimetering be thoroughly familiar with the pitfalls that can be
encountered when planimetering and with the mathematical programs for volume calculations.
A geoscientist may spend months working on a prospect with the end result being a net pay
isochore map prepared to estimate the hydrocarbon volume for the prospect. If the isochore map
is planimetered incorrectly, as a result of carelessness or a lack of understanding of the planime-
tering procedures, a viable project can be mistakenly rejected. We highly recommend that all
planimetered work be spot-checked by the geologist or reservoir engineer who prepared the net
pay isochore map.
Reservoir Characterization
After the identification of hydrocarbons in a permeable stratigraphic unit, a plan must be estab-
lished to estimate the in-place volumes of these hydrocarbons and what might be expected to be
recovered. Recoverable hydrocarbons are referred to as hydrocarbon reserves. Because the pro-
ductive characteristics of a reservoir or field may not be fully known until it is maturely devel-
oped, offset analogous reservoirs need to be considered as models to the development of any
new discovery. With a model to follow, various reservoir-engineering parameters such as reser-
voir porosity, permeability, and water saturation can be estimated to allow volumetric analyses
of the reservoir or field to estimate potentially recoverable reserves.
Estimation of Reserves
Reservoir bulk volume is calculated from net hydrocarbon isochore maps, as discussed in Chap-
ter 10. From detailed log analysis, rock properties can be identified. As an example, interstitial
water saturation must be estimated for reserve determinations. Detailed reservoir petrophysics
780 Chap. 14 / Isochore and Isopach Maps
for use in calculating reserves is beyond the scope of this book; however, a presentation of gen-
eral reservoir engineering may be helpful, and it is outlined here.
Using the letter symbols G and N, initial in-place volumes of oil and gas can be determined
by the following equations.
and
where
G = original gas-in-place, in cu ft
N = original oil-in-place, in barrels
φ = effective porosity, fraction
Swi = interstitial water saturation, fraction
B = formation volume factor, dimensionless
43,560 = cu ft per acre-foot
7758 = barrels per acre-foot
subscript o = oil-bearing zone
subscript g = gas-bearing zone
Bgi = standard cu ft/reservoir cu ft
Boi = reservoir barrels/stock tank barrels
Gas Reservoirs. Equation (14-4) is used to estimate original gas-in-place. There are several un-
known factors in the equation that must be determined. The formation volume factor (FVF) is de-
fined as the relationship of gas volumes from surface conditions to reservoir conditions. For gas
reservoirs, Bgi is expressed in standard cubic feet per cubic foot, SCF/cu ft. The porosity is expressed
as a fraction of the bulk volume, and the interstitial water (Sw) is a fraction of the pore volume.
To determine the unit recovery for a gas reservoir, the final reserve volume per acre-foot is
determined based on the reservoir drive mechanism. This fractional recovery, or recovery factor,
represents the difference between the initial unit-in-place gas and the final, or abandonment,
unit-in-place gas.
100(G − Ga)
Recovery Factor ( RF ) =
G
or
where
This recovery factor is indicative of depletion drive reservoirs, where interstitial water satu-
ration remains unchanged and, conversely, gas saturation remains constant. The other end of the
spectrum with regard to drive mechanisms is a strong water drive, where produced gas is being
replaced by encroaching water (there is no appreciable pressure loss and Bgi = Bga ). The recov-
ery factor for a water-drive gas reservoir, which is representative of the change in gas and water
saturations in the reservoir due to production, is shown in the following equation.
Oil Reservoirs. Oil reservoirs are often more difficult to analyze for a number of reasons, in-
cluding the presence in some reservoirs of both oil and gas. If an oil reservoir is found without
free gas, the oil is said to be undersaturated. An oil reservoir with a free gas cap is indicative of a
saturated oil reservoir.
Equation (14-5) is used to volumetrically determine original oil-in-place. The variables in
Eq. (14-5) are very similar to those discussed for Eq. (14-4). If we consider an undersaturated oil
reservoir under a strong water drive, the recovery factor is based on the following equation.
Where there is an initial gas cap, the oil is saturated. In such cases, the reservoir can be pro-
duced under drive mechanisms other than water drive. These include dissolved gas, gas cap, or a
combination drive. The opposite end of the recovery factor spectrum from a water drive reservoir
is that of a dissolved gas drive reservoir.
cast future performance. It is good practice to monitor both production and reservoir pressure
data. These points of reference are invaluable in evaluating reservoir heterogeneities and are ap-
plicable to material balance equations.
Various performance curves can be used to evaluate reserves and forecast future production
trends. The most common performance curves are those which plot the production of oil, gas,
and water versus time. In many instances, performance evaluation is by far the most accurate
method for estimating the original in-place volume of hydrocarbons, estimating recoverable re-
serves, and forecasting future performance. The following is a partial list of the types of perfor-
mance curves that can be plotted.
A detailed discussion of these performance curves is beyond the scope of this book. We
again refer you to Craft and Hawkins (1959).
where
For example, with a 10-deg bed dip, an interval with a TST of 100 ft has a vertical thickness
of 101.5 ft; at a 20-deg dip, the vertical thickness is 106.5 ft. At a 45-deg dip, however, the verti-
cal thickness becomes 141 ft, which is significantly different than the stratigraphic thickness. If
the true vertical thickness is the value contoured, the map is more correctly referred to as an iso-
chore map. Therefore, net sand and net pay maps are isochore maps.
Interval Isopach Maps 783
TVT
TST =
(sin α tan φ) + cos α (14-11)
where
Figure 14-42 Effect of changing bed dip on TVT of a stratigraphic unit, where TST is constant.
784 Chap. 14 / Isochore and Isopach Maps
Figure 14-43 shows the equation and an example problem. This equation takes into account
the dip of the upper and lower surfaces and the vertical thickness of the interval.
Finally, in the case of deviated wells, the measured log thickness must be first corrected to
TVT, and then corrected to TST. The procedure for vertical thickness conversions is discussed in
this chapter and in Chapter 4.
To avoid making laborious stratigraphic thickness calculations, the graph in Fig. 14-44 can
be used to calculate TST if the vertical thickness and the dips of the upper and lower surfaces are
known. The horizontal axis represents the dip of the upper surface, and the vertical axis repre-
sents the correction factor. The curves within the graph represent the values obtained by sub-
tracting the dip of the upper surface from the dip of the lower surface (note positive and negative
values). Consider the following example.
Data:
Figure 14-43 Cross section showing the geometric relationship between two horizons that have
different angles of dip. Equation (14-11) is used in this type of situation to convert TVT to TST. (Pre-
pared by C. Harmon. Modified from Tearpock and Harris 1987. Published by permission of Tenneco
Oil Company.)
Interval Isopach Maps 785
2. Enter the graph on the horizontal axis at 20 deg and project vertically until the line inter-
sects the curve with value equal to the dip of the lower surface minus the dip of the upper
surface. In this case, it is the +10-deg curve.
3. From the intersection with the curve, project horizontally to the left to intersect the vertical
axis and determine the correction factor. In this case, it is 0.88.
Therefore,
Figure 14-44 Graph derived from Eq. (14-11) and used to determine the correction factor for con-
verting TVT of a stratigraphic interval to TST, where the upper and lower surfaces dip at different an-
gles. (Prepared by C. Harmon. From Tearpock and Harris 1987. Published by permission of Tenneco
Oil Company.)
786 Chap. 14 / Isochore and Isopach Maps
(a)
Figure 14-45 (a) Which line, A or B, represents the true stratigraphic thickness of the designated
interval? (b) True (1:1 scale) cross section of the seismic interval shown in (a). (Prepared by C. Har-
mon. From Tearpock and Harris 1987. Published by permission of Tenneco Oil Company.)
Interval Isopach Maps 787
(b)
Figure 14-45 (continued)
line B on the seismic line and cross section. The reason for this pitfall is that the seismic line, at
this depth, has about a 2:1 vertical exaggeration. In this case, you would post a larger thickness
for the interval than is actually present. To obtain corrected data points, you need to apply Eq.
(14-11), which uses the dip of the top and bottom horizons, and the vertical thickness of the in-
terval. The graph in Fig. 14-44 can also be used to determine the stratigraphic thickness.
In summary, seismic information can be a valuable source of interval thickness data, as long
as you are aware of the visual distortion inherent in seismic data and properly account for it in
the calculation of stratigraphic thicknesses.
APPENDIX
788
Appendix 789
Figure A-2 Map constructed using the interpretive contouring technique. (Reproduced from
Analysis of Geologic Structures by John M. Dennison, by permission of W. W. Norton & Company,
Inc. Copyright©1968 by W. W. Norton & Company, Inc.)
Figure A-3 Map constructed using the mechanical contouring technique. (Reproduced from
Analysis of Geologic Structures by John M. Dennison, by permission of W. W. Norton & Company,
Inc. Copyright©1968 by W. W. Norton & Company, Inc.)
Appendix 791
Figure A-4 Map constructed using the equal-spaced contouring technique. (Reproduced from
Analysis of Geologic Structures by John M. Dennison, by permission of W. W. Norton & Company,
Inc. Copyright©1968 by W. W. Norton & Company, Inc.)
Figure A-5 Map constructed using the parallel contouring technique. (Reproduced from Analysis
of Geologic Structures by John M. Dennison, by permission of W. W. Norton & Company, Inc. Copy-
right©1968 by W. W. Norton & Company, Inc.)
REFERENCES
Adams, G. F., 1957, Block diagrams from perspective grids: Journal of Geoscience Education, v. 5, no. 2,
p. 10–19.
Allan, U. S., 1989, Model for hydrocarbon migration and entrapment within faulted structures: American Asso-
ciation of Petroleum Geologists Bulletin, v. 73, p. 803–811.
Allaud, L., 1976, Vertical net sandstone determination for isopach mapping of hydrocarbon reservoirs: Ameri-
can Association of Petroleum Geologists Bulletin, v. 60, p. 2150–2153.
Allen, C. R., 1962, Circum-Pacific faulting in the Philippines-Taiwan region: Journal of Geophysical Research,
v. 67, p. 4795–4812.
Anspach, D. H., S. E. Tripp, R. E. Berlitz, and J. A. Gilreath, 1987, Postdevelopment analysis of producing
shelf-slope environments of deposition, High Island area: Transactions Gulf Coast Association of Geological
Societies, v. 37, p. 1–10.
Anstey, N. A., 1977, Seismic Interpretation: The Physical Aspects: Boston, International Human Resources De-
velopment Corporation Press, 637 p.
______, 1982, Simple Seismics for the Petroleum Geologist, the Reservoir Engineer, the Well-Log Analyst, the
Processing Technician, and the Man in the Field: Boston, International Human Resources Development Cor-
poration Press, 168 p.
Archie, G. E., 1942, The electrical resistivity log as an aid in determining some reservoir characteristics: Trans-
actions American Institute of Metallurgical Engineers, v. 146, p. 54–62.
______, 1947, Electrical resistivity an aid in core-analysis interpretation: American Association of Petroleum
Geologists Bulletin, v. 31, p. 350–366.
Atwater, G. I., and M. J. Foreman, 1959, Nature of growth of southern Louisiana salt domes and its effect on
petroleum accumulation: American Association of Petroleum Geologists Bulletin, v. 43, p. 2592–2622.
Atwater, G. I., and E. E. Miller, 1965, The effect of decrease in porosity with depth on future development of
oil and gas reserves in South Louisiana (Abs.): American Association of Petroleum Geologists Bulletin,
v. 49, p. 334.
Aydin, A., and A. Nur, 1982, Evolution of pull-apart basins and their scale independence: Tectonics, v. 1,
p. 91–105.
792
References 793
______, 1985, The type and role of stepovers in strike-slip tectonics: in Biddle, K. T. and N. Christie-Blick,
eds., Strike-Slip Deformation, Basin Formation, and Sedimentation: Society of Economic Paleontologists
and Mineralogists, Special Publication No. 37, p. 35–44.
Badgley, P. C., 1959, Structural Methods for the Exploration Geologist and a Series of Problems for Structural
Geology Students: New York, Harper & Bros., 280 p.
Badley, M. E., 1985, Practical Seismic Interpretation: Boston, International Human Resources Development
Corporation Press, 266 p.
Baldwin, B., and C. O. Butler, 1985, Compaction curves: American Association of Petroleum Geologists Bul-
letin, v. 69, p. 622–626.
Ball, S. M., 1982, Exploration application of temperatures recorded on log-headings – An up-the-odds method
of hydrocarbon-charged porosity prediction: American Association of Petroleum Geologists Bulletin, v. 66,
p. 1108–1123.
Bally, A. W., 1983, Seismic Expression of Structural Styles: American Association of Petroleum Geologists
Bulletin, Studies in Geology Series No. 15, v. 3., p. 167–172.
Bally, A. W., P. L. Gordy, and G. A. Stewart, 1966, Structure, seismic data, and orogenic evolution of southern
Canadian Rocky Mountains: Canadian Society of Petroleum Geologists Bulletin, v. 14, p. 337–381.
Banks, C. J., and J. Wharburton, 1986, “Passive-roof” duplex geometry in the frontal structures of the Kirthar
and Sulaiman Mountain Belts, Pakistan: Journal of Structural Geology, v. 8, p. 229–237.
Barnett, J. A. M., J. Mortimer, J. H. Rippon, J. J. Walsh, and J. Waterson, 1987, Displacement geometry in the
volume containing a single normal fault: American Association of Petroleum Geologists Bulletin, v. 71,
p. 925–937
Beckwith, R. H., 1941, Trace-slip faults: American Association of Petroleum Geologists Bulletin, v. 25, no. 12,
p. 2181–2193.
______, 1947, Fault problems in fault planes: Geological Society of America Bulletin, v. 58, p. 79–108.
Bell, W. G., 1956, Tectonic setting of Happy Springs and nearby structures in the Sweetwater Uplift area, cen-
tral Wyoming, in Geological Record of the American Association of Petroleum Geologists, Rocky Mountain
Section.
Bengston, C. A., 1980, Statistical curvature analysis methods for interpretation of dipmeter data: Oil and Gas
Journal, June 23 issue, p. 172–190.
Bennison, G. M., 1975, An Introduction to Geological Structures and Maps, third ed.: London, Edward Arnold
Ltd., 144 p.
Berg, O. R., and D. G. Woolverton, 1985, Seismic Stratigraphy II – an Integrated Approach to Hydrocarbon
Exploration: American Association of Petroleum Geologists Memoir 39, 276 p.
Berg, R. R., 1967, Point bar origin of Fall River sandstone reservoirs, northeastern Wyoming: Society of Petro-
leum Engineers, Paper 1953, 6 p.
Billings, M. P., 1954, Structural Geology, second ed.: New York, Prentice-Hall, 514 p.
______, 1972, Structural Geology, third ed.: Englewood Cliffs, NJ, Prentice-Hall, 606 p.
Biot, M. A., 1961, Theory of folding of stratified viscoelastic media and its implication in tectonics and oro-
genesis: Geological Society of America Bulletin, v. 72, p. 1595–1620.
Bischke, R. E., 1974, A model of convergent plate margins based on the Recent tectonics of Shikoku, Japan:
Journal of Geophysical Research, v. 79, p. 4845–4857.
______, 1990, Applied Structural Balancing: Gulf Coast Association of Geological Societies Short Course,
40th Annual Convention, Lafayette, LA.
______, 1994a, The compressional off-structure problem: Houston Geological Society Bulletin, May issue,
p. 29–34.
______, 1994b, Interpreting sedimentary growth structures from well log and seismic data (with examples):
American Association of Petroleum Geologists Bulletin, v. 7, p. 873–892.
______, 2002, Structural analysis of Ridge Basin, California: An example of extensional strike-slip folding, in
Crowell, J., ed., The Tectonics and Sedimentation of Ridge Basin, California: Geological Society of America
Memoir, in review.
Bischke, R. E., W. Finley and D. J. Tearpock, 1999, Growth analysis (∆d/d): Case histories of the resolution of
correlation problems as encountered while mapping around salt: Transactions Gulf Coast Association of Ge-
ological Societies, v. 49, p. 102–110.
Bischke, R. E., and J. Suppe, 1990a, Calculating sand/shale ratios from growth normal fault dips on seismic
profiles: Transactions Gulf Coast Association of Geological Societies, v. 40, p. 39–49.
Bischke, R. E., and J. Suppe, 1990b, Geometry of rollover: Origin of complex arrays of antithetic and synthetic
crestal faults (Abs.): American Association of Petroleum Geologists Bulletin, v. 74, p. 611.
794 References
Bischke, R. E., J. Suppe, and R. del Pilar, 1990, A new branch of the Philippine Fault system, as observed from
aeromagnetic and seismic data: Tectonophysics, v. 183, p. 243–264.
Bischke, R. E., and D. J. Tearpock, 1993, A method for estimating gross sand percentage and reservoir thick-
ness from seismic section: Example from Segno Field, Polk Co., Texas: Houston Geological Society Bul-
letin, May issue, p. 22–23, 40–43.
Bischke, R. E., and D. J. Tearpock, 1999, A graphical dip domain technique for projecting large growth faults
to depth using imaged hanging wall structure: Transactions Gulf Coast Association of Geological Societies,
v. 49, p. 112–120.
Bishop, M. S., 1960, Subsurface Mapping: New York, John Wiley & Sons, 198 p.
Blake, G. H., 1991, Review of the Neogene biostratigraphy of the Los Angeles Basin and implications for basin
evolution, in Biddle, K. T., ed., Active Margin Basins: American Association of Petroleum Geologists Mem-
oir 52, p. 135–184.
Bloomer, R. R., 1977, Depositional environments of a reservoir sandstone in west-central Texas: American As-
sociation of Petroleum Geologists Bulletin, v. 61, no. 3, p. 344–359.
Boeckelman, W., unpublished notes.
Boos, C. M., and M. F. Boos, 1957, Tectonics of eastern flank and foothills of Front Range, Colorado: Ameri-
can Association of Petroleum Geologists Bulletin, v. 41, no. 12, p. 2603–2676.
Bott, W. F., Jr., and T. T. Tieh, 1987, Diagenesis and high fluid pressures in the Frio and Vicksburg shales,
Brooks County, Texas: Transactions Gulf Coast Association of Geological Societies, v. 37, p. 323–334.
Bouvier, J. D., C. H. Kaars-Sijpesteijn, D. F. Kleusner, C. C. Onyejekwe, and R. C. Van Der Pal, 1989, Three-
dimensional seismic interpretation and fault sealing investigations, Nun River Field, Nigeria: American Asso-
ciation of Petroleum Geologists Bulletin, v. 73, p. 1397–1414.
Bower, T. H., 1947, Log map, new type of subsurface map: American Association of Petroleum Geologists
Bulletin, v. 127, no. 7, p. 340–349.
Boyer, S. E., 1986, Styles of folding within thrust sheets: examples from Appalachian and Rocky Mountains of
the U.S.A. and Canada: Journal of Structural Geology, v. 8, p. 325–339.
Boyer, S. E., and D. Elliott 1982, Thrust systems: American Association of Petroleum Geologists Bulletin,
v. 66, no. 9, p. 1196–1230.
Brabb, E. E., 1989, Geologic map of Santa Cruz County, California: U.S. Geological Survey Miscellaneous In-
vestigations Series Map I-1905, scale 1:62,500.
Branson, R. B., 1991, Productive trends and production history, South Louisiana and adjacent offshore, in
Goldthwaite, D., ed., An Introduction to Central Gulf Coast Geology: New Orleans Geological Society,
p. 61–70.
Brenneke, J. C., 1995, Analysis of fault traps: World Oil, December issue, p. 63–71.
Bristol, H. M., 1975, Structural geology and oil production of northern Gallatin County and southernmost
White County, Illinois: Illinois State Geological Survey, Illinois Petroleum 105, 20 p.
Bristol, H. M., and J. D. Treworgy, 1979, The Wabash Valley fault system in southeastern Illinois: Illinois State
Geological Survey, Circular 509, 19 p.
Broussard, M. J., and B. E. Lock, 1995, Modern analytical techniques for fault surface seal analysis: A Gulf
Coast case history: Transactions Gulf Coast Association of Geological Societies, v. 45, p. 87–93.
Brown, A. R., 1999, Interpretation of Three-Dimensional Seismic Data, fifth ed.: American Association of Pe-
troleum Geologists Memoir 42, 510 p.
Brown, G. G., D. L. Katz, et al., 1948, Natural Gasoline and Volatile Hydrocarbons: Tulsa, OK, National Gaso-
line Association of America.
Brown, L. F., and W. L. Fisher, 1984, Stratigraphic Interpretation and Petroleum Exploration: American Asso-
ciation of Petroleum Geologists Bulletin, Continuing Education Course Note Series No. 16.
Brown, R. L., 1981, Thickness Corrections for Deviated Wells – A Simplified Technique Utilizing the Apparent
Dip along a Well Path: Contribution to the 1981 Geological/Petrophysical Conference, Houston, TX.
Brown, R. N., 1980, History of exploration and discovery of Morgan, Ramadan and July Oil Fields, Gulf of
Suez, Egypt, in Miall, A. D., ed., Facts and Principles of World Petroleum Occurrence: Canadian Society of
Petroleum Geologists Memoir 6, p. 733–764.
Brown, W. G., 1982, New Tricks for Old Dogs – A Short Course in Structural Geology: American Association
of Petroleum Geologists, Southwest Section Meeting, Wichita Falls, TX, 79 p.
______, 1984a, Working with Folds: American Association of Petroleum Geologists, Structural Geology
School Course Notes.
______, 1984b, Basement Involved Tectonics, Foreland Areas: American Association of Petroleum Geologists,
Continuing Education Course Series No. 26, 92 p.
References 795
Bruce, C. H., 1973, Pressured shale and related sediment deformation: Mechanism for development of regional
contemporaneous faults: American Association of Petroleum Geologists Bulletin, v. 57, p. 878–886.
Brune, J. N., T. L. Henyey, and R. F. Roy, 1969, Heat flow, stress and rate of slip on the San Andreas Fault,
California: Journal of Geophysical Research, v. 74, p. 3821–3827.
Bucher, W. H., 1933, The Deformation of the Earth’s Crust: Princeton, NJ, Hafner Publishing Company, 518 p.
Busch, D. A., 1959, Prospecting for stratigraphic traps: American Association of Petroleum Geologists Bul-
letin, v. 43, no. 12, p. 2829–2843.
______, 1974, Stratigraphic Traps in Sandstones – Exploration Techniques: American Association of Petro-
leum Geologists Memoir 21, 174 p.
Busch, D. A., and D. A. Link, 1985, Exploration Methods for Sandstone Reservoirs: Tulsa, OK, Oil and Gas
Consultants, Inc., 327 p.
Busk, H. G., 1929, Earth Flexures; Their Geometry and Their Representation and Analysis in Geological Sec-
tion, with Special Reference to the Problem of Oil Finding: Cambridge, Cambridge University Press, 106 p.
______, 1957, Earth Flexures, Their Geometry and Their Representation and Analysis in Geological Section,
with Special Reference to the Problem of Oil Finding: New York, William Trussell, 106 p.
Cape, D. C., S. McGeary, and G. A. Thompson, 1983, Cenozoic normal faulting and the shallow structure of
the Rio Grande Rift near Socorro, New Mexico: Geological Society of America Bulletin, v. 94, p. 3–14.
Challinor, J., 1933, The “throw” of a fault: Geological Magazine, v. 70, p. 384–393.
Chamberlin, T. C., 1897, The method of multiple working hypothesis: Journal of Geology, v. 5, p. 837–848.
Chamberlin, R. T., 1910, The Appalachian folds of central Pennsylvania: Journal of Geology, v. 18,
p. 228–251.
Chatellier, J-Y., and C. Porras, 2001, The Multiple Bischke Plot Analysis, a simple and powerful graphic tool to
integrate stratigraphic studies: American Association of Petroleum Geologists, Annual Convention Program
Book, p. A34-A35.
Chenoweth, P. A., 1972, Unconformity Traps, in King, R. E., ed., Stratigraphic Oil and Gas Fields – Classifica-
tion, Exploration Methods, and Case Histories: American Association of Petroleum Geologists Memoir 16,
p. 42–46.
Chinnery, M. A., 1963, The stress changes that accompany strike-slip faulting: Bulletin of the Seismological
Society of America, v. 53, p. 921–932.
Christensen, A. F., 1983, An example of a major syndepositional listric fault, in Bally, A. W., ed., Seismic Ex-
pression of Structural Styles: American Association of Petroleum Geologists Bulletin, Studies in Geology
No. 15, v. 2, p. 2.3.1–36 to 2.3.1–40.
Clark, J. C., 1959, Problems of fault nomenclature: American Association of Petroleum Geologists Bulletin,
v. 43, no. 11, p. 2653–2674.
Clark, J. C., and J. D. Reitman, 1973, Oligocene stratigraphy, tectonics, and paleogeography southwest of the
San Andreas Fault, Santa Cruz Mountains and Gabilan Range, California Coastal Ranges: United States Ge-
ological Survey, Professional Paper 783, 18 p.
Clark, M. M., K. K. Lienkaemper, J. J. Harwood, D. S. Lajoie, K. R. Matti, J. C. Perkins, J. A. Rymer, A. M.
Sarna-Wojcicki, A. M. Sharp, J. D. Sims, J. C. Tinsley, and J. I. Ziony, 1984, Preliminary slip-rate table and
map of late-Quaternary faults of California: United States Geological Survey Open-File Report 84–106, 12 p.
Clark, S. K., 1943, Classification of faults: American Association of Petroleum Geologists Bulletin, v. 27,
p. 1245–1265.
Cloos, E., 1942, Distortion of stratigraphic thickness due to folding: Proceedings of the National Academy of
Science, v. 28, no. 10, p. 401–407.
______, 1968, Experimental analysis of Gulf Coast fracture patterns: American Association of Petroleum Ge-
ologists Bulletin, v. 52, p. 420–444.
Coates, J., 1945, The construction of geological sections: The Quarterly Journal, The Geological Mining and
Metallurgical Society of India, v. 17, p. 1–11.
Cole, W., 1969, Reservoir Engineering Manual, second ed.: Houston, Gulf Publishing Company, 385 p.
Coleman, J. M., 1982, Deltas: Processes of Deposition and Models for Exploration: Boston, International
Human Resources Development Corporation, 124 p.
______, 1988, Dynamic changes and processes in the Mississippi River Delta: Geological Society of America
Bulletin, v. 100, p. 999–1015.
Craft, B. C., and M. F. Hawkins, 1959, Applied Petroleum Reservoir Engineering: Englewood Cliffs, NJ, Pren-
tice Hall, 437 p.
Crans, W., G. Mandl, and J. Haremboure, 1980, On the theory of growth faulting: A geomechanical delta
model based on gravity sliding: Journal of Petroleum Geology, v. 2, p. 265–307.
796 References
Crowell, J. C., 1948, Template for spacing structure contours: American Association of Petroleum Geologists
Bulletin, v. 32, p. 2290–2294.
______, 1950, Geology of Hungry Valley area, southern California: American Association of Petroleum Geolo-
gists Bulletin, v. 34, p. 1623–1646.
______, 1959, Problems in fault nomenclature: American Association of Petroleum Geologists Bulletin, v. 43,
p. 2653–2674.
______, 1962, Displacement along the San Andreas fault, California: Geological Society of America Special
Paper 71, 61 p.
______, 1974a, Origin of late Cenozoic basins in southern California, in Dickinson, W. R., ed., Tectonics and
Sedimentation: Society of Economic Paleontologists and Mineralogists, Special Publication No. 22,
p. 190–204.
______, 1974b, Sedimentation along the San Andreas Fault, California, in Dott, R. H., Jr. and Shaver, R. H.,
eds., Modern and Ancient Geosynclinal Sedimentation: Society of Economic Paleontologists and Mineralo-
gists, Special Publication No. 19, p. 292–303.
______, 1982, The tectonics of Ridge Basin, Southern California, in Crowell, J. C., and M. Link, eds., Geo-
logic History of Ridge Basin, Southern California: Pacific Section, Society of Economic Paleontologists and
Mineralogists, p. 25–42.
______, 2002a, Introduction to geology of Ridge Basin, southern California, in Crowell, J. C., ed., Evolution of
Ridge Basin, Southern California: An Interplay of Sedimentation and Tectonics: Geological Society of
America Memoir, in review.
______, 2002b, Overview of rocks bordering Ridge Basin, southern California, in Crowell, J. C., ed., Evolution
of Ridge Basin, Southern California: An Interplay of Sedimentation and Tectonics: Geological Society of
America Memoir, in review.
______, 2002c, Tectonics of Ridge Basin, southern California, in Crowell, J. C., ed., Evolution of Ridge Basin,
Southern California: An Interplay of Sedimentation and Tectonics: Geological Society of America Memoir,
in review.
Crowell, J. C. and M. H. Link, eds., 1982, Geologic history of Ridge Basin, southern California: Pacific Sec-
tion, Society of Economic Paleontologists and Mineralogists, 304 p.
Currie, J. B., 1952, Three-dimensional method for solution of oil field structures: American Association of Pe-
troleum Geologists Bulletin, v. 36, p. 889–890.
______, 1956, Role of concurrent deposition and deformation of sediments in development of salt-dome
graben structures: American Association of Petroleum Geologists Bulletin, v. 40, p. 1–16.
Dahlen, F. A., and J. Suppe, 1988, Mechanics, growth, and erosion of mountain belts: Geological Society of
America, Special Paper 218, p. 161–178.
Dahlen, F. A., J. Suppe, and D. Davis, 1984, Mechanics of fold-and-thrust belts and accretionary wedges: Co-
hesive Coulomb theory: Journal of Geophysical Research, v. 89, p. 10,087–10,101.
Dahlstrom, C. D. A., 1969, Balanced cross-sections: Canadian Journal of Earth Sciences, v. 6, p. 743–757.
______, 1970, Structural geology in the eastern margin of the Canadian Rocky Mountains: Bulletin of Cana-
dian Petroleum Geology, v. 18, p. 332–406.
Dart, R. L., and H. S. Swolfs, 1992, Subparallel faults and horizontal-stress orientations: An evaluation of in-
situ stresses inferred from elliptical wellbore enlargements, in Larsen, R. M., H. Brekke, and B. T. Larson,
eds., Structural and Tectonic Modeling and Its Applications to Petroleum Geology: Norwegian Petroleum
Society, Special Publication No. 1, p. 519–529.
Davis, D., and T. Engelder, 1985, The role of salt in fold-and-thrust belts: Tectonophysics, v. 119, p. 67–88.
Davis, D., J. Suppe, and F. A. Dahlen, 1983, Mechanics of fold-and-thrust belts and accretionary wedges: Jour-
nal of Geophysical Research, v. 88, p. 1153–1172.
Davis, T. L., 1987, Seismic facies analysis: pitfalls and applications in cratonic basins: The Leading Edge, v. 6,
p. 18–23, 37.
Davison, I., 1986, Listric normal fault profiles: Calculation using bed-length balance and fault displacement:
Journal of Structural Geology, v. 8, p. 209–210, p. 1560–1580.
______, 1987, Normal fault geometry related to sediment compaction and burial: Journal of Structural Geol-
ogy, v. 9, p. 393–401.
Dawers, N. H., and J. R. Underhill, 2000, The role of fault interaction and linkage in controlling synrift strati-
graphic sequences: Late Jurassic, Statfjord East Area, North Sea: American Association of Petroleum Geolo-
gists Bulletin, v. 84, p. 45–64.
De Paor, D. G., and G. Eisenstadt, 1987, Stratigraphy and structural consequences of fault reversal: An exam-
ple from the Franklinian Basin, Ellesmere Island: Geology, v. 15, p. 948–949.
References 797
Dennis, J. G., 1972, Structural Geology: New York, Ronald Press, 532 p.
Dennison, J. M., 1968, Analysis of Geologic Structures: New York, W. W. Norton & Co., 209 p.
Dickey, P. A., 1981, Petroleum Development Geology, second ed.: Tulsa, OK, PennWell Publishing Co., 428 p.
______, 1986, Petroleum Development Geology, third ed.: Tulsa, OK, PennWell Publishing Co., 530 p.
Dickinson, G., 1954, Subsurface interpretation of intersecting faults and their effects upon stratigraphic hori-
zons: American Association of Petroleum Geologists Bulletin, v. 38, no. 5, p. 854–877.
Dickinson, W. R., and D. R. Seely, 1979, Structure and stratigraphy of forearc regions: American Association
of Petroleum Geologists Bulletin, v. 63, p. 2–31.
Dixon, J. M., 1975, Finite strain and progressive deformation in models of diapiric structures: Tectonophysics,
v. 28, p. 89–124.
Dobrovolny, J. S., 1951, Descriptive geometry for geologists: American Association of Petroleum Geologists
Bulletin, v. 35, p. 1674–1686.
Doebl, F., and R. Teichmuller, 1979, Zur Geologie und heutigen Geothermik mittleren Oberrhein-Graben:
Fortschritte in der Geologie von Rheinland und Westfalen, v. 27, p. 1–17.
Dolly, E. D., Geological techniques utilized in Trap Spring Field discovery, Railroad Valley, Nye County,
Nevada, in Newman, G. W., and H. D. Goode, eds., Basin and Range Symposium and Great Basin Field Con-
ference: Rocky Mountain Association of Geologists and Utah Geological Association, p. 455–467.
Donath, F. A., and R. B. Parker, 1964, Folds and folding: Geological Society of America Bulletin, v. 75,
p. 45–62.
Donn, W. L., and J. A. Shimer, 1958, Graphic Methods in Structural Geology, New York, Appleton-Century-
Crofts, 180 p.
Doveton, J. H., 1986, Log Analysis of Subsurface Geology: New York, John Wiley & Sons, 273 p.
Dula, W. F. Jr., 1991, Geometric models of listric normal faults and rollover folds: American Association of Pe-
troleum Geologists Bulletin, v. 75, p. 1609–1625.
Duran, L. G., 1951, Trigonometric and graphical solution of problems in structural mapping: World Oil, v. 133,
no. 6, p. 94–98.
Dury, G. H., 1952, Map Interpretation: London, Pittman, 203 p.
Eardley, A. J., 1938, Graphic treatment of folds in three dimensions: American Association of Petroleum Geol-
ogists Bulletin, v. 22, p. 483–489.
Eastman Christensen, Magnetic single-shot operations manual: Salt Lake City, UT, Eastman Christensen, 38 p.
Eastman Whipstock, Introduction to directional drilling: Houston, TX, Eastman Whipstock, 38 p.
Ebanks, W. J., Jr., and J. F. Weber, 1982, Development of a shallow heavy-oil deposit in Missouri: Oil and Gas
Journal, September 27 issue, p. 222–234.
Edwards, M. B., 2001, Wilcox Regional Study – South Texas Zapata Phase: Unpublished multi-client report.
Elliott, D., 1976, The energy balance and deformation mechanisms of thrust sheets: Philosophical Transactions
of the Royal Society of London, Series A, v. 283, p. 289–312.
______, 1983, The construction of balanced cross sections: Journal of Structural Geology, v. 5, p. 101.
Emery, K. O., 1980, Continental margins – classification of petroleum prospects: American Association of Pe-
troleum Geologists Bulletin, v. 64, p. 297–315.
Engelder, T., and R. Engelder, 1977, Fossil distortion and decollement tectonics of the Appalachian Plateau:
Geology, v. 5, p. 457–460.
Ewing, T. E., 1983, Growth faults and salt tectonics in the Houston Diapir Province – relative timing and explo-
ration significance: Transactions Gulf Coast Association of Geological Societies, v. 33, p. 83–90.
Fagin, R. A., J. E. Trusty, L. R. Emmet, and M. K. Mayo, 1991, MWD resistivity tool guides bit horizontally in
thin bed: Oil & Gas Journal, December 9 issue, p. 62–65.
Faill, R. T., 1969, Kink band structure in the Valley and Ridge Province, central Pennsylvania: Geological Soci-
ety of America Bulletin, v. 80, p. 2539–2550.
______, 1973, Kink band folding, Valley and Ridge Province, Pennsylvania: Geological Society of America
Bulletin, v. 84, p. 1289–1314.
Fallaw, W. C., 1973, Grabens on anticlines in Gulf Coastal Plain, and thinning of sedimentary section in down-
thrown block: American Association of Petroleum Geologists Bulletin, v. 57, p. 198–203.
Ferrill, D. A., J. A. Stamatakos, and D. Simmons, 1999, Normal fault corrugation: Implications for growth and
seismicity of active normal faults: Journal of Structural Geology, v. 21, p. 1027–1038.
Fisher, W. L., and J. H. McGowen, 1967, Depositional systems in the Wilcox Group of Texas and their relation-
ship to occurrence of oil and gas: Transactions Gulf Coast Association of Geological Societies, v. 27,
p. 105–125.
798 References
Flinn, D., 1962, On folding during three-dimensional progressive deformation: Quarterly Journal of the Geo-
logical Society, v. 118, p. 385–433.
Forgotson, J. M., Jr., 1960, Review and classification of quantitative mapping techniques: American Associa-
tion of Petroleum Geologists Bulletin, v. 44, p. 83–100.
Fox, F. G., 1959, Structure and accumulation of hydrocarbons in southern Foothills, Alberta, Canada: Ameri-
can Association of Petroleum Geologists Bulletin, v. 43, p. 992–1025.
Frey, M. G., and W. H. Grimes, 1970, Bay Marchand-Timbalier Bay-Callou Island Salt Complex, Louisiana: in
Halbouty, M. T., ed., Giant Oil and Gas Fields of the Decade 1968–1978: American Association of Petroleum
Geologists Memoir 14, p. 277–291.
Gatewood, L. E., 1970, Oklahoma City Field – anatomy of a giant: in Halbouty, M. T., ed., Giant Oil and Gas
Fields of the Decade 1968–1978: American Association of Petroleum Geologists Memoir 14, p. 223–254.
Geikie, A., 1903, Text-Book of Geology, fourth ed.: New York, Macmillan and Company, v. 1.
Geiser, P. A., 1988, The role of kinematics in the construction and analysis of geological cross sections in de-
formed terrains: in Mitra, G., and S. Wojtal, eds., Geometrics and Mechanisms of Thrusting, with Special
Reference to the Appalachians: Geological Society of America, Special Paper 222, p. 47–76.
Ghignone, J. I., and G. Deandrade, 1970, General geology and major oilfields of Reconcavo Basin, Brazil: in
Halbouty, M. T., ed., Giant Oil and Gas Fields of the Decade 1968–1978: American Association of Petroleum
Geologists Memoir 14, p. 337–358.
Gibbs, A. D., 1983, Balanced cross section construction from seismic sections in areas of extensional tectonics:
Journal of Structural Geology, v. 5, p. 153–160.
______, 1984, Structural elevation of extensional basin margins: Quarterly Journal of the Geological Society,
v. 141, p. 609–620.
Giles, A. B., and D. H. Wood, 1983, Oakwood salt dome, east Texas: Geologic framework, growth history, and
hydrocarbon production: Texas Bureau of Economic Geology, Geological Circular 83–1, 55 p.
Gill, W. D., 1953, Construction of geological sections of folds with steep-limb attenuation: American Associa-
tion of Petroleum Geologists Bulletin, v. 37, p. 2389–2406.
Goguel, J., 1962, Tectonics, second ed.: San Francisco, W. H. Freeman and Co., 384 p.
Gordy, P. L., and F. R. Frey, 1975, Geological cross sections through the Foothills: Foothills Fieldtrip Guide
Book, Canadian Society of Petroleum Geologists/Canadian Society of Exploration Geophysicists.
Gordy, P. L., F. R. Frey, and D. K. Nossin, 1977, Geological guide for the Waterton-Glacier Park Field Confer-
ence: Canadian Society of Petroleum Geologists, 93 p.
Gow, K. L., 1962, Segno Field, Polk County, Texas, in Denham, R. L., ed., Typical Oil and Gas Fields of
Southeast Texas: Houston, TX, Houston Geological Society, p. 187–191.
Gries, R. R., and R. C. Dyer, eds., 1985, Seismic Exploration of the Rocky Mountain Region: Rocky Mountain
Association of Geologists and Denver Geophysical Society, 300 p.
Groshong, R. H., Jr., 1975, “Slip” cleavage caused by pressure solution in a bucklefold: Geology, v. 3,
p. 411–413.
______, 1989, Half graben structures: balanced models of extensional fault-bend folds: Geological Society of
America Bulletin, v. 101, p. 96–105.
Guiraud, M., and M. Seguret, 1985, A releasing solitary overstep model for the Late Jurassic – Early Cretaceous
Soria strike-slip basin (northern Spain), in Biddle, K. T., and N. Christie-Blick, Strike-Slip Deformation,
Basin Formation, and Sedimentation: Society of Economic Paleontologists and Mineralogists, Special Publi-
cation No. 37, p. 159–175.
Gwinn, V. E., 1964, Thin-skinned tectonics in the Plateau and northwestern Valley and Ridge province of the
Central Appalachians: Geological Society of America Bulletin, v. 75, p. 863–900.
Halbouty, M. T., 1967, Salt domes: Gulf Region, United States and Mexico: Houston, Gulf Publishing Com-
pany, 425 p.
______, 1979, Salt Domes: Gulf Region, United States and Mexico, second ed.: Houston, Gulf Publishing
Company, 561 p.
______, 1982, The deliberate search for the subtle trap: American Association of Petroleum Geologists Mem-
oir 32, 351 p.
Hamblin, W. K., 1965, Origin of “reverse drag” on the downthrown side of normal faults: Geological Society
of America Bulletin, v. 76, p. 1145–1164.
Handley, E. J., 1954, Contouring is important: World Oil, v. 138, p. 106–107.
Harding, T. P., 1973, Newport-Inglewood Trend, California – an example of wrenching style of deformation:
American Association of Petroleum Geologists Bulletin, v. 57, p. 97–116.
References 799
______, 1974, Petroleum traps associated with wrench faults: American Association of Petroleum Geologists
Bulletin, v. 58, p. 1290–1304.
______, 1976, Predicting productive trends related to wrench faults: World Oil, v. 182, no. 7, p. 64–69.
______, 1976, Tectonic significance and hydrocarbon trapping consequences of sequential folding synchro-
nous with San Andreas faulting, San Joaquin Valley, California: American Association of Petroleum Geolo-
gists Bulletin, v. 60, p. 356–378.
______, 1984, Graben hydrocarbon occurrences and structural style: American Association of Petroleum Geol-
ogists Bulletin, v. 68, p. 333–362.
______, 1985, Seismic characteristics and identification of negative flower structures, positive flower struc-
tures, and positive structural inversion: American Association of Petroleum Geologists Bulletin, v. 69, p.
582–600.
______, 1990, Identification of wrench faults using subsurface structural data: Criteria and pitfalls: American
Association of Petroleum Geologists Bulletin, v. 74, p. 1590–1609.
Harding, T. P., and J. D. Lowell, 1979, Structural styles, their plate-tectonic habitats, and hydrocarbon traps in
petroleum provinces: American Association of Petroleum Geologists Bulletin, v. 63, p. 1016–1058.
Harrington, J. W., 1951, The elementary theory of subsurface structural contouring: Transactions American
Geophysical Union, v. 32, p. 77–80.
Hartman, R. R., D. J. Teskey, and J. L. Friedberg, 1971, A system for rapid digital aeromagnetic interpretation:
Geophysics, v. 36, p. 891–918.
Hay, J. T. C., 1978, Structural development in the northern North Sea: Journal of Petroleum Geology, v. 1,
p. 65–77.
Heisey, E. L., ed., 1977, Rocky Mountain Thrust Belt Geology and Resources: Casper, WY, Wyoming Geologi-
cal Society, 787 p.
Herold, S. C., 1933, Projection of dip angle on profile section: American Association of Petroleum Geologists
Bulletin, v. 17, p. 740–742.
Heybroek, P., U. Haanstra, and D. A. Erdman, 1967, Observations on the geology of the North Sea area: Pro-
ceedings of the 7th World Petroleum Congress, v. 2, p. 905–916.
Hill, M. L., 1942, Graphic method for some geologic calculations: American Association of Petroleum Geolo-
gists Bulletin, v. 26, p. 1155–1159.
______, 1947, Classification of faults: American Association of Petroleum Geologists Bulletin, v. 31,
p. 1669–1673.
______, 1959, Dual classification of faults: American Association of Petroleum Geologists Bulletin, v. 43,
p. 217–237.
Hill, M. L., and T. W. Dibblee, 1953, San Andreas, Garlock, and Big Pine faults, California (A study of the
character, history, and tectonic significance of their displacements): Geological Society of America Bulletin,
v. 64, p. 443–458.
Hills, E. S., 1963, Elements of Structural Geology: London, Methuen & Co., 483 p.
Hooper, N. J., J. G. M. Raven, and M. J. Kilpatrick, 1992, Computer modeling of multiple surfaces with faults:
The Ivanhoe Field, Outer Moray Firth Basin, U.K. North Sea, in Hamilton, D. E., and T. A. Jones, eds.,
American Association of Petroleum Geologists Bulletin, Computer Applications in Geology No. 1,
p. 161–174.
Horsfield, W. T., 1977, An experimental approach to basement-controlled faulting: Geologie en Mijnbouw,
v. 56, p. 363–370.
______, 1980, Contemporaneous movement along crossing conjugate normal faults: Journal of Structural Ge-
ology, v. 2, p. 305–310.
Hossack, J. R., 1983, A cross-section through the Scandinavian Caledonides constructed with the aid of
branch-line maps: Journal of Structural Geology, v. 5, p. 103–111.
Hubbert, M. K., 1937, Theory of scale models as applied to the study of geologic structures: Geological Soci-
ety of America Bulletin, v. 77, p. 1247–1264.
Hubbert, M. K., and W. W. Rubey, 1959, Role of fluid pressure in mechanics of overthrust faulting: Geological
Society of America Bulletin, v. 70, p. 115–166.
Hull, C. E., and H. R. Warman, 1968, Asmari Oil Fields of Iran: London, British Petroleum Company,
p. 428–437.
Illies, J. H., 1981, Mechanism of graben formation: Tectonophysics, v. 73, p. 249–266.
Ivanhoe, L. F., 1956, Integration of geological data on seismic sections: American Association of Petroleum
Geologists Bulletin, v. 40, p. 1016–1023.
800 References
Jackson, M. P. A., and W. E. Galloway, 1984, Structural and depositional styles of Gulf Coast Tertiary conti-
nental margins: Application to hydrocarbon exploration: American Association of Petroleum Geologists Bul-
letin, Continuing Education Course Note Series No. 25, 226 p.
Jackson, M. P. A., and S. J. Seni, 1983, Geometry and evolution of salt structures in a marginal rift basin of the
Gulf of Mexico, east Texas: Geology, v. 11, p. 131–135.
Jaeger, J. C., 1962, Elasticity, Fracture and Flow with Engineering and Geological Application: London,
Methuen, p. 268.
Jaeger, J. C., and N. G. W., Cook, 1969, Fundamentals of rock mechanics: London, Methuen, 513 p.
Jaeger, L. G., 1966, Cartesian Tensors in Engineering Science: New York, Pergamon Press, 166 p.
Jageler, A. H., and D. R. Matuszak, 1972, Use of well logs and dipmeters in stratigraphic-trap exploration:
American Association of Petroleum Geologists Memoir 16, p. 107–135.
James, D. M. D., ed., 1984, The geology and hydrocarbon resources of Negora Brunei Darussalam: Bandar
Seri Begawan, Muzium Brunei, 169 p.
Jenyon, M. K., 1988, Fault-salt wall relationships, southern North Sea: Oil & Gas Journal, September 5 issue,
p. 76–81.
Johanson, D. B., 1987, Structural evolution of Grand Lake Field, Cameron Parish, Louisiana: Transactions
Gulf Coast Association of Geological Societies, v. 37, p. 113–122.
Jones, P. B., 1971, Folded faults and sequence of thrusting in Alberta Foothills: American Association of Petro-
leum Geologists Bulletin, v. 55, p. 292–306.
______, 1982, Oil and gas beneath east-dipping underthrust faults in the Alberta Foothills: Rocky Mountain
Association of Geologists, v. 1, p. 61–74.
______, 1988, Balanced cross-sections – an aid to structural interpretation: The Leading Edge, v. 7, no. 8,
p. 29–31.
Katz, D. L., et al., 1966, How water displaces gas from porous media: Oil and Gas Journal, January 10 issue,
p. 55–60.
Kay, M., 1945, Paleogeographic and palinspastic maps: American Association of Petroleum Geologists Bul-
letin, v. 29, p. 426–450.
______, 1954, Isolith, isopach, and palinspastic maps (Geological Note): American Association of Petroleum
Geologists Bulletin, v. 38, p. 916–917.
Kelley, V., 1971, Geology of Pecos Country, southern New Mexico: New Mexico Bureau of Mines and Mineral
Resources, Memoir 24, 79 p.
Kennedy, W. Q., 1946, The Great Glen Fault: Quarterly Journal of the Geological Society, v. 102, p. 41–76.
Kerans, C., and H. S. Nance, 1991, High frequency cyclicity and regional depositional patterns of the Grayburg
Formation, Guadalupe Mountains, New Mexico, in Meader-Roberts, S. J., M. P. Candelaria, and G. E. Moore
eds., Sequence Stratigraphy, Facies, and Reservoir Geometries of the San Andres, Grayburg and Queen For-
mations, Guadalupe Mountains, New Mexico: Society of Economic Paleontologists and Mineralogists, Spe-
cial Publication 91–32, p. 53–70.
Klein, G. D., 1985 Sandstone Depositional Models for Exploration for Fossil Fuels, third ed.: Boston, Interna-
tional Human Resources Development Corporation, 209 p.
Krumbein, W. C., 1942, Criteria for subsurface recognition of unconformities: American Association of Petro-
leum Geologists Bulletin, v. 26, p. 36–62.
______, 1948, Lithofacies maps and regional sedimentary stratigraphic analysis: American Association of Pe-
troleum Geologists Bulletin, v. 32, p. 1909–1923.
______, 1952, Principles of facies map interpretation: Journal of Sedimentary Petrology, v. 22, p. 200–211.
Kuhme, A. K., 1987, Seismic interpretation of reefs: The Leading Edge, v. 6, no. 8, p. 60–65.
Kyte, D. G., D. N. Meehan, and T. R. Svor, 1994, Method of maintaining a borehole in a stratigraphic zone dur-
ing drilling: United States Patent 5,311,951 (May 17, 1994).
Lafayette Geological Society, 1964, Typical oil and gas fields of Southwestern Louisiana: Lafayette, LA,
Lafayette Geological Society, v. 1, 41 p.
______, 1970, Typical oil and gas fields of Southwestern Louisiana: Lafayette, LA, Lafayette Geological Soci-
ety, v. 2, 30 p.
______, 1973, Offshore Louisiana Oil & Gas Fields: Lafayette, Louisiana, LA Geological Society, v. 3, 30 p.
Lamb, C. F., 1980, Painter Reservoir Field – giant in Wyoming Thrust Belt: American Association of Petro-
leum Geologists Bulletin, v. 64, p. 638–373.
Lamerson, P. R., 1982, The Fossil Basin and its relationship to the Absaroka thrust system, Wyoming and Utah,
in Powers, R. B., ed., Geologic Studies of the Cordilleran Thrust Belt: Denver, CO, Rocky Mountain Associ-
ation of Geologists, p. 279–340.
References 801
Langstaff, C. S., and D. Morrill, 1981, Geologic Cross Sections: Boston, International Human Resources De-
velopment Corporation Press, 108 p.
Laubscher, H. P., 1961, Die Fernschilbhypothese der Jurafaltilng: Eclogae Geologicae Helvetiae, v. 54,
p. 222–282.
______, 1977, Fold Development in the Jura: Tectonophysics, v. 37, p. 337–362.
Leake, J. and F. Shray, 1991, Logging while drilling keeps horizontal well on small target: Oil & Gas Journal,
September 23 issue, p. 53–59.
LeBlanc, R. J., 1972, Geometry of sandstone reservoir bodies, in Cook, T. D., ed., Symposium on Underground
Waste Management and Environmental Implications: American Association of Petroleum Geologists Memoir
18, p. 133–190.
Lelek, J. J., 1982, Geologic factors affecting reservoir analysis, Anschutz Ranch East Field, Utah-Wyoming:
Society of Petroleum Engineers 15th Annual Offshore Technology Conference, Society of Petroleum Engi-
neers Paper 10992.
Leroy, L. W., ed., 1950, Subsurface Geologic Methods – A Symposium, second ed.: Golden, CO, Colorado
School of Mines, 1156 p.
Leroy, L. W., D. O. Leroy, and J. W. Raese, eds., 1977, Subsurface Geology, fourth ed.: Golden, CO, Colorado
School of Mines, 941 p.
Leroy, L. W., D. O. Leroy, J. W. Raese, and S. D. Schwochow, eds., 1987, Subsurface Geology, fifth ed.:
Golden, CO, Colorado School of Mines, 1002 p.
Leroy, L. W., and J. W. Low, 1954, Graphic Problems in Petroleum Geology: New York, Harper and Bros., 238
p.
Levorsen, A. I., 1943, Discovery thinking: American Association of Petroleum Geologists Bulletin, v. 27,
p. 887–928.
______, 1954, Geology of Petroleum: San Francisco, W. H. Freeman and Co., 703 p.
Ley, H. H., 1930, Structure contouring: American Association of Petroleum Geologists Bulletin, v. 14,
p. 103–105.
Lindsay, R. F., 1991, Grayburg Formation (Permian-Guadalpian): Comparison of reservoir characteristics and
sequence stratigraphy in the Northwestern Central Basin Platform with outcrops in the Guadalupe Moun-
tains, New Mexico, in Meader-Roberts, S. J., M. P. Candelaria, and G. E. Moore eds., Sequence Stratigraphy,
Facies, and Reservoir Geometries of the San Andres, Grayburg and Queen Formations, Guadalupe Moun-
tains, New Mexico: Society of Economic Paleontologists and Mineralogists, Special Publication 91–32,
p. 111–118.
Link, M., 2002, Sedimentation of Ridge Basin, in Crowell, J. C., ed., The Tectonics and Sedimentation of
Ridge Basin, California: Geological Society of America Memoir, in review.
Link, M. H., and R. H. Osborne, 1982, Sedimentary facies of Ridge Basin, southern California, in Crowell,
J. C., and M. H. Link, Geologic History of Ridge Basin, Southern California: Society of Economic Paleontol-
ogists and Mineralogists, Pacific Section, p. 63–78.
Link, M. H., C. K. Taylor, N. G. Muñoz, J. E. Bueno, and P. J. Muñoz, 1996, 3-D seismic examples from Cen-
tral Lake Maracaibo, Maraven’s Block 1 Field, Venezuela, in Weimer, P., and T. L. Davis, eds., Application of
3-D Seismic Data to Exploration and Production: American Association of Petroleum Geologists, Studies in
Geology No. 42, and Society of Exploration Geophysicists Geophysical Developments Series No. 5: Tulsa,
OK, American Association of Petroleum Geologists and Society of Exploration Geophysicists, p. 69–82.
Link, P. K., 1982, Basic Petroleum Geology: Tulsa, OK, Oil and Gas Consultants International, Inc., 235 p.
Link, T. A., 1949, Interpretations of foothills structures, Alberta, Canada: American Association of Petroleum
Geologists Bulletin, v. 33, p. 1475–1501.
Lisowski, M., W. H. Prescott, J. C. Savage, and M. J. S. Johnston, 1990, Geodetic estimate of coseismic slip
during the 1989 Loma Prieta, California, earthquake: Geophysical Research Letters, v. 17, p. 1437–1440.
Lister, G. S., M. A. Etheridge, and P. A. Symonds, 1986, Detachment faulting and the evolution of passive con-
tinental margins: Geology, v. 14, p. 246–250.
Lock, B. E., 1989, Subsurface Geological lnvestigations, unpublished course notes.
Lock, B. E., and S. L. Voorhies, 1988, Sequence stratigraphy as a tool for interpretation of the Cockfield/Yegua
in Southwest Louisiana: Transactions Gulf Coast Association of Geological Societies, v. 39, p. 123–131.
Low, J. W., 1951, Subsurface maps and illustrations, in Leroy, L. W., ed., Subsurface Geologic Methods — A
Symposium, second ed.: Golden, CO, Colorado School of Mines, p. 894–969.
Lowell, J. D., 1985, Structural Styles in Petroleum Exploration: Golden, CO, Oil & Gas Consultants Interna-
tional, 460 p.
Lyle, H. N., 1951, Southwest Texas faults: Oil & Gas Journal, v. 49, no. 41, p. 108–112.
802 References
Mackenzie, D. B., 1972, Primary stratigraphic traps in sandstone: in King, R. E., ed., Stratigraphic Oil and Gas
Fields – Classification, Exploration Methods, and Case Histories: American Association of Petroleum Geolo-
gists Memoir 16, p. 47–66.
Mackin, J. H., 1950, The down-structure method of viewing geologic maps: Journal of Geology, v. 58,
p. 55–77.
Macurda, D. B., Jr., 1987, Seismic interpretation of transgressive and progradational sequence: The Leading
Edge, v. 6, no. 4, p. 18–21.
Malhase, J., 1927, Constructing geologic sections with unequal scales: American Association of Petroleum Ge-
ologists Bulletin, v. 11, p. 755–757.
Mannhard, G. W., and D. A. Busch, 1974, Stratigraphic trap accumulation in southwestern Kansas and north-
western Oklahoma: American Association of Petroleum Geologists Bulletin, v. 58, p. 447–463.
Marshak, S., and G. Mitra, 1988, Basic Methods of Structural Geology: Englewood Cliffs, NJ, Prentice-Hall,
446 p.
Martin, R. G., 1980, Distribution of salt structures in the Gulf of Mexico: Map and descriptive text: U.S. Geo-
logical Survey Map MF-1213.
May, R. S., K. D. Ehman, G. G. Gray and J. C. Crowell, 1993, A new angle on the tectonic evolution of the
Ridge Basin, a “strike-slip” basin in Southern California: Geological Society of America Bulletin, v. 105,
p. 1357–1372.
Mayuga, M. N., 1970, Geology and development of California’s giant – Wilmington Oil Field, in Halbouty,
M. T., ed., Geology of Giant Petroleum Fields: American Association of Petroleum Geologists Memoir 14,
p. 158–184.
McClay, K., and M. Bonora, 2001, Analog models of restraining stepovers in strike-slip fault systems: Ameri-
can Association of Petroleum Geologists Bulletin, v. 85, p. 233–260.
Medwedeff, D. A., 1988, Structural analysis and tectonic significance of Late-Tertiary and Quaternary com-
pressive-growth folding, San Joaquin Valley, California [Ph.D. thesis]: Princeton University, 184 p.
______, 1989, Growth fault-bend folding at southeast Lost Hills, San Joaquin Valley, California: American As-
sociation of Petroleum Geologists Bulletin, v. 73, p. 54–67.
Medwedeff, D. A., and J. Suppe, 1986, Growth-fault bend folding – precise deformation kinematics, timing
and rates of folding and faulting from syntectonic sediments: Geological Society of America, Abstracts with
Programs, v. 18, p. 692.
Merret R., and R. W. Almendinger, 1991, Estimates of strain due to brittle faulting: Sample of fault popula-
tions: Journal of Structural Geology, v. 13, p. 735–738.
Merritt, J. W., 1946, Geotechniques of oil exploration: Oil Weekly, v. 121, no. 5, p. 17–26.
Mertie, J. B., Jr., 1947, Calculation of thickness in parallel folds: Geological Society of America Bulletin, v. 58,
p. 779–802.
______, 1947, Delineation of parallel folds and measurement of stratigraphic dimensions: Geological Society
of America Bulletin, v. 58, p. 779–802.
______, 1948, Application of Brianchon’s theorem to construction of geologic profiles: Geological Society of
America Bulletin, v. 59, p. 767–786.
Mertosono, S., 1975, Geology of Pungut and Tandun Oil Fields, Central Sumatra: Indonesian Petroleum Asso-
ciation (June issue), p. 165–179.
Meyer, H. J., and H. W. McGee, 1985, Oil and gas fields accompanied by geothermal anomalies in Rocky
Mountain region: American Association of Petroleum Geologists Bulletin, v. 69, p. 933–945.
Mitra, S., 1986, Duplex structures and imbricate thrust systems: Geometry, structural position, and hydrocar-
bon potential: American Association of Petroleum Geologists Bulletin, v. 70, p. 1087–1112.
______, 1988, Three-dimensional geometry and kinematic evolution of the Pine Mountain thrust system,
southern Appalachians: Geological Society of America Bulletin, v. 100, p. 72–95.
______, 1993, Geometry and kinematic evolution of inversion structures: American Association of Petroleum
Geologists Bulletin, v. 77, p. 1159–1191.
Mitra, S., and J. Namson, 1989, Equal-area balancing: American Journal of Science, v. 289, p. 563–599.
Moody, J. D., 1973, Petroleum exploration aspects of wrench-fault tectonics: American Association of Petro-
leum Geologists Bulletin, v. 57, p. 449–476.
Montgomery, S. L., 1997a, Raster logs may be basis for a geologic workstation: Oil and Gas Journal, April 7
issue, p. 84–88.
Montgomery, S. L., 1997b, A thing of beauty: Oil and Gas Investor, August issue, p. 49–51.
References 803
Morley, C. K., 1999, Patterns of displacement along large normal faults: Implications for basin evolution and
fault propagation, based on examples from East Africa: American Association of Petroleum Geologists Bul-
letin, v. 83, p. 613–634.
Morley, C. K., R. A. Nelson, T. L. Patton, and S. G. Munn, 1990, Transfer zones in the East African Rift Sys-
tem and their relevance to hydrocarbon exploration in rifts: American Association of Petroleum Geologists
Bulletin, v. 74, p. 1234–1253.
Mosar, J., and J. Suppe, 1988, Fault propagation folds: models and examples from the Pre-Alps and the Jura:
6ieme Reunion du Groupe Tectonique Suisse, Neuchatel, 8/9.
Mount, V. S. and J. Suppe, 1987, State of stress near the San Andreas fault: Implications for wrench tectonics:
Geology, v. 15, p. 1143–1146.
______, 1992, Present-day stress orientations to active strike-slip faults: California and Sumatra: Journal of
Geophysical Research, v. 97, p. 11,995–12,013.
Mount, V. S., J. Suppe, and S. C. Hook, 1990, A forward modeling strategy for balancing cross sections: Amer-
ican Association of Petroleum Geologists Bulletin, v. 74, p. 521–531.
Murray, G. E., 1968, Salt structures of Gulf of Mexico basin – a review, in Braunstein, J., and G. D. O’Brien,
eds., Diapirism and Diapirs: American Association of Petroleum Geologists Memoir 8, p. 99–121.
Namson, J., 1981, Structure of the western Foothills Belt, Miaoli-Hsinchu area, Taiwan, (1) Southern Part: Pe-
troleum Geology of Taiwan, no. 18, p. 31–51.
Narr, W., and J. Suppe, 1994, Kinematics of basement-involved compressive structures: American Journal of
Science, v. 294, p. 802–860.
Nelson, P. H. H., 1980, Role of reflection seismic in development of Nembe Creek Field, Nigeria, in Halbouty,
M. T., ed., Giant Oil and Gas Fields of the Decade 1968–1978: American Association of Petroleum Geolo-
gists Memoir 30, p. 565–576.
New Orleans Geological Society, 1965, Oil & Gas Fields of Southeast Louisiana, v. 1: New Orleans, LA, New
Orleans Geological Society.
______, 1967, Oil & Gas Fields of Southeast Louisiana, v. 2: New Orleans, LA, as above.
______, 1983, Oil & Gas Fields of Southeast Louisiana, v. 3: New Orleans, LA, as above.
______, 1988, Offshore Louisiana Oil & Gas Fields, v. 2: New Orleans, LA, as above.
Nilsen, T. H., and R. J. McLaughlin, 1985, Comparison of tectonic framework and depositional patterns of the
Hornelen strike-slip basins of Norway and the Ridge and Little Sulphur Creek strike-slip basins of Califor-
nia, in Biddle, K. T., and N. Christie-Blick., eds., Strike-slip Deformation, Basin Formation and Sedimenta-
tion: Society of Economic Paleontologists and Mineralogists, Special Publication No. 37, p. 79–103.
Nunns, A. G., 1991, Structural restoration of seismic and geological sections in extensional regimes: American
Association of Petroleum Geologists Bulletin, v. 75, p. 278–297.
O’Brien, C., 1988, Pragmatic migration: A method for interpreting a grid of 2D migrated seismic data: The
Leading Edge, v. 7, no. 2, p. 24–29.
Obert, L. and W. I. Duval, 1967, Rock Mechanics and the Design of Structures in rock: New York, John Wiley
& Sons, Inc., 650 p.
Ocamb, R. D., 1961, Growth faults of south Louisiana: Transactions Gulf Coast Association of Geological So-
cieties, v. 11, p. 139–175.
Oppenheimer, D. H., 1990, Aftershock slip behavior of the 1989 Loma Prieta earthquake: Geophysical Re-
search Letters, v. 17, p. 1199–1202.
Pacht, J. A., B. Bowen, B. L. Shaffer, and W. R. Pottorf, 1992, Systems tracts, seismic facies, and attribute
analysis within a sequence stratigraphic framework — example from offshore Louisiana Gulf Coast, in
Rhodes, E. G., and T. F. Moslow, eds., Marine Clastic Reservoirs: New York, Springer-Verlag, p. 21–38.
Page, D., 1859, Handbook of Geological Terms and Geology: Edinburgh, London, William Blackwood &
Sons, 415 p.
Payton, C. E., ed., 1977, Seismic stratigraphy — applications to hydrocarbon exploration: American Associa-
tion of Petroleum Geologists Memoir 26, 516 p.
Peacock, D. C. P., and D. J. Sanderson, 1991, Displacements, segment linkage, and relay ramps in normal fault
zones: Journal of Structural Geology, v. 13, p. 721–733.
Pelto, C. R., 1954, Mapping of multicomponent systems: Journal of Geology, v. 62, p. 501–511.
Pennebaker, Paul E., 1972, Vertical net sandstone determination for isopach mapping of hydrocarbon reser-
voirs: American Association of Petroleum Geologists Bulletin, v. 56, p. 1520–1529.
Perrier, R., and J. Quiblier, 1974, Thickness changes in sedimentary layers during compaction history: Methods
for quantitative evaluation: American Association of Petroleum Geologists Bulletin, v. 58, p. 507–520.
804 References
Perry, W. J., D. H. Roeder, and D. R. Lageson, North American thrust-faulted terraines: American Association
of Petroleum Geologists Bulletin, Reprint Series No. 27, 466 p.
Peters, W. C., 1987, Exploration and Mining Geology, fourth ed.: Golden, CO, Colorado School of Mines, 696 p.
Pettijohn, F. J., P. E. Potter, and R. Siever, 1972, Sand and Sandstone: New York, Springer-Verlag, 618 p.
______, 1987, Sand and Sandstone, second ed.: New York, Springer-Verlag, 553 p.
Pollard, D. D., and P. Segall, 1987, Theoretical displacements and stresses near fractures in rock, with applica-
tions to faults, joints, veins, dikes and solution surfaces, in Atkinson, B. K., ed., Fracture Mechanics of Rock:
New York, Academic Press, p. 277–350.
Posamentier, H. W., M. T. Jervey, and P. R. Vail, 1988, Eustatic controls on clastic deposition Ι, in Wilgus,
C. K., B. J. Hastings, H. Posamentier, J. C. Van Wagoner, C. A. Ross, and C. G. St. C. Kendall, eds., Sea
Level Changes: An Integrated Approach: Society of Economic Paleontologists and Mineralogists, Special
Publication 42, p. 109–124.
Posamentier, H. W., and P. R. Vail, 1988, Eustatic controls on clastic deposition ΙΙ, in Wilgus, C. K., B. J. Hast-
ings, H. Posamentier, J. C. Wagoner, C. A. Ross, and C. G. St. C. Kendall, eds., Sea Level Changes: An
Integrated Approach: Society of Economic Paleontologists and Mineralogists, Special Publication 42,
p. 125–154.
Potter, P. E., 1963, Late Paleozoic sandstones of the Illinois Basin: Illinois State Geological Survey, Report of
Investigations 217, 92 p.
Powers, R. B., ed., 1982, Geologic Studies of the Cordilleran Thrust Belt, 3 volumes: Denver, CO, Rocky
Mountain Association of Geologists.
Price, R. C., 1986, The southeastern Cordillera: Thrust faulting, tectonic wedging and delamination of the
lithosphere: Journal of Structural Geology, v. 8, p. 239–254.
Ragan, D. M., 1985, Structural Geology: An Introduction to Geometrical Techniques: New York, John Wiley &
Sons, 393 p.
Ramsay, J. G., 1967, Folding and fracturing of rocks: New York, McGraw-Hill Inc., 568 p.
Rees, F. B., 1972, Methods of mapping and illustrating stratigraphic traps: in King, R. E., Stratigraphic Oil and
Gas Fields – Classification, Exploration Methods, and Case Histories: American Association of Petroleum
Geologists Memoir 16, p. 168–221.
Reid, H. F., 1909, Geometry of faults: Geological Society of America Bulletin, v. 20, p. 171–196.
Reid, H. F., W. M. Davis, A. C. Lawson, F. L. Ransome, and Committee, 1913, Report of the committee on the
nomenclature of faults: Geological Society of America Bulletin, v. 24, p. 163–186.
Reiter, W. A., 1947, Contouring fault planes: World Oil, v. 126, no. 7, p. 34–35.
Remmelts, G., 1995, Fault-related salt tectonics in the southern North Sea, The Netherlands: in Jackson, M. P. A.,
D. G. Roberts and S. Snelson, eds., Salt tectonics: A global perspective: American Association of Petroleum
Geologists Memoir 65, p. 261–272.
Rettger, R. E., 1929, On specifying the type of structural contouring: American Association of Petroleum Geol-
ogists Bulletin, v. 13, p. 1559–1560.
Rich, J. L., 1932, Simple graphical method for determining true dip from two components and for constructing
contoured maps from dip observations: American Association of Petroleum Geologists Bulletin, v. 16,
p. 92–94.
______, 1934, Mechanics of low-angle overthrust faulting as illustrated by Cumberland thrust block, Virginia,
Kentucky, and Tennessee: American Association of Petroleum Geologists Bulletin, v. 18, p. 1584–1596.
______, 1935, Graphical method for eliminating regional dip: American Association of Petroleum Geologists
Bulletin, v. 19, p. 1538–1540.
______, 1951, Three critical environments of deposition and criteria for recognition of rocks deposited in each
of them: Geological Society of America Bulletin, v. 62, p. 1–20.
Rider, M. H., 1978, Growth faults in Carboniferous of Western Ireland: American Association of Petroleum
Geologists Bulletin, v. 62, p. 2191–2213.
Robinson, J. P., 1982, Petroleum exploration in southeastern Arizona: anatomy of an overthrust play, in Pow-
ers, R. B., ed., Geologic Studies of the Cordilleran Thrust Belt: Rocky Mountain Association of Geologists,
p. 665–674.
Roeder, D. H., 1973, Subduction and orogeny: Journal of Geophysical Research, v. 78, p. 5005–5024.
______, 1983, Hydrocarbons and geodynamics of folding belts: Rocky Mountain Association of Geologists,
Continuing Education Short Course Notes.
Roux, W. F., Jr., 1978, The development of growth fault structures: American Association of Petroleum Geolo-
gists Bulletin, Structural Geology School Notes.
Rowan, M. G., and R. Kligfield, 1989, Cross-section restoration and balancing as aid to seismic interpretation
in extensional terrains: American Association of Petroleum Geologists Bulletin, v. 73, p. 955–966.
References 805
Rowan, M. G., and R. Linares, 2000, Fold evolution matrices and axial surface analysis of fault bend folds: Ap-
plication to the Medina Anticline, eastern Cordillera, Colombia: American Association of Petroleum Geolo-
gists Bulletin, v. 84, p. 741–764.
Royse, F., Jr., M. A. Warner, and D. C. Reese, 1975, Thrust belt structural geometry and related stratigraphic
problems, Wyoming-ldaho-northern Utah, in Bolyard, D. W., ed., Symposium on Deep Drilling Frontiers in
the Central Rocky Mountains: Rocky Mountain Association of Geologists, p. 41–54.
Sachnik, F. L., and R. D. More, 1983, Southern Appalachian faulting and folding, in Bally, A. W., ed., 1983,
Seismic expression of structural styles: American Association of Petroleum Geologists, Studies in Geology
No. 15, v. 3, p. 3.4.1–79 to 3.4.1–82.
Sanchez, R., J. Y. Chatellier, R. de Sifontes, N. Parra, and P. Muñoz, 1997, Multiple Bischke Plot Analysis, a
powerful method to distinguish between tectonic or sedimentary complexity and miscorrelations: Memoras
del Primero Congreso Latinoamericano de Sedimentologia, Sociedad Venezolana de Geologos, Tomo II,
p. 257–264.
Sanford, A. R., 1959, Analytical and experimental study of simple geologic structures: Geological Society of
America Bulletin, v. 70, p. 19–52.
Santiago-Acevedo, J., 1980, Giant fields of the Southern Zone, Mexico: in Halbouty, M. T., ed., Giant Oil and
Gas Fields of the decade 1968–1978: American Association of Petroleum Geologists Memoir 30, p. 339–385.
Schlumberger, 1987, Log Interpretation Principles/Applications: Schlumberger Educational Services.
Schmid, C. F., and E. H. Maccannell, 1955, Basic problems, techniques and theory of isopleth mapping: Jour-
nal of The American Statistical Association, v. 50, no. 269, p. 220–239.
Scholle, P. A., and D. Spearing, 1982, Sandstone Depositional Environments: American Association of Petro-
leum Geologists Memoir 31, 410 p.
Sclater, J. G., and P. A. F. Christie, 1980, Continental stretching: An explanation of the post-mid-Cretaceous
subsidence of the central North Sea Basin: Journal of Geophysical Research, v. 85, p. 3711–3739.
Schwartz, S. Y., D. L. Orange, and R. S. Anderson, 1994, Complex fault interactions in a restraining bend on
the San Andreas Fault, Southern Santa Cruz Mountains, California, in Simpson, R. W., ed., The Loma Prieta,
California, Earthquake of October 17, 1989: Tectonic Processes and Models: United States Geological Sur-
vey Professional Paper 1550-F, p. F49–54.
Sebring, L., Jr., 1958, Chief tool of the petroleum exploration geologist: The subsurface structural map: Ameri-
can Association of Petroleum Geologists Bulletin, v. 42, p. 561–587.
______, 1958, Subsurface map: Underground guide for oil men: Oil and Gas Journal, v. 56, no. 27, p. 186–189.
Sengbush, R. L., 1986, Petroleum Exploration: A Quantitative Introduction: Boston, International Human Re-
sources Development Corporation Press, 233 p.
Seni, S. J., and M. P. A. Jackson, 1983, Evolution of salt structures, east Texas diapir province, part 1: Sedi-
mentary record of halokinesis: American Association of Petroleum Geologists Bulletin, v. 67, p. 1219–1244.
Serra, O., 1985, Sedimentary Environments From Wireline Logs: Schlumberger, 211 p.
Setchell, J., 1958, A nomogram for determining true stratum thickness: Shell Trinidad EP 28884, Abstract
in PA Bulletin, No. 127/128: N. V. DeBataafache Petroleum Maatschappij, The Hague, Production Depart-
ment, p. 8.
Shaw, J. H., R. E. Bischke, and J. Suppe, 1994, Relations between folding and faulting in the Loma Prieta epi-
central zone: Strike-slip fault bend folding, in Simpson, R. W., ed., The Loma Prieta, California, Earthquake
of October 17, 1989: Tectonic Processes and Models: United States Geological Survey Professional Paper
1550-F, p. F3–21.
Shaw, J. H., S. C. Hook, and E. P. Sitohang, 1997, Extensional fault-bend folding and synrift deposition: An
example from the Central Sumatra Basin, Indonesia: American Association of Petroleum Geologists Bulletin,
v. 81, p. 367–379.
Shaw, J. H. and J. Suppe, 1994, Active faulting and growth folding in the eastern Santa Barbara Channel, Cali-
fornia: Geological Society of America Bulletin, v. 106, p. 607–626.
______, 1996, Earthquake hazards of active blind-thrust faults under the central Los Angeles basin, California:
Journal of Geophysical Research, v. 101, p. 8623–8642.
Shelton, J. W., 1984, Listric normal faults: An illustrated summary: American Association of Petroleum Geolo-
gists Bulletin, v. 68, p. 801–815.
Sheriff, R. E., 1973, Encyclopedic Dictionary of Exploration Geophysics: Tulsa, OK, Society of Exploration
Geophysicists, 266 p.
______, 1980, Seismic Stratigraphy: Boston, International Human Resources Development Corporation Press.
______, 1989, Geophysical Methods: Englewood Cliffs, NJ, Prentice-Hall, 605 p.
Silver, A., 1982, Techniques of Using Geologic Data: Oklahoma City, OK, IED Exploration.
806 References
Singer, J. M., 1992, An example of log interpretation in horizontal wells: The Log Analyst, March-April issue,
p. 85–95.
Situmorang, B., Siswoyo, E. Thajib, and F. Paltrinieri, 1976, Wrench fault tectonics and aspects of hydrocarbon
accumulation in Java: Indonesian Petroleum Association, June issue, p. 53–67.
Skilbeck, C. G., and Lennox, eds., 1984, Seismic Atlas of Australia and New Zealand Sedimentary Basins:
Sydney, Earth Resources Foundation, 301 p.
Smith, D. A., 1966, Theoretical consideration of sealing and non-sealing faults: American Association of Petro-
leum Geologists Bulletin, v. 50, p. 363–374.
______, 1980, Sealing and nonsealing faults in Louisiana Gulf Coast Salt Basin: American Association of Pe-
troleum Geologists Bulletin, v. 64, p. 145–172.
Smith, D. A., and F. A. E. Reeve, 1970, Salt piercement in shallow Gulf Coast salt structures: American Associ-
ation of Petroleum Geologists Bulletin, v. 54, p. 1271–1289.
Smoluchowski, M., 1909, Folding of the earth’s surface: Information of mountain chains: Academy of Science
Cracovie Bulletin, v. 6, p. 3–20.
Sneider, R. M., F. H. Richardson, D. D. Paytner, R. E. Eddy, and I. A. Wyant, 1977, Predicting reservoir rock
geometry and continuity in Pennsylvanian reservoirs, Elk City Field, Oklahoma: Journal of Petroleum Tech-
nology, v. 29, p. 851–866.
Spencer, E. W., 1977, Introduction to the Structure of the Earth, second ed.: New York, McGraw-Hill, 640 p.
Stephenson, E. A., and D. D. Haines, 1946, Preparation of contour maps: Oil and Gas Journal, v. 45, no. 15,
p. 115.
______, 1946, Use of contour maps: Oil and Gas Journal, v. 45, no. 16, p. 131.
Stewart, W. A., 1950, Unconformities: in Leroy, L. W. ed., Subsurface Geologic Methods – A Symposium, sec-
ond ed.: Golden, CO, Colorado School of Mines.
Strahler, A. N., 1948, Geomorphology and structure of the West Kaibab fault zone and Kaibab Plateau, Ari-
zona: Geological Society of America Bulletin, v. 59, p. 513–540.
Straley, W. H., III, 1932, Some notes on the nomenclature of faults: Studies for Students: Chicago, University
of Chicago, p. 756–763.
Stockwell, C. H., 1947, The use of plunge in the construction of cross-sections of folds: Geological Association
of Canada, v. 3, p. 97–121.
Stone, D. S., Wrench faulting and Rocky Mountain tectonics: The Mountain Geologist, v. 6, no. 2, p. 67–79.
Stude, G. R., 1978, Depositional environments of Gulf of Mexico South Timbalier Block 54 salt dome and salt
dome growth models: Transactions Gulf Coast Association of Geological Societies, v. 28, p. 627–646.
Suppe, J., 1980, Imbricate structure of western Foothills Belts, south-central Taiwan: Petroleum Geology of
Taiwan, no. 17, p. 1–16.
______, 1983, Geometry and kinematics of fault-bend folding: American Journal of Science, v. 283, p. 684–721.
______, 1985, Principles of Structural Geology: Englewood Cliffs, NJ, Prentice-Hall, 537 p.
______, 1988, Short course on cross-section balancing in petroleum structural geology: Pacific Section, Ameri-
can Association of Petroleum Geologists.
Suppe, J., and Y. L. Chang, 1983, Kink method applied to structural interpretation of seismic sections, western
Taiwan: Petroleum Geology of Taiwan, no. 19, p. 29–47.
Suppe, J. G., G. T. Chou, and S. C. Hook, 1992, Rates of folding and faulting from growth strata, in McClay,
K. R., ed., Thrust Tectonics: New York, Chapman and Hall, p. 105–121.
Suppe, J., and D. A. Medwedeff, 1984, Fault-propagation folding: Abstracts with Programs, Geological Society
of America Bulletin, v. 16, p. 670.
Suppe, J., and D. A. Medwedeff, 1990, Geometry and kinematics of fault-propagation folding: Eclogae Geo-
logicae Helvetiae, v. 83, p. 409–454.
Suppe, J., and J. Namson, 1979, Fault-bend origin of frontal folds in the western Taiwan fold-and-thrust belt:
Petroleum Geology of Taiwan, no. 6, p. 1–18.
Sutter, H. H., 1947, Exaggeration of vertical scale of geologic sections: American Association of Petroleum
Geologists Bulletin, v. 31, p. 318–339.
Sylvester, A. G., 1988, Strike-slip faults: Geological Society of America Bulletin, v. 100, p. 1666–1703.
Tanner, J. H., III, 1967, Wrench fault movements along Washita Valley Fault, Arbuckle Mountain area, Okla-
homa: American Association of Petroleum Geologists Bulletin, v. 51, p. 126–141.
Taylor, J. C., 1973, Recent developments at Signal Hill, Long Beach oil field: in Guidebook for the Society of
Economic Paleontologists and Mineralogists, Annual Meeting with the American Association of Petroleum
Geologists and the Society of Economic Geophysicists, Field Trip 1, p. 16–25.
References 807
Tearpock, D. J., 1997, Project Management and Synergistic Team Development: Short Course Training Man-
ual, Subsurface Consultants & Associates, LLC.
______, 1998, The ten commandments of quality interpretation and mapping: GeoLOGIC, a quarterly techni-
cal newsletter of Subsurface Consultants & Associates, Fall issue.
______, 1999, Rebuttal to Expert Witness Reports of John C. Davis, Michael J. Economides, Jack C. Gold-
stein, and Harry F. Manbeck, Jr.: Union Pacific Resources Company, Plaintiff v. Chesapeake Energy Corpo-
ration and Chesapeake Operating, Inc., Defendants. C.A. No. 4–96CV-726-Y, Deposition Exhibit 10,
Tearpock.
______, 2000, Outsourcing: A key to success in the Twenty-first Century: Energy Houston, v. 2.
Tearpock, D. J., and R. E. Bischke, 1980, The structural analysis of the Wissahickon Schist near Philadelphia,
Pennsylvania: Geological Society of America Bulletin, v. 91, p. 644–647.
Tearpock, D. J., and R. E. Bischke, 1990, Mapping throw in place of vertical separation: a costly subsurface
mapping misconception: Oil and Gas Journal, July 16, v. 88, no. 29, p. 74–78.
Tearpock, D. J., and R. E. Bischke, 1991, Applied Subsurface Geological Mapping, first ed.: Upper Saddle
River, NJ, Prentice-Hall, 648 p.
Tearpock, D. J., R. E. Bischke, and J. L. Brewton, 1994, Quick Look Techniques for Prospect Evaluations:
Lafayette, LA, Subsurface Consultants & Associates, 286 p.
Tearpock, D. J., and J. C. Brenneke, 2001a, Shared earth modeling: A true multidisciplinary approach to E&P:
Energy Houston, v. 4, no. 1, p. 40–45.
______, 2001b, Multidisciplinary teams, integrated software for shared-earth modeling key E&P success: Oil
and Gas Journal, December 10, p. 84–88.
Tearpock, D. J., and J. Harris, 1987, Subsurface Geological Mapping Techniques – A Training Manual: Ten-
neco Oil Co.
Tearpock, D. J., and J. Harris, 1990, Isopach maps and their application in subsurface mapping: Lafayette Geo-
logical Society, Continuing Education Short Course Notes.
Tearpock, D. J., and J. Harris, 1990, Applied Subsurface Mapping Techniques: Gulf Coast Association of Geo-
logical Societies Short Course, Annual Convention, Lafayette, LA.
Tearpock, D. J., and J. Harris, 1990, Quantitative Mapping Techniques: Houston Geological Society, Continu-
ing Education Short Course Notes.
Tearpock, D. J., and H. Pousson, 1990, A three-dimensional correction factor equation for directionally drilled
wells: Transactions Gulf Coast Association of Geological Societies, v. 40.
Thomas, W. A., 1968, Contemporaneous normal faults on flanks of Birmingham anticlinorium, central Al-
abama: American Association of Petroleum Geologists Bulletin, v. 52, p. 2123–2136.
Thorogood, J. L., 1986, Well surveying: past progress, current status and future needs: World Oil, January
issue, p. 87–91.
______, 1990, Instrument performance models and their application to directional surveying operations: Soci-
ety of Petroleum Engineers, Paper 18051.
Thorsen, C. E., 1963, Age of growth faulting in southeast Louisiana: Transactions Gulf Coast Association of
Geological Societies, v. 13, p. 103–110.
Tobias, S., 1990, Expansion profiles and sequence stratigraphy: A new way to identify systems tracts, sequence
boundaries and eustatic histories: in Armentrout, J. M., and B. F. Perkins, Sequence Stratigraphy as an Explo-
ration Tool – Concepts and Practices in the Gulf Coast; Eleventh Annual Research Conference, Gulf Coast
Section, Society of Economic Paleontologists and Mineralogists: Austin, TX, Earth Enterprises, p. 351–361.
Todd, R. G., and R. M. Mitchum, 1977, Seismic stratigraphy and global changes of sea level, Part 8: Identifica-
tion of Upper Triassic, Jurassic, and Lower Cretaceous seismic sequences in Gulf of Mexico and offshore
West Africa, in Payton, C. E., ed., Seismic Stratigraphy – Applications to Hydrocarbon Exploration: Ameri-
can Association of Petroleum Geologists Memoir 26, p. 145–163.
Travis, Russell B., 1978, Graphic determination of stratigraphic and vertical thicknesses in deviated wells:
American Association of Petroleum Geologists Bulletin, v. 63, p. 845–866.
Trusheim, F., 1960, Mechanism of salt migration in northern Germany: American Association of Petroleum
Geologists Bulletin, v. 44, p. 1519–1541.
Tucker, P. M., 1982, Pitfalls Revisited: Tulsa, OK, Society of Exploration Geophysicists, 19 p.
______, 1988, Seismic contouring: A unique skill: Society of Exploration Geophysicists, v. 53, no. 6, p. 741–749.
Tucker, P. M., and H. J. Yarston, 1973, Pitfalls in seismic interpretation: Society of Exploration Geophysicists,
Monograph No. 2, 50 p.
Turk, L. B., 1950, Significance and use of lap-out maps in prospecting for oil and gas (Abs.): American Associ-
ation of Petroleum Geologists Bulletin, v. 34, p. 625.
808 References
Usdansky, S. I., and R. H. Groshong, Jr., 1984, Analytical extrapolation of cross sections of vertical drape folds
by digital computer: Geological Society of America, Abstract with Programs, Part A, v. 16, p. 258.
______, 1984, Comparison of analytical models for dip-domain and fault-bend folding: Geological Society of
America, Abstract with Programs, Part B, v. 16, p. 680.
Vail, P. R., and W. Wornardt, Jr., 1991, An integrated approach to exploration and development in the 90’s: Well
log-seismic sequence stratigraphy analysis: Transactions Gulf Coast Association of Geological Societies
v. 41, p. 630–650.
Vanderhill, J. B., 1991, Depositional setting and reservoir characteristics of Lower Queen (Permian Guadalu-
pian) sandstones, Keystone (Colby) Field, Winkler County, Texas, in Meader-Roberts S., M. P. Candelaria,
and G. E. Moore, eds., Sequence stratigraphy, facies, and reservoir geometries of the San Andeas, Grayburg
and Queen Formations, Guadalupe Mountains, New Mexico: Society of Economic Paleontologists and Min-
eralogists, Special Publication 91–32, p. 119–130.
Van Wagoner, J. C., R. M. Mitchum, K. M. Campion, and V. D. Rahmanian, 1990, Siliciclastic sequence
stratigraphy in well logs, cores, and outcrops: concepts for high-resolution correlation of time and facies:
American Association of Petroleum Geologists, Methods in Exploration 7, 55 p.
Van Wagoner, J. C., H. W. Posamentier, R. M. Mitchum, P. R. Vail, J. F. Sarg, T. S., Loutit, and J. Hardenbol,
1988, An overview of the fundamentals of sequence stratigraphy and key definitions: in Wilgus, C. K., B. S.
Hastings, C. G. St. C. Kendall, H. W. Posamentier, C. A. Ross, and J. C. Van Wagoner, eds., Sea Level
Changes: An Integrated Approach: Society of Economic Paleontologists and Mineralogists, Special Publica-
tion No. 42, p. 39–45.
Vogler, H. A., and B. A. Robison, 1987, Exploration for deep geopressure gas: Corsair Trend, offshore Texas:
American Association of Petroleum Geologists Bulletin, v. 71, p. 777–787.
Wadsworth, A. H., Jr., 1953a, Percentage of thinning chart – new techniques in subsurface geology: American
Association of Petroleum Geologists Bulletin, v. 37, no. 1, p. 158–162.
______, 1953b, The percentage of thinning chart: Oil and Gas Journal, v. 51, no. 43, p. 72–73.
Wallace, R. E., 1968, Notes on stream channel offsets by the San Andreas Fault, Southern Coastal Ranges, Cal-
ifornia, in Dickinson, W. R. and Grantz, A., eds., Proceedings of Conference on Geologic Problems of San
Andreas Fault System: Palo Alto, CA, Stanford University Press, v. 11, p. 6–21.
______, 1973, Surface fracture patterns along San Andreas fault, in Kovach, R. L., and A. Nur, eds., Proceed-
ings of Conference on Geologic Problems of San Andreas Fault System: Palo Alto, CA, Stanford University
Press, v. 13, p. 248–258.
Weber, K. J., and E. Daukora, 1976, Petroleum geology of the Niger Delta: Proceedings of the Ninth World Pe-
troleum-Congress, v. 2, p. 209–221.
Weber, K. J., G. Mandl, W. F. Pilaar, F. Lehner, and R. G. Precious, 1978, The role of faults in hydrocarbon mi-
gration and trapping in Nigerian growth fault structures: 10th Annual Society of Petroleum Engineers Off-
shore Technological Conference, Preprint no. OTC-3356, p. 2643–2653.
Weiss, L. E., 1972, The Minor Structures of Deformed Rocks; A Photographic Atlas: New York, Springer-Ver-
lag, 431 p.
Wernicke, B., 1985, Uniform-sense normal simple shear of the continental lithosphere: Canadian Journal of
Earth Sciences, v. 22, p. 108–125.
Wernicke, B., P. L. Guth, and G. L. Axen, 1984, Tertiary extensional tectonics in the Sevier thrust belt of south-
ern Nevada, in Lintz, J., ed., Western Geological Excursions, v. 4: Guidebook for the Annual Meeting of the
Geological Society of America: Reno, NV, Mackay School of Mines, p. 473–510.
West, J., and H. Lewis, 1982, Structure and palinspastic reconstruction of the Absaroka Thrust, Anschutz
Ranch area, Utah and Wyoming: Rocky Mountain Association of Geologists, 1982 Annual Symposium,
p. 633–639.
Wharton, J. B., Jr., 1948, Isopachous maps of sand reservoirs: American Association of Petroleum Geologists
Bulletin, v. 32, p. 1331–1339.
Wheeler, R. L., 1980, Cross-strike structural discontinuities: Possible exploration tool for natural gas in Ap-
palachian Overthrust Belt: American Association of Petroleum Geologists Bulletin, v. 64, p. 2166–2178.
White, N. J., J. A. Jackson, and D. P. McKenzie, 1986, The relationship between the geometry of normal faults
and that of the sedimentary layers in their hanging walls: Journal of Structural Geology, v. 8, p. 897–909.
Wiggins, G. B., and D. J. Tearpock, 1985, Methods in oil and gas reserves estimation: Prospects, newly discov-
ered and developed properties: American Association of Petroleum Geologists, Annual Convention Short
Course.
Wilcox, R. E., T. P. Harding, and D. R. Seely, 1973, Basic wrench tectonics: American Association of Petro-
leum Geologists Bulletin, v. 57, p. 74–96.
References 809
Wilgus, C. K., H. Posamentier, C. A. Ross, and C. G. St. C. Kendall, eds. 1988, Sea level changes: An in-
tegrated approach: Society of Economic Paleontologists and Mineralogists Special Publication No. 42,
p. 39–45.
Wilson, C. W., and R. G. Stearns, 1958, Structure of the Cumberland Plateau, Tennessee: Geological Society of
America Bulletin, v. 69, p. 1283–1296.
Winker, C. D., and M. B. Edwards, 1983, Unstable progradational clastic shelf margins, in Stanley, D. J., and
Moore, G. T., eds., The Shelfbreak: Critical Interface on Continental Margins: Society of Economic Paleon-
tologists and Mineralogists Special Publication No. 33, p. 139–157.
Winker, C. D., R. A. Morton, T. E. Ewing, and D. D. Garcia, 1983, Depositional setting, structural style, and
sandstone distribution in three geopressured geothermal areas, Texas Gulf Coast: University of Texas, Bureau
of Economic Geology, Report of Investigations 134, 60 p.
Winn, R. D., Jr., M.G. Bishop, and P. S. Gardner, 1985, Shallow-water and sub-storm-base deposition of Lewis
Shale in Cretaceous Western Interior seaway, south-central Wyoming: American Association of Petroleum
Geologists Bulletin, v. 71, p. 859–881.
Withjack, M. O., Q. T. Islam, and P. R. La Pointe, 1995, Normal faults and their hanging wall deformation: An
experimental study: American Association of Petroleum Geologists Bulletin, v. 79, p. 1–18.
Wolff, C. J. M., and J. P. De Wardt, 1981, Borehole position uncertainty – analysis of measuring methods and
deviation of systematic error model: Journal of Petroleum Technology, v. 33, p. 2339–2350.
Wood, M. F., 1981, Depositional environments of the Apple Canyon Sandstone, Ridge Basin, central Trans-
verse Range, California [M. S. thesis]: Los Angeles, University of Southern California, 266 p.
Woodbury, H. O., I. B. Murray, Jr., P. J. Pickford, and W. H. Akers, 1973, Pliocene and Pleistocene depocen-
ters, outer continental shelf, Louisiana and Texas: American Association of Petroleum Geologists Bulletin,
v. 57, p. 2428–2439.
Woodbury, H. O., I. B. Murray, Jr., and R. E. Osbourne, 1980, Diapirs and their relation to hydrocarbon accu-
mulation, in Miall, A. D., ed., Facts and Principles of World Petroleum Occurrence: Canadian Society of Pe-
troleum Geologists Memoir 6, p. 119–142.
Woodward, N. B., 1987, Stratigraphic separation diagrams and thrust belt structural analysis: Thirty-Eighth
Field Conference, 1987 Wyoming Geological Association Guidebook, p. 69–77.
Woodward, N. B., S. E. Boyer, and J. Suppe, 1985, An Outline of Balanced Cross Sections: University of Ten-
nessee Department of Geological Sciences, Studies in Geology 11, second ed.: Knoxville, TN, University of
Tennessee, 170 p.
Worrall, D. M., and S. Snelson, 1989, Evolution of the Gulf of Mexico, with emphasis on Cenozoic growth
faulting and the role of salt, in Bally, A. W., and A. R. Palmer, The Geology of North America, v. A: Boulder,
CO, Geological Society of America, p. 97–137.
Wright, T. L., 1991, Structural geology and tectonic evolution of the Los Angeles basin, California, in Biddle,
K. T., Active Margin Basins: American Association of Petroleum Geologists Memoir 52, p. 35–134.
Wyoming Geological Association, 1987, The Thrust Belt Revisited, Thirty-Eighth Field Conference Guide-
book: Casper, WY, Wyoming Geological Association, 403 p.
Xiao, H., F. A. Dahlen, and J. Suppe, 1991, Mechanics of extensional wedges: Journal of Geophysical Re-
search, v. 96, p. 10,301–10,318.
Xiao, H., and J. Suppe, 1988, Origin of rollover: Geological Society of America, Annual Convention, Abstracts
with Programs, p. A109.
______, 1989, Role of compaction in the listric shape of growth normal faults: American Association of Petro-
leum Geologists Bulletin, v. 73, p. 777–786.
______, 1992, Origin of rollover: American Association of Petroleum Geologists Bulletin, v. 76, p. 509–529.
Xiaohan, L., 1983, Perturbations de contraintes liées aux structures cassantes dans les calcaires fins du Langue-
doc. Observations et simulations mathématiques [unpubl. 3 eme cycle thesis]: Montpellier, Université des
Sciences et Techniques du Languedoc, 130 p.
Yeats, R. S., and J. M. Beall, 1991, Stratigraphic controls of oil fields in the Los Angeles Basin, in Biddle,
K. T., ed., Active Margin Basins: American Association of Petroleum Geologists Memoir 52, p. 221–235.
Zheng, Z., J. Kemeny, and N. G. Cook, 1989, Analysis of borehole breakouts: Journal of Geophysical Re-
search, v. 94, p. 7171–7182.
Zoback , M. D., D. Moos, L. G. Mastin, and R. N. Anderson, 1985, Well bore breakouts and in situ stress: Jour-
nal of Geophysical Research, v. 90, p. 5523–5530.
Zoback, M. D., et al., 1987, New evidence on the state of stress of the San Andreas fault system: Science,
v. 238, p.1105–1111.
This page intentionally left blank
Index
811
812 Index
Boomerang fault, Yucca Mountains, computer-aided structural direct technique (triangulation), 23,
Nevada, 614, 617 modeling and balancing, 27–28
Bottom water reservoir, 729–30 515–20 indirect technique (gridding), 23–26
cross section, 729 cross section consistency, 527–29 surface modeling, 23
defined, 728 line length exercise, 513–15 Computer-based fault seal analysis,
net oil isochore map for, 729 picking thrust faults, 523–27 247–50
structure map, 729 pin lines, 513 application of 3D modeling and
three-dimensional model of, 729 retrodeformation, 520–23 visualization technique, 247–48
Bow and Arrow Rule, 515, 529, 533 volume accountability rule, fault-seal potential, 248–50
Bow Valley, Canadian Rocky Mountains, 510–11 Computer-based log correlation, 108–17
523, 526 Clay analog models of strike-slip advantages of, 117
Bows in fault surfaces, 612–614 faulting, 680–81 fault identification example, 114–17
Box structures, 574–77 Coherency data, 617, 619 on-screen log correlation, 109–14
Branch point, 536 Collapse folding, 428–29 Conformable geology and multi-surface
Brazos Ridge, Gulf of Mexico, 591–92, Colombia, Eastern Condillera, 298 stacking, 31–34
594, 603, 605, 605–7, 612–14, Colorado: Conjugate kink structures, 578
620, 620–34, 696 Rocky Mountains, 521 Conservation of fault size, See Additive
Corsair Fault, 603, 607–9, 611–14 Uinta Basin, 202–4 property of faults
Breaks in slope, MBPA, 692 Combined vertical separation, 283–89 Conservation of vertical separation, See
Breakthrough, faults, 560 Committee on the Nomenclature of Additive property of faults
Brittle theory of crystal deformation, Faults, 253–54 Consistency check, Structural
508–9 Compensating fault, 271–73 interpretation, 520, 722
Brunei, 314, 429, 610, 715, 721 Compensating fault pattern, 277–80, Constant contour interval, importance
Build rate, directionally drilled wells, 44 410–15 of, 14–15
Burgentine Lake Field, Texas, 599–600, Complex growth structures, 715–18 Constant dip domain method, 534, See
602 Composite type logs, 65 also Kink method
Busk method approximation, 533–34 Composite show logs, 65, 67 approximation
Busk method of segmented circular arcs, Compressional faulting: Constant-thickness fold, 539–40
529 intersecting compressional faults, Contemporaneous faults, 314; See also
292–94 Growth faults
ramp and flat thrust faults, 294–98 Contour interval, 14–15, 334–36
C reverse faults, 289, 291–92 Contour license, 340
single compressional faults, 292 Contour lines:
Cactus-Nispero Field, Mexico, 407 thrust faults, 291–92 defined, 8
California: Compressional growth structures, 715 spacing of, 11, 335
Little Sulphur Creek Basin, 679 Compressional inversion structures, 718 Contour maps, 8–9
Newport-Inglewood Trend, 652 Compressional plate boundaries and defined, 8
Rosecrans Oil Field, 440, 444 thrust faults, 291 and sound geologic principles, 11
Transverse Ranges, 652, 680–81 Compressional restraining bend, 645 Contourable data, and associated
Wilmington Field, Los Angeles Compressional structures, 199, 506–83 contour map, 9
Basin, 440, 445 cross section consistency, 527–29 Contouring, 8–42, 324–356
Canadian Rocky Mountains, 519, 523, cross section construction, 529–33 anticlinal structures, 340
526–27, 577 depth to detachment calculations, change or reversal in the direction of
Central Luzon Valley-Llocos Basins, 545–47 dip, 339
Philippines, 646 mechanical stratigraphy, 508–10 chosen reference, 334
Central Sumatra Basin, Indonesia, structural balancing, 506–8, 507–8 closed structural lows, 340
591–93, 614 benefits of, 507 computer-based contouring, 23–42
Chaotic zone, 281 classical balancing, 508, 510–27 contouring faulted surfaces on the
Checkshot survey, 143 nonclassical balancing, 508 computer, 34–37
Cherry Hill Fault, California, 652–54 ultimate goals of, 507 direct technique (triangulation),
Chevron folds, 578 Compressional tectonic mapping 23, 27–28
Chile, 652 techniques: indirect technique (gridding),
China, 595 reverse faults, 447–50 23–26
Chuhuangkeng Anticline, Taiwan, 576 thrust faults, 450–53 surface modeling, 23
Circle method, 372–78 Computer-aided structural modeling and conformable geology and multi-
Classical balancing, 508, 508–27, 548 balancing, 515–20 surface stacking, 31–34
techniques, 510–27 Computer-based contouring, 23–44 constructing contours in groups of
area accountability, 511 contouring faulted surfaces on the several lines, 337
bed length consistency, 511–12 computer, 34–37 contour compatibility, 339
Index 813
Diapiric salt structures, 197–99 nonmagnetic surveys, 50 correlation of vertical wells and
Diapiric salt tectonic mapping survey errors, 54 directional wells, 84–86
techniques: uncertainties, 52–54 estimating the missing section for
contouring the salt surface, 432–33 directional tools used for normal faults, 86–99
hydrocarbon traps, 431–32 measurements, 48–50 log correlation plan, 81–84
salt-fault intersection, 433 directional well plan, 47–48 by hand, 62
salt-sediment intersection, 433–35 directional well plots, 54–56 horizontal wells, 101–7
Dickinson method, 421 drop rate, 44 direct detection of bed boundaries,
Digital raster well log formats, 108 general terminology, 44 102
Dip analysis, 520 horizontal wells, 43, 44–46 modeling log response of bed
Dip domains, 597 extended reach wells, 44 boundaries and fluid contacts,
constant dip domain method, 534, long-radius horizontal wells, 44 102
See also Kink method short-radius horizontal wells, true stratigraphic depth (TSD)
approximation 44–46 method, 105–7
curved dip domain method, 534, See KOP (kick-off point), 44 true vertical depth (TVD) cross
also Busk method offshore wells, drilled from a single section, 102–4
approximation platform, 43 procedures/guidelines, 61–65
graphical dip domain technique for, ramp angle, 44 stratigraphic type log, 65–66
595–602 vertical point, 44 using the computer, 108–117
Dip rate increases, and faulting, 338 Disconformity, 393 vertical wells, 69–80
Dip reversal, 339 Discontinuities on ∆d/d plots, 690 fault determinations, 73–76
Dip cross section, 175 Distortion in cross sections, 201 faults versus variations in
Dip spectral analysis, 569–74 Documentation: stratigraphy, 73
Dipmeter recording within a stratigraphic log correlation, 129–33 log correlation plan, 69–71
unit and unconformities, 533 three-dimensional seismic data, 465 Electric log cross sections, 180–82
Dipping beds, 91–99 Domal structures, 340 En echelon faulted structures, 614–620
three-dimensional correction factor, contouring, 340 En echelon folded structures, 643
94–96 Downlap, 692–94 Equal-spaced contouring, 19–21, 342–
TVT calculation using, 98–99 Downthrown fault block, 71 343, 791
two-dimensional correction factor, Downthrown (hanging wall) restored Europe, 431
93–94 top, 123 Expansion fault, use of term, 683
TVT calculation using, 97–98 Downward-dying growth faults, Expansion Index Method, 683–87
Direct contouring technique 718–21 benefits to using, 686
(triangulation), 23, 27–28 Drain hole wells, 44–46 disadvantages to using, 687
Direction of well deviation, 51 Drift indicator tool, 50 Extended reach wells, 44
Directional surveys, and fault surface Drop rate, directionally drilled wells, 44 Extensional faulting:
maps, 322–29 Duplex structures, 565–68 bifurcating fault pattern, 273–77
Directional wells: combined vertical separation,
pitfalls of, 327–29 283–89
and repeated sections, 329 E compensating fault pattern, 277–80
Directionally drilled well log intersecting fault pattern, 281–83
correlation, 81–99 East Painter Reservoir Field, Wyoming, Extensional folding along strike-slip
correlation of vertical wells and direc- Nugget Sandstone, 199–200 faults, 668–79
tionally drilled wells, 84–86 East Texas Oil Field, 393 geometry of, 672–79
estimating the missing section for Eastern Condillera, Colombia, S.A., 298 Hungry Valley Fault, 674–79
normal faults, 86–99 Edge water reservoir: Ridge Basin geology, 669–72
dipping beds, 91–99 basic construction, 730–42 Extensional releasing bends, 645
horizontal beds, 86–91 cross section, 731 Extensional structures, 197
log correlation plan, 81–84 defined, 728 balancing/interpretation, 584–634
Directionally drilled wells, 43–59 net oil isochore map for, 731 compaction effects along growth
application of, 43–44 structure map, 731 normal faults, 620–34
build rate, 44 three-dimensional model of, 731 inverting fault dips to determine
common types of, 44–46 Elastic dislocation models, 641 sand/shale ratios or percent
controlled directional drilling, 43 Electric log correlation, 60–133 sand, 627–34
defined, 43 basics concepts of, 71–72 prospect example, 623–27
directional surveys composite type logs, 65 cross structures, 612–14
calculations, 50–52 composite show logs, 65, 67 downward dying growth faults,
free gyroscope surveys, 52 defined, 61 602–12
magnetic surveys, 50, 52 directionally drilled wells, 81–99 hanging wall (rollover) anticlines
Index 815
Coulomb collapse theory, 585–89 Fault maps, 251–331 fault-seal potential, 248–50
growth sedimentation, 589–95 directional surveys and fault surface fault surface sections constructed by
origin of, 584–95 maps, 322–29 hand, 243–47
keystone structures, 602–12 fault data determined from seismic Fault shape, and fold shape, 558
projecting large growth faults to information, 298–314 Fault slicing, 243
depth, 597 seismic and well log data Fault surface maps, 342, 597, 620
determining Coulomb collapse integration, 305–8 and directional surveys, 322–29
angles from rollover structures, seismic pitfalls, 308–14 and repeated sections, 329
601–2 fault data determined from well logs, compressional areas, 446
graphical dip domain technique 260–63 construction of, 263–71, 305–8
for, 595–602 fault displacement, definition of, fault-displacement mapping,
procedures for, 597–600 255–56 307–8
rollover geometry features, fault patterns, 271–98 techniques, 267–71
596–97 compressional faulting, 289–98 contouring guidelines, 266–67
strike-ramp pitfall, 614–20 extensional faulting, 271–89 kinks in, 636–37
synthetic and antithetic faults, fault surface maps: loop-tied, 620
602–12 construction of, 263–71, 305–8 Fault surfaces, quality-checking in map
three-dimensional effects, 612–14 construction techniques, 267–71 and seismic views, 479–80
Extensional tectonic mapping contouring guidelines, 266–67 Fault throw, defined, 255
techniques: fault terminology, 253–55 Fault traces, 345, 369–83, 497–99
bifurcating fault pattern, 415–19 growth faults, 314–22 equation to determine heave, 382–83
combined vertical separation, 426 estimating the vertical separation equation to determine radius of circle,
compensating fault pattern, 410–15 (missing section) for, 314–20 380–81
exceptions to, 408 heave, 255 new circle method, 378–80
growth faults, 427–30 mathematical relationship of throw to Rule of 45, 371–72
intersecting fault pattern, 419–26 vertical separation, 256–60 tangent method/circle method,
missing section, 255 372–78
repeated section, 255 Fault vertical separation, 34–35
F throw, 255 Fault wedges, 732, 766–69
vertical separation, 254, 329 conventional method for mapping,
Fatigue Wash Fault, 614 Fault overlap, 369 766–67
Yucca Mountains, Nevada, 614, accurate, drawing, 495–99 defined, 766
617 Fault patterns, 271–98 Mid-Trace Method (Mid-Point
Fault bend folds, 291, 445, 549–52 antithetic fault, 271 Method), 767–69
animated models of, 518 compensating fault, 271 Fault zone, 640
Fault boundaries, 497–99 compressional faulting, 289–98 Faulted surfaces:
Fault data: intersecting compressional faults, contouring, 342–56
determined from seismic, 298–313 292–94 across normal faults, 345–55
determined from well logs, 260–263 ramp and flat thrust faults, 294–98 across reverse faults, 355–56
Fault displacement definition on seismic single compressional faults, 292 checking structure maps (error
data, 298–307 extensional faulting, 271–89 analysis), 350–54
Fault-displacement mapping, 307–8 bifurcating fault pattern, 273–77 fault trace, 345
Fault gap mode, 237 combined vertical separation, legitimate contouring of throw,
Fault gaps, See also Fault traces 283–89 354–55
accurate, drawing, 495–99 compensating fault pattern, 277–80 mapping throw across a fault,
defined, 369, 383 intersecting fault pattern, 281–83 349–50
width of, 371 master fault, 271 projecting contours within the
Fault heave, defined, 383 normal faulting, defined, 271 fault gap, 347
Fault ID, using seismic data, 302 Fault polygons, 35, 497–99 contouring on the computer, 34–37
Fault integration, on a workstation, mislocated, common causes for, 501 limitations, 37
501–5 Fault propagation breakthrough, 563 procedure for, 36–37
Fault interpretation on seismic data, Fault propagation folds, 445, 549, Faults, additive property of, 387–90
465–80 558–74 Fence diagrams, 223–27
quality-checking fault surfaces in animated models of, 518 construction of, 224–27
map and seismic views, balancing, 560–64 Field production history, reservoirs,
479–80 Fault seal analysis, 240–50 781–82
reconnaissance, 469 computer-based, 247–50 Finished illustration cross section, 182,
strategies, 472–79 application of 3D modeling and 185–93
well control, integrating, 469–72 visualization technique, 247–48 Flap map, 453
816 Index
Fold-and-thrust belts, types of, 291 Growth anticline, 702–3 Gyroscopic survey tools, 50
Fold geometry, 551–52, 556–74, 639–40 Growth axial surface, 590–91, 590–93 Gyroscopic surveys, 52
Fold shape, and fault shape, 558 Growth compressional style, 702–3
Fold shape, predicting, 558 Growth-fault surface map construction,
Folded fold, 551 321–22 H
Footwall, 620 Growth faults, 71, 620–34
Forward dips, 570–71 compaction effects along, 620–34 Hachured lines, 15, 336
Fossil Basin, Utah/Wyoming, 298 estimating the vertical separation Half-graben structure, 592
Four-way closures, 592–94 (missing section) for: Hand contouring, 16–22
Fourth-dimensional time factor, 682 restored top method, 314–15 equal-spaced contouring, 19–20
Fracture porosity, 554 single well method, 315–20 interpretive contouring, 20–21
Framework horizons, 462–63 Expansion Index Method mechanical contouring, 16–18
selecting, 480–87 benefits to using, 686 parallel contouring, 18–19
Frazier Mountain Thrust, California, 669 disadvantages to using, 687 Handpicking seismic events, as infill
Frictional theory of crystal deformation, growth-fault surface map strategy, 492
508–9 construction, 321–22 Hanging wall anticlines, and growth
Full thickness area, reservoir, 732, 730. and hanging wall anticlines, 314 faults, 314
and rollovers, 314 Hanging wall geometry, and positioning
Growth interval, 695 wells, 660–61
G Growth sedimentary section, 592–94 Hanging wall (rollover) anticlines:
Growth strata, and piercing line example of a rollover structure,
Gabon, 431 evidence, 645–46 591–95
Garlock Fault, 646–47 Growth structures, 682–722 growth sedimentation, 589–95
Gas reservoirs, 780–81 analysis of, 685–86 origin of, 584–95
General map symbols, 789 Expansion Index Method, 683–87 Heave, 255, 256, 347, 383
Geologic cross sections, See Cross growth inversion structures, 683 equation to determine, 382–83
sections growth reverse faults, 683 subsurface petroleum-related
Geologic license, 15 growth strike-slip faults, 683 structure mapping, 347
Geophysical data: growth thrust faults, 683 High-angle thrust faults, 292
integration in subsurface mapping, Multiple Bischke Plot Analysis High-resolution marker correlations, 117
134–74 (MBPA)/∆d/d methods, 687–94 Hogsback ramp and flat thrust fault, 298
assumptions/limitations, 135–36 common extensional growth Hope Fault, New Zealand, 646, 648–49
Glenwood Syncline, California, 659–63 patterns, 690–92 Horizon integration, on a workstation,
Goguel’s Law of Volume Conservation, unconformity patterns, 692–96 501–5
671 vertical separation versus depth Horizon interpretation, 480–95
Golden Meadow Field, Lafourche (VS/d) method, 711–21 three-dimensional seismic data,
Parish, Louisiana, 286, examples of analysis of VS/d 480–95
725–27 plots, 713–21 framework horizons, selecting,
Good Hope Field, St. Charles Parish, technique, 712–13 480–87
Louisiana, 67 Guadalupe Mountains, New Mexico, infill strategies, 492–95
Graben blocks, 431 700 interpretation strategy, 488–92
Grand Isle Ash at Grand Isle Block, Gulf Coast, 431, 622 well data, tying, 487–88
435–39 Gulf of Mexico, 135, 171, 264, 304–5, Horizon mis-tie problem, 155
Graphic scale, 15, 336 314, 404, 408, 416, 429, 584, Horizontal displacements, 637–38, 645
Grayburg Formation, Guadalupe 589, 595, 599–600, 610, and strike-slip faults, 645
Mountains, Texas, 700–701 622–23, 698, 709–11, 757 Horizontal Slice Method, 775, 776–77
Great Glen Fault, Scotland, 646 Austin Chalk, 46 Horizontal well log correlation, 101–7
Green River Basin, Wyoming, 113, 236 bows in, 614 direct detection of bed boundaries,
Gridding, 23–26 Brazos Ridge, 591–92, 594, 605, 102
estimating values at grid nodes, 26 605–7, 612–14, 620, 696 modeling log response of bed
selecting neighbors, 24–26 East Cameron Block 270 Field, boundaries and fluid contacts,
natural neighbors, 26 429–30 102
nearest “n” neighbors, 24–26 growth faults, 714–15 true stratigraphic depth (TSD)
steps involved in, 24 Mississippi Canyon Block 194, 56 method, 105–7
Gridding a horizon, as infill strategy, Rob L Horizon, 709–11 true vertical depth (TVD) cross
494 South Pelto area, 109, 114 section, 102–4
Gridding and contouring, 499 Tertiary section in, 140 Horizontal wells, 43, 44–46, 101–7
Growth, 682–84 time—depth table from, 144 extended reach wells, 44
history, 683–85, 722 West Cameron Block 192, 216–17 long-radius horizontal wells, 44
and structural styles, 683 Gulf of Mexico Basin, 304 purpose of drilling, 46
Index 817
short-radius horizontal wells, 44–46 basis of, 725 Kink method approximation, 533,
and water coning, 46 contouring the hydrocarbon wedge, 534–45, 573, 582
Hornelen Basin, Norway, 679 742–52 applications, 539–45
Horse structure, 565–66 limited well control/evenly Kinks, 617–19
Hungry Valley Fault, Ridge Basin, distributed impermeable rock, in fault surface maps, 636–37
California, 669–79, 674–79 742–44 KOP (kick-off point), directionally
Huntington Beach Field Anticline, walking wells, 744–52 drilled wells, 44
Newport-Inglewood Trend, mapping the top of structure
701–2 versus the top of porosity,
Hydro-fracturing experiments, 642 763–66 L
Hydrocarbon reserve estimation, 779 net pay isochore map, 725
Hydrocarbon traps, 431–32 construction, 730 Lafourche Parish, Louisiana, Golden
Hydrocarbon-water contacts, 102 construction for a reservoir Meadow Field, Lafourche
containing oil and gas, Parish, 286
737–42 Lateral changes in depositional
I construction for a single-phase environment, 193
reservoir, 732–37 Lateral displacements on strike-slip
IEPS (Integrated Exploration and net sand isochore maps, 725 faults, 645–56
Production System), 755 purpose of, 723 piercing point/piercing line evidence,
Idealized fold shape, 554 reservoir volume determinations 645–46
Imbricate structures, 564–574 from, 775–79 defined, 645
Inactive axial surface, 589, 589–90 Horizontal Slice Method, 775, isopach/isochore information, 646
Index contour, 15, 336 776–77 local restoration, 648–56
Indigo Bayou area, Iberville Parish, selecting a method, 779 pre-growth strata, 645
Louisiana, 271 Vertical Slice Method, 775, regional restoration, 646–48
Indirect contouring technique (gridding), 777–78 releasing bends, 648, 649–52
23–26 vertical thickness restraining bends, 648, 652–56
Indonesia, 584, 595 determining, 752–59 surface features, 645
Central Sumatra Basin, 591–93, 614 and fluid contacts in deviated Lateral ramp, 527
Infill strategies, 492–95 wells, 759–63 Lateral velocity changes, managing,
autopicking, 494–95 impact of correction factors, 170–71
gridding a horizon, 494 757–59 Law of Sines, 258
handpicking, 492 Isochore piling, 774 Least Squares gridding algorithm, 270
interpolation, 494 Isometric projections, 228 Lewis Shale, Wyoming, 113–14
Inner limit of water (ILW), 732, 745 Isopach maps, 723–87 Lift-off folds, 551
Integrated structure map, 357–369 defined, 723–24 Lift-off structures, 574–77
Interactive fault picking/gapping tools, interval isopach maps, 782–87 Line length balancing exercise, 513–15
117 construction, using seismic data, Line of bifurcation, 275
Interference structures, 578–83 786–87 Line of intersection, 281
Interpolation, as infill strategy, 494 true stratigraphic thickness (TST) Line of termination, 278
Interpretation workflow, developing, 464 from well logs, 783–85 Listric growth normal faults, 314
Interpretive contouring, 20–21, 342, 790 purpose of, 723 Lithologic characteristics of a fault zone,
Intersecting compressional faults, 240–43
292–94 Lithology logs, 520
Intersecting fault pattern, 281–83, J Little Sulphur Creek Basin, California,
419–26 679
Intersecting horst-graben faults, 282 Jura Mountains, Switzerland, 291, 545, Log correlation, 60–133
Interval isopach maps, 782–87 574 annotation and documentation,
construction, using seismic data, Jurassic Louann Salt, 686 129–33
786–87 Juxtaposition of rock units, 240–50 computer-based log correlation,
defined, 723–24 108–17
true stratigraphic thickness (TST) fault identification example,
from well logs, 783–85 K 114–17
Inversion structures, 715–18 on-screen log correlation, 109–14
structural styles, 717–18 KB, 61 transition to, 108
Isochore maps, 723–87 Kinematic processes, 548–49 correlation type log, 65–69
basic construction of, 728–42, Kinematics, 680 defined, 61
729–30 Kink band folds, 20 electric log correlation:
bottom water reservoir, 729–30 Kink bands, 551 basic concepts in, 71–72
edge water reservoir, 730–42 Kink law, 540–42, 662 defined, 61
818 Index
M N
O
Magnetic directional surveys, 50, 52 Narrow window correlation displays,
Map symbols, 788 117 Oblique cross section, 175
Marchand-Timbalier-Caillou Island Salt Natural neighbors, 26 Occam’s Razor, 337
Massif, 431, 433 Nearest “n” neighbors, 24–26 Offshore wells, drilled from a single
Master fault, 271, 314 Negative slope, 695, 715 platform, 43
Master synthetic fault, 602 Net/gross ratio, 728 Oil reservoirs, engineering, 781
Measured depth (MD), 44, 51, 61 Net/gross ratio map, 730, 742–44 Oil wedge, 730
Measured log thickness (MLT), 99–101 Net pay isochore maps, 725–728 On-screen log correlation, 109–14
Measurement-while-drilling (MWD) basic construction, 728 Onlap, 692–94
surveying, 50 bottom water reservoir, 729 Optimistic contouring, 338
Mechanical contouring, 16–18, 28, edge water reservoir, 730 Orinoco heavy oil belt, Venezuela, 46
341–44, 790 construction for a reservoir con- Out-of-sequence thrusts, 563, 564
Medina Anticline, Colombia, 450–53 taining oil and gas, 737–42 Overpressured shale zones, 605
Index 819
Overthrust faults, 292 Principal plane of stress, 642 Reservoir volume determinations,
Overthrust paradox, 508 Problem-solving cross sections, 183–85 775–79
Project plan, developing, 463–64 Horizontal Slice Method, 775,
Projected Slope gridding algorithm, 270 776–77
P Projection of wells, 204–17 from isochore maps, 775–79
cross sections, 204–17 selecting a method, 779
Painter Reservoir Field, Wyoming, 445 deviated wells, 213–17 Vertical Slice Method, 775, 777–78
Nugget Sandstone, 199–200 down-dip projection, 213 Restored surface method, computer
Panel diagrams, See Fence diagrams normal to the line of section mapping, 35–36
Paper-based log correlation, 108, (minimum distance method), limitations, 37
108–13, 117 213 procedure for, 36–37
Parallel contouring, 18–19, 342, 791 plunge projection, 206–9 Restored top estimation, 123–28
Parallel fold, 539 into a seismic line, 217 deviated wells, 125–28
Pecos County, New Mexico, 662 strike projection, 209–12 vertical wells, 123–25
Peripheral faults, 431 Pull-apart map, 453 Restraining bends, 636, 648–49
Pessimistic contouring, 338 Retrodeformation, 520–24
Phantoming, 141 Retrodeforming a structural cross
Philippine Fault, 639, 640, 642, 644, 646 section, complexity of, 518
Q Reversal of dip, 339
Philosophical Doctrine, 3–7, 457, 462,
465, 506 “Reverse drag folds”, 585
Quality of subsurface structural and
Picking/posting data, 166–69 Reverse faults, 289, 291–92
mapping methods, 2–3
extracting the data, 169 mapping, 447–50
Queen Formation, Guadalupe
posting the information, 169 mapping the surface of, 292
Mountains, 700–701
types of data from seismic, 166–69 Ridge Basin, California, 669–79
Picking thrust faults, 523–27 Rocky Mountains, 548
Piercing lines, 658 Rollover anticlines, origin of, 584–95
Piercing point/piercing line evidence: R Rollovers:
lateral displacements, 645–46 and growth faults, 314
defined, 645 Rabbit Island Fault, Gulf of Mexico, Roof thrust, 577
isopach/isochore information, 719–21 Rosecrans Oil Field, California, 440,
646 Rabbit Island Salt Spine, 431 444
local restoration, 648–56 Radial faults, 431 Rule of 45, 371–72
mountain ranges, 646 Radius of circle, equation to determine, Ryckman Creek Field, Wyoming, 445
offset zoned diapirs, 646 380–81
pre-growth strata, 645 Radius of curvature directional survey
regional restoration, 646–48 calculation, 50 S
releasing bends, 648, 649–52 Ramp angle, 44
restraining bends, 648, 652–56 directionally drilled wells, 44 “S” vergence, 582
Pin lines, 513, 513–14 Ramp geometry, 509 Sabine Uplift, Louisiana-Texas border,
Pine Mountain thrust region, Ramp and flat thrust faults, 294–98 393
Appalachians, 552 Raster well log formats, 108 Salt diapirs, 197–99, 431–40
Pitfalls in seismic interpretation, 308 Raton Basin, Colorado, 579 completed structural picture, 435–40
Plunge projection, 206–9 Reconnaissance seismic interpretation, contouring the salt surface, 432–33
Porosity top mapping, 763–766 469 hydrocarbon traps, 431–32
Post-depositional normal faults, 714 Recovery factor, 780–81 salt-fault intersection, 433
Post-depositional reverse faults, 714 Regional dip, 337–38 salt-sediment intersection, 433–35
“Postage stamp” map, 14 Relative age principle, 682 San Andreas Fault, California, 639, 642,
Posting data, 137–38 Releasing bends, 636, 648–49 646, 649, 652, 658–63, 703
Pre-Alps, 575 Repeated sections, 117–22, 255 San Gabriel Fault, California, 669–79
Pre-growth strata: and directional wells, 329 San Jacinto Fault, California, 649
isopach/isochore information based and fault surface maps, 329 Sand-shale distribution, in a reservoir, 728
on, 646 identification of a previously Sandstone/shale ratio formula, 628
and piercing line evidence, 645 unrecognized repeated section, Savanna Creek Duplex, Canada, 519
Preliminary structure mapping, using a 122 Scotland, Great Glen Fault, 646
computer, 495–501 Reservoir engineering, 779–82 Segno Field, Polk County, Texas,
drawing accurate fault gaps/overlaps, estimation of reserves, 779–81 630–33
495–99 gas reservoirs, 780–81 Seismic data, See also Three-
gridding and contouring, 499–501 oil reservoirs, 781 dimensional seismic data
Pressure differentials across a fault zone, field production history, 781–82 integration of well data and, for
240–43 reservoir characterization, 779 structure mapping, 390–92
820 Index
Seismic data (cont.) principal plane of, 642 contour compatibility, 400–403
translating seismic time models into and strike-slip faults, 640–42 contour compatibility across faults
seismic depth models, 518 measurements across, 640–42 application of, 403–7
tying, 146 trajectories, 641 exceptions to, 404–7
Seismic interpretation, 134–174, 308– Strike cross section, 175 contouring
314, 456–505 Strike-slip displacements, 635 anticlinal structures, 340
basic principles of, 134–174 scaling factors for, 656–57 change or reversal in the direction
pitfalls of, 308–14 Strike-slip fault tectonics, 441–45 of dip, 339
3D data, 456–505 Strike-slip faulted structures, 199 chosen reference, 334
Seismic line, projection of wells into, Strike-slip faults, 635–81 closed structural lows, 340
217 balancing, 658–79 constructing contours in groups of
Seismic reflection analysis, 575 bend widths, 657 several lines, 337
Seismic time section, 141–43 compressional folding along, 658–68 contour compatability, 339
vertical dimension on, 143 criteria for, 643–45 contour interval, 334–35
Semangko Fault, Indonesia, 639, 642 defined, 441 contour license, 340
Separation, on a fault, 255–56 extensional folding along, 668–79 contour spacing, 335–36, 340
Sequence boundaries, 150 geometry of, 672–79 control points, 337
Sequence stratigraphic analysis, 150 Ridge Basin geology, 669–72 domal structures, 340
Sequence stratigraphy, 134 and horizontal displacements, 645 equal-spaced contouring, 342
Setchell’s equation, 753–54, 760–63 interpretation, problem of, 637–38 faulted surfaces, 342–56
correction factor (CF) in, 761–62 lack of data, 656 graphic scale, 336
Shale resistivity markers (SRMs), 73 lateral displacements, 645–56 guidelines, 334–41
Shared earth model, 463 piercing point/piercing line hachured lines, 336–37
Short-radius horizontal wells, 44–46 evidence, 645–46 index contour, 336
Sideswipe, 135 surface features, 645 interpretive contouring, 342
Signal Hill, California, 652–656, 702 mapping, 636–43 mechanical contouring, 341
Single compressional faults, 292 strain ellipse model, 638–40 optimistic, 338
“Single shot” magnetic survey, 50 stresses, 640–42 parallel contouring, 342
Slip, on a fault, 255–56 overinterpretation, 656 in pencil, 337
Small velocity problems, accounting for, stress measurements across, 640–42 pessimistic, 338
171–73 Strong horizontal velocity gradient, 170 reflect as a geologic interpretation,
Smooth contours, 337 Structural balancing, 507–8 338
Snakehead structure, 558 basis of, 507 regional dip, 337–38
Sonic logs, 520 benefits of, 507 smooth style of, 337
South America: classical balancing, 508 structural noses, 340–41
Andes Mountains, 62 computer-aided, 515–20 widening of contours, 340
Eastern Condillera, Colombia, 298 nonclassical balancing, 508 fault gaps, defined, 383
South Marsh Island (SMI), offshore Gulf ultimate goals of, 507 fault heave, defined, 383
of Mexico, 305 Structural cross sections, 176–82, fault overlap, 369
South Pelto area, Gulf of Mexico, 109, 237–40 fault traces/fault gaps, 369–83
114 construction of, using the computer, equation to determine heave,
Southern Permian Basin, North Sea, 237–40 382–83
431, 435 defined, 176 equation to determine radius of
St. Charles Ranch, Texas, 619 drawing, 176–79 circle, 380–81
Stacked Multiple Bischke Plot, 705 drawing with the same horizontal and new circle method, 378–80
use of, 709–11 vertical scales, 180 Rule of 45, 371–72
Stacking, in computer mapping, 31–34, electric log sections, 180–82 tangent method/circle method,
36–37 stick sections, 182 372–78
Static mis-ties, 160 Structural fence diagram, 227 generic case study, 383–87
Stick cross sections, 182 Structural highs, 340 integration of seismic and well data
Straight-line cross sections, 182, 196–97 Structural modeling, 518–20 for, 390–92
Strain ellipse model, 638–40 Structural (nongrowth) fault, 307–8 manual integration of fault map and,
Strategies, fault interpretation from Structural noses, 340–41 357–69
seismic data, 472–79 contouring, 340–41 normal faults, 357–66
Stratigraphic cross sections, 182–83 Structural styles, and growth, 683 restored tops, 362–66
construction of, 235–37 Structure maps, 332–455 reverse faults, 366–69
Stratigraphic fence diagram, 227 additive property of faults, 387–90 mapping across vertical faults, 397–98
Stratigraphic type log, 65–66 closely spaced horizons, 400–403 multiple horizon mapping, 453–55
Stress, 640–42, 680–81 completeness of the work undertaken, discontinuity of structure with
as a mathematical concept, 640 453–55 depth, 455
Index 821
subsurface structure map: Syntectonic sedimentation, 714, 722 quality-checking fault surfaces in
construction of, 333 Synthetic seismograms, 140 map and seismic views, 479–80
importance and reliability of, 332 tying well data to, 152–54 reconnaissance, 469
tectonic habitat mapping techniques, strategies, 472–79
407–53 well control, integrating, 469–72
compressional tectonics, 445–53 T framework interpretation and
diapiric salt tectonics, 431–40 mapping, 462–63
extensional tectonics, 408–30 Taiwan, Chuhuangkeng Anticline, 576 horizon and fault integration on a
strike-slip fault tectonics, 440–45 Tangent method, 372–78 workstation, 501–5
top of structure versus top of porosity, Tangential directional survey horizon interpretation, 480–95
398–400 calculation, 50 framework horizons, selecting,
unconformities, 392–97 Taranaki Basin, New Zealand, 550 480–87
angular unconformity, 393 Teamwork, 463 infill strategies, 492–95
defined, 392–93 Tectonic habitat mapping techniques, interpretation strategy, 488–92
disconformity, 393 407–53 well data, tying, 487–88
mapping techniques, 393–97 compressional tectonics, 445–53 interpretation of, 456–505
Subsea true vertical depth (SSTVD), 44, reverse faults, 447–50 interpretation workflow, developing,
61 thrust faults, 450–53 464
Subsurface geological maps, importance diapiric salt tectonics, 431–40 optimizing displays for better results,
of, 1–2 completed structural picture, 458–62
Subsurface geoscientists: 435–40 optimizing the data, 457
and accurate geologic interpretations, contouring the salt surface, Philosophical Doctrine, 456–57
5–6 432–33 preliminary structure mapping,
role of, 2 hydrocarbon traps, 431–32 495–501
Subsurface mapping: salt-fault intersection, 433 drawing accurate fault
basics of, 1–7 salt-sediment intersection, 433–35 gaps/overlaps, 495–99
Philosophical Doctrine of accurate extensional tectonics, 408–30 gridding and contouring, 499–501
interpretation/mapping, 3–7 bifurcating fault pattern, 415–19 project plan, developing, 463–64
Subsurface mapping, from seismic data, combined vertical separation, 426 project setup, 458
134–174, 456–505 compensating fault pattern, teamwork, 463
data extraction, 166–73 410–15 workstation project, organizing, 465
converting time to depth, 169–73 exceptions to, 408 Three-dimensional structural
extracting the data, 169 growth faults, 427–30 workstation software, 518
information posting, 169 inclusions, 408 Three-dimensional views, cross sections,
picking/posting, 166–69 intersecting fault pattern, 419–26 223–34
types of data from seismic, strike-slip fault tectonics, 440–45, fence diagrams, 223–27
166–69 441–45 isometric projections, 228
data validation and interpretation, Tectonostratigraphic anomalies, 688–89 log maps, 223
139–66 Texas: three-dimensional reservoir analysis
examining the seismic sections, Bethel Dome, 440 model, 228–34
139–43 Burgentine Lake Field, 599–600, 602 3D seismic data sets, advantage of, 143
tying seismic data, 143–66 St. Charles Ranch, 619 Throw, 255, 256, 262, 347, 383
integration of geophysical data in, Wilcox strata, 239, 242 mathematical relationship of, to
134–74 Woodbine Sandstones, 393, 395 vertical separation, 256–60
assumptions/limitations, 135–36 Theory of Elasticity, 680 and missing section/repeated section,
process, 136–38 Thick-skinned strike-slip environment, 255
3D data, 456–505 643 subsurface petroleum-related
Subsurface maps, types of, 6 Thick-skinned vertical faults, 656 structure mapping, 347
Subsurface petroleum geology, Thin-limb fold, 539–40 Throw map, 307
objectives of, 1 Thin-skinned compressional Thrust faults, 291–92
Subtle unconformities, 692 environment, 643 in structure mapping, 450–53
Superposition principle, 682 Thin-skinned, low-angle faults, 656 Tight streak, 399–400, 763
Suppe’s Assumptions, 549 Three-dimensional perspective, structure Time—depth table, 169–70
Surface modeling, 23 contour map, 9–12 Time section, 141–43
Symmetric fold type, 549–52 Three-dimensional reservoir analysis Top of porosity map, 398
Synclinal breakthrough, 560, 563 model, 228–34 Total depth (TD), 65
Synclines, contouring of, 340 Three-dimensional seismic data, Totco tool, 50
Syndepositional faults, 71, 314 465–505 Trace fault, on a map, 35
Synergistic teams, 463 documentation, 465 Transfer structures, 529
Synergy, 138, 463 fault interpretation, 465–80 Transfer zone, 293–94
822 Index
Transverse Ranges, California, 652, Upthrown (hanging wall) restored top, key point in, 745
680–81 123 procedure for, 745–48
Transverse structures, 612–14 Utah, Fossil Basin, 298 purpose of, 744
Trapezoidal directional survey reservoirs with significant shale
calculation, 50 intervals, 748–52
Triangle zones, 527, 577–78 V Water coning, and horizontal wells, 46
with passive roof backthrust, 577–78 Water wedge, 730
Triangulation, 23, 27–28 Velocity data: Wedge structures, 577–78
steps involved in, 27–28 problems with, 170–171 Wedge zone, 732, 738–39, 764
True dip seismic line, 164 Venezuela, 527–28, 705–9, 715, 717 Well control, integrating in fault
True stratigraphic thickness (TST), Vertical conductivity of a fault, 241–43 interpretation, 469–72
79–80, 99 Vertical exaggeration, 176, 180, 201–4 Well data:
logs, 754–55, 782–85 Vertical point, deviated well, 44 integration of seismic data and,
True vertical depth thickness (TVDT), Vertical resolution of a well log, 140 134
99, 754–55 Vertical seismic profile (VSP), 153–54, for structure mapping, 390–92
True vertical depth (TVD), 44, 61 488 tying, 487–88
cross section, 102–4 Vertical separation, 34–35, 37, 217–18, Well deviation angle, 51
logs, 755–57 254, 261–63, 289, 305, 329 Well deviation surveys, 520
True vertical thickness (TVT), 31–32, combined, 283–89 Well log sticks, in cross sections, 176
79, 86, 99, 329, 752, 752–59 for a growth fault, 321–22 Well logs, fault data determined from,
logs, 755–57 mathematical relationship of throw 260–63
three-dimensional correction factor to, 256–60 West Cameron Block 192, northern Gulf
equation, 94–96 Vertical separation map, 307, 320–322 of Mexico, 216–17
Two-dimensional structural workstation Vertical separation versus depth (VS/d) West Ridge cross fault, Yucca
software, 518 method, 711–21 Mountains, Nevada, 617
Tying seismic data: analysis examples, 713–21 Wharton Method, 732, 738
concepts of complex growth structures, 715–18 Wheeler Ridge, California, 578, 581
loop-tying as a proof of downward-dying growth faults, Whitney Canyon Field, Wyoming,
correctness, 146–48 718–21 445
tying loops, rationale for, 143–46 growth normal faults, 714–15 Wilcox strata, South Texas, 239, 242
mis-ties, 160–66 growth reverse faults, 715 Wilmington Field, Los Angeles Basin,
migration mis-ties, 160–66 post-depositional (nongrowth) California, 440, 445
static mis-ties, 160 faults, 714 Woodbine Sandstones, East Texas, 393,
procedures in, 149–60 technique, 712–13 395
annotating the well information, Vertical Slice Method, 775, 777–78 Workstation, horizon and fault integra-
150 Vertical thickness: tion on, 501–5
contemplating the data, 149 and fluid contacts in deviated wells, Workstation project, organizing,
picking a reflection to interpret 759–63 465
and map, 149–50 impact of correction factors, 757–59 Workstation software, 518
tying the faults, 154–55 Vertical well log correlation, 69–80 Wyoming, 652
tying the lines and horizons, 155–60 fault determinations, 73–76 Fossil Basin, 298
tying well data to seismic with faults versus variations in strati- Green River basin, 113, 236
checkshot information, 150–52 graphy, 73 Utah backarc fold-and-thrust belt
tying well data to seismic with log correlation plan, 69–71 fields, 445
synthetic seismograms, 152–54 pitfalls in, 79–80
stratigraphic variations, 76–78
Vertical wells, 69–80, 123–25, 327 Y
U log correlation plan, 69–71
restored top estimation, 123–25 Y-O Buckle, New Mexico, 668
Uinta Basin, Colorado, 202–4 Violin Breccia, San Gabriel Fault, Yucca Mountains, Nevada, 614, 617
Unconformities, 128–29, 690–91, 711 California, 669
angular, 129 Visualization software, 247
on dipmeter, 129 Volume accountability rule, 510–11 Z
on electric logs, 129 Volumetric configuration of reservoirs,
mapping techniques, 392–97 774–775 “Z” vergence, 582
and hydrocarbon traps, 128, 392–97 Zagros collisional belt, Iran, 445
recognizing, during correlation, 129 Zayante Fault, California, 703–4
Unconformity wedge zone, 397 W Zero bed dip, 752–53
Undersaturated oil, 781 Zigzag cross sections, 196
Unit isochore map, 32–34 Walking wells, 744–52 Zone of combined vertical separation,
United States, growth faults, 714–15 defined, 744 283