You are on page 1of 310

Seals, Traps,

and the
Petroleum System

Edited by
R.C. Surdam
Institute for Energy Research
University of Wyoming
Laramie, Wyoming

AAPG Memoir 67

Published by
The American Association of Petroleum Geologists
Tulsa, Oklahoma, U.S.A. 74101
Copyright © 1997
The American Association of Petroleum Geologists
All Rights Reserved
Published 1997
Printed and bound in the United States of America

ISBN:0-89181-347-0

Seals, traps, and the petroleum system/edited by R.C. Surdam.


p. cm. -- (AAPG memoir: 67)
Includes bibliographical references (p. ) and index.
ISBN 0-89181-347-0 (alk. paper)
1. Traps (Petroleum geology ) I. Surdam, Ronald C. II. Series.
TN870.57.S43 1997 97-20684
553.2’8--dc21 CIP

AAPG grants permission for a single photocopy of an item from this publication for personal use. Authorization
for additional copies of items from this publication for personal or internal use is granted by AAPG provided that
the base fee of $3.00 per copy is paid directly to the Copyright Clearance Center, 222 Rosewood Drive, Danvers,
Massachusetts 01923. Fees are subject to change. Any form of electronic or digital scanning or other digital trans-
formation of portions of this publication into computer-readable and/or transmittable form for personal or cor-
porate use requires special permission from, and is subject to fee charges by, the AAPG.

Association Editor: Kevin T. Biddle


Science Director: Richard Steinmetz
Publications Manager: Kenneth M. Wolgemuth
Special Projects Editor: Anne H. Thomas
Production: Custom Editorial Productions, Inc., Cincinnati, Ohio

THE AMERICAN ASSOCIATION OF PETROLEUM GEOLOGISTS (AAPG) DOES NOT ENDORSE


OR RECOMMEND ANY PRODUCT OR SERVICES THAT MAY BE CITED, USED OR DISCUSSED
IN AAPG PUBLICATIONS OR IN PRESENTATIONS AT EVENTS ASSOCIATED WITH THE AAPG.

This books and other AAPG publications are available from:

The AAPG Bookstore Geological Society Publishing House


P.O. Box 979 Unit 7, Brassmill
Tulsa, OK 74101-0979 Enterprise Centre
Telephone: (918) 584-2555; Brassmill Lane
or (800) 364-AAPG (USA—book orders only) Bath BA1 3JN
Fax: (918) 584-2652 United Kingdom
or (800) 898-2274 (USA—book orders only) Tel 0225-445046
Fax 0225-442836

Australian Mineral Foundation Canadian Society of Petroleum Geologists


AMF Bookshop #505, 206 7th Avenue S.W.
63 Conyngham Street Calgary, Alberta T2P 0W7
Glenside, South Australia 5065 Canada
Australia Tel (403) 264-5610
Tel (08) 379-0444
Fax (08) 379-4634

ii

About the Editor


Ronald C. Surdam serves as director of the Institute for Energy Research


(IER), a multidisciplinary research organization founded in November 1993 at
the University of Wyoming to address problems relating to the exploitation of
natural gas resources, clastic diagenesis, thermal modeling, source rock matura-
tion, petrophysics, and pressure compartmentalization. He received an A.B. in
geology in 1961 and a Ph.D. in geology in 1967 from the University of California
at Los Angeles. Dr. Surdam joined the geology faculty at the University of
Wyoming in 1966, and has served as professor of geology since 1973. He was
selected as an AAPG Distinguished Lecturer (1985–86), the Don R. and Patricia
Boyd Distinguished Lecturer in Petroleum Exploration (1990), and the AAPG
Roy M. Huffington Distinguished Lecturer (1995–96). Dr. Surdam has served on
the National Science Foundation Geology/Geochemistry Advisory Panel
(1980–83), the U.S. Continental Scientific Drilling and Review Group (1990), the
American Chemical Society Advisory Panel (1988–93), and was elected Fellow in
the Geological Society of America (1989). He has presented 171 invited lectures,
taught numerous short courses and seminars for a variety of corporations and
scientific societies, and is the author or coauthor of 170 publications. Dr. Surdam
previously served as Associate Editor of the Bulletin of the Geological Society of
America (1989), and edited AAPG Memoir 37.
Dr. Surdam’s current and recent research interests include the San Joaquin
and Coastal basins of California, the Gulf Coast, the Western Canada Basin, the
Rocky Mountain Laramide basins, the Potiguar and Renconcavo basins of Brazil,
the San Jorge Basin of Argentina, the Sinu Basin of Colombia, the Maracaibo
Basin of Venezuela, the Mahakam Delta of Indonesia, and the Gippsland Basin
of Australia. In all of these areas, he is generally interested in the petroleum sys-
tem, with special interest in source-reservoir rock relations; fluid-flow character-
istics, including sealing mechanisms; clastic diagenesis; and determination of the
spatial attributes of hydrocarbon accumulations.

iii

Acknowledgments

Many thanks and congratulations to the following authors and technical reviewers who worked diligently to
make this volume possible:

AUTHORS
David M. Allard Henry P. Heasler John T. Leftwich Hugh W. Reid
C.D. Atkinson Susan J. Hippler T. Leslie Leith John H. Sales
George W. Bolger Huseyin Is Debi T. Maucione John Sneider
Peter J. Boult Zun Sheng Jiao Randi S. Martinsen Robert M. Sneider
Terry Engelder John G. Kaldi John W. Neasham Ronald C. Surdam
A.E. Fallick Mark R. Krolow James C. Niemann Paul N. Theologou
J. Foden James Krushin Ron A. Noble Kadir Uygur
M. Arif Yükler

REVIEWERS
William R. Almon William P. Iverson John W. Neasham Stephen A. Sonnenberg
David K. Baskin John G. Kaldi Peter D’Onfro Robert M. Sneider
Mark P. Fischer Randi S. Martinsen Olivier Poix Ronald J. Steele
Willim Higgs John McKay John H. Sales Brian Towler
Peter Huntoon Norman R. Morrow Leta K. Smith Charles L. Vavra
John Warme

I would like thank the Gas Research Institute (GRI) of Chicago for generously supporting various research pro-
grams at the Institute of Energy Research (IER); it is fair to say that the bulk of the work presented in Section III of
the memoir would not have been undertaken without GRI support. In fact, the innovative ideas of Dave Powley
(GRI; formerly of Amoco) provided the impetus to study and apprehend concepts relating to pressure seals and
compartments; subsequently, GRI has provided many resources and encouragement to accomplish this task.
In addition, I would like to acknowledge Alice Rush, David Copeland, and Laura Vass of the IER at the
University of Wyoming for their contributions to the volume. Allory Deiss (IER) created and improved graphics
for several authors in the memoir, and has been an invaluable part of this effort. I would also like to thank my
co-editor Kathy Kirkaldie (IER), who kept the effort progressing by accomplishing large and small tasks with
enthusiasm and optimism and always with a great deal of encouragement.

iv

Foreword

P etroleum explorationists, in the ever increasingly


difficult search for hydrocarbons, presently place
significant emphasis on integrated studies of
After all, in many places in the world, our knowledge
of structural settings and source rock attributes have
been enhanced greatly, yet in these areas forecasting
source, conduit, seal and reservoir lithologies, and efficiencies do not begin to approach 68%.
traps (Magoon and Dow, 1994, and others in AAPG This AAPG memoir is a serious attempt to reduce
Memoir 60). R.J. Murris (1984), in the introduction to uncertainty with respect to hydrocarbon traps and
AAPG Memoir 35, offers a real and graphic display of seals. Armed with the knowledge available in this vol-
the importance of such integrated approaches to oil ume and the experiences willingly shared by its
and gas exploration. In his figure 1, based on an actual authors, explorationists will be able to get beyond the
case of 165 prospects, Murris (1984) nicely illustrates essential knowledge threshold and begin to take the
that ranking prospects by trap size alone increases steps necessary to acquire a better conceptual under-
forecasting efficiency of hydrocarbon accumulations standing of hydrocarbon traps and seals and produce
by 28% over a random drilling order. For quite some the diagnostic tools we so desperately need to expedite
time it has been apparent that the majority of success- their development. In my opinion, in most frontier
ful wells are sited on structural closures, but it has prospective areas, 70% forecasting efficiencies will
been equally apparent that the majority of unsuccess- only become a reality when techniques and technolo-
ful wells also are sited on structural closure. Murris gies are developed to detect, delineate, and quantify
(1984) demonstrates that by including geochemical the capacity of sealing lithologies, and the critical char-
charge and retention parameters (e.g., source potential, acteristics of trapping mechanisms are documented.
maturation history, and expulsion efficiency), a fore- This memoir is divided into the following sections:
casting efficiency as high as 63% can be obtained. (I) Seal Characteristics: Processes Controlling Sealing
Murris (1984) argues that it is possible—by integrating Capacity; (II) Traps: Hydrocarbon Seals in a Regional
source rock studies with structural aspects during the Context; and (III) Pressure Seals and Fluid
evaluation of a prospect—to reduce exploration uncer- Compartments. R.M. Sneider, J.S. Sneider, G. Bolger,
tainty significantly. and J.W. Neasham start Section I with a description
Magoon and Dow (1994) promoted a holistic of a new technique to measure sealing capacity using
approach to the evaluation of a “petroleum system” cuttings (i.e., hydrocarbon column heights). Prior to
by formalizing the integration of data on essential this work, analytical techniques were restricted to
elements and processes characterizing a petroleum core samples; because the coring of fine-grained
accumulation. Ultimately, they hoped that this lithologies occurs mainly by accident, quantitative
approach would reduce the risk of exploring for data on seals remained scarce. With the advent of the
hydrocarbon plays and prospects. They considered techniques outlined in this paper, essential seal data
source rocks, reservoir rocks, seal rocks, and over- should become much more readily available. A paper
burden rocks as the essential elements of a petrole- by R. Noble, J. Kaldi, and C.D. Atkinson discusses the
um system, and trapped formation and the genera- role of hydrocarbons in fine-grained lithologies and
tion, migration, and accumulation of petroleum, the relates it to seal performance. J. Krushin’s paper
vital processes affecting it. examines the use of pore-throat distribution to seal
After much discussion on petroleum systems with performance. In contrast, J. Niemann and M.R.
many exploration geologists, it has become clear to Krolow discuss a method to determine if faults exhib-
me that the least understood element in a petroleum it sealing or leaking characteristics.
system typically is the seal, and the least understood Section II of this memoir addresses seals in a
process in a petroleum system is the trapping of regional context; in other words, the trapping of
hydrocarbons. To be sure, if explorationists are to hydrocarbons. J. Sales starts the discussion with a
reach the 68% forecasting efficiency described by summary of his lifetime effort to understand the
Murris (1984), we must substantially improve our relationship between seal strength and trap closure.
understanding of seals and trapping mechanisms. This paper is followed by J. Kaldi and C.D. Atkinson

v
vi Surdam

with a neat case history that demonstrates the value In the last paper in this section, R.C. Surdam intro-
of evaluating sealing potential. Next the memoir duces a new exploration paradigm used in the search
includes papers by S.J. Hippler, and T.L. Leith and for anomalously pressured gas/condensate accumu-
A. Fallick that discuss the role of fractures, diagene- lations, or so-called "pressure compartments." The
sis, and organics in seals associated with North Sea memoir concludes with an appendix by H. Reid, in
oil fields. D. Allard discusses fault seal interpreta- which analyses of drillstem test data are used to more
tion techniques that may improve our understand- accurately assess the leakage potential and reservoir
ing of basin-specific seal integrity. P. Boult, P. quality of tight formations, even potential seals.
Theologou, and J. Foden continue the discussion of In summary, I believe that every explorationist can
hydrocarbon traps with a chapter on seal leakage learn something from the articles contained in this
and capillary properties in the Eromanga Basin of volume. I hope all who read it will more fully appre-
central Australia. This section of the memoir con- ciate the role of seals and traps in determining the
cludes with a discussion of carbonate hydrocarbon spatial distribution of hydrocarbon accumulations.
reservoirs and seals in southeast Turkey by K. This AAPG memoir is the outgrowth of a very timely
Uygur, H. Is, and M. Yükler. Hedberg Conference sponsored by the AAPG—a con-
Section III of the memoir deals with pressure seals ference that clearly demonstrated communication is
and pressure compartments, or anomalously pres- the key to integrated scientific endeavors.
sured hydrocarbon accumulations. R.C. Surdam, Z. Good luck and success in your search for the ever-
Jiao, and H. Heasler start the discussion by describing elusive hydrocarbon.
anomalously pressured gas accumulations in Rocky
Mountain Laramide basins. They suggest that gas-sat- Ronald C. Surdam
urated pressure compartments are a new class of
hydrocarbon accumulation. Next, R. Martinsen deter- REFERENCES
mines the difference between a stratigraphic trap and
a fluid-pressure compartment. Stratigraphic elements Magoon, L.B., and W.G. Dow, 1994, The petroleum
and pressure seals are discussed by Z. Jiao and R.C. system, in L.B. Magoon and W.G. Dow, eds., The
Surdam, whereas T. Engelder and J. Leftwich present petroleum system from source to trap: AAPG
a case history of the role of faults in the establishment Memoir 60, p. 3–24.
of pressure seals and compartments. D. Maucione Murris, R.J., 1984, Introduction, in G. Demaison and
and R.C. Surdam next discuss the remote detection of R.J. Murris, eds., Petroleum geochemistry and
regional pressure seals using seismic reflection data. basin evolution: AAPG Memoir 35, p. x-xii.

vi

Table of Contents

Foreword R.C. Surdam

Section I Seal Characteristics: Processes Controlling Sealing Capacity


1 Comparison of Seal Capacity Determinations: Core vs. Cuttings . . . . . . . . . . . . . . . . . . . . . . .1
R.M. Sneider, J.S. Sneider, G.W. Bolger, and J.W. Neasham

2 Oil Saturation in Shales: Applications in Seal Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .13


R.A. Noble, J.G. Kaldi, and C.D. Atkinson

3 Seal Capacity of Nonsmectite Shale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31


J.T. Krushin

4 Delineation of a Pressure Fault Seal from Shale Resistivities . . . . . . . . . . . . . . . . . . . . . . . . . .49


J.C. Niemann and M.R. Krolow

Section II Traps: Hydrocarbon Seals in a Regional Context


5 Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution
of Oil and Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .57
J.K. Sales

6 Evaluating Seal Potential: Example from the Talang Akar Formation,


Offshore Northwest Java, Indonesia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .85
J.G. Kaldi and C.D.Atkinson

7 Microstructures and Diagenesis in North Sea Fault Zones: Implications


for Fault-Seal Potential and Fault Migration rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .103
S.J. Hippler

8 Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field,


Norwegian North Sea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .115
T.L. Leith and A.E. Fallick

9 Fault Leak Controlled Trap Fill: Rift Basin Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .135


D.M. Allard

10 Capillary Seals Within the Eromanga Basin, Australia: Implications


for Exploration and Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .143
P.J. Boult, P.N. Theologou, and J. Foden

11 Reservoir Characterization of Cretaceous Mardin Group Carbonates in


Bölükya-Cukurtas and Karakus Oil Fields, SE Turkey: A Petrographic
and Petrophysical Comparison of Overthrust and Foreland Zones . . . . . . . . . . . . . . . . .169
K. Uygur, H. Is, and M.A.Yükler

vii
viii Table of Contents

Section III Pressure Seals and Fluid Compartments


12 Anomalously Pressured Gas Compartments in Cretaceous Rocks
of the Laramide Basins of Wyoming: A New Class of
Hydrocarbon Accumulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .199
R.C. Surdam, Z.S. Jiao, and H.P. Heasler

13 Stratigraphic Controls on the Development and Distribution of


Fluid-Pressure Compartments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .223
R.S. Martinsen

14 Characteristics of Anomalously Pressured Cretaceous Shales in the


Laramide Basins of Wyoming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .243
Z.S. Jiao and R.C. Surdam

15 A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields . . . . . . . . . . . . .255
T. Engelder and J.T. Leftwich

16 Seismic Response Characteristics of a Regional-Scale Pressure


Compartment Boundary, Alberta Basin, Canada . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .269
D.T. Maucione and R.C. Surdam

17 A New Paradigm for Gas Exploration in Anomalously Pressured


“Tight Gas Sands” in Rocky Mountain Laramide Basins . . . . . . . . . . . . . . . . . . . . . . . . .283
R.C. Surdam

Appendix Evaluating Seal Facies Permeability and Fluid Content from


Drill-Stem Test Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .299
H.W. Reid

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .313
AAPG
Wishes to thank the following
for their generous contributions
to

Seals, Traps, and the Petroleum System

Amoco Production Company

Mobil Exploration and Producing Technical Center

PetroTech Associates

John K. Sales

Contributions are applied against the production


costs of publication, thus directly reducing the book’s
purchase price and making the volume
available to a greater audience.

ix
Sneider, R.M., J.S. Sneider, G.W. Bolger, and J.W.
Neasham, 1997, Comparison of seal capacity
determinations: conventional cores vs. cuttings, in
R.C. Surdam, ed., Seals, traps, and the petroleum
system: AAPG Memoir 67, p. 1–12.

Chapter 1

Comparison of Seal Capacity Determinations:


Conventional Cores vs. Cuttings
Robert M. Sneider
John S. Sneider
Robert M. Sneider Exploration, Inc.
Houston, Texas, U.S.A.

George W. Bolger
John W. Neasham1
PetroTech Associates
Houston, Texas, U.S.A.

ABSTRACT
Comparison of hydrocarbon column heights (HCHs) calculated from seals
recovered in conventional cores with HCHs calculated by using cuttings
from the same interval indicates that mercury/air capillary pressure mea-
surements of cuttings can be extremely useful to estimate seal capacity. An
empirical adjustment factor (EAF), expressed in psi, needs to be added to the
capillary pressure value determined on cuttings to approximate that mea-
sured with mercury/air capillary pressure of conventional cores.
For top and lateral seals that are the result of lithologic changes (as
opposed to fault seals), good to excellent agreement is found between the
hydrocarbons actually trapped in fields and the HCH calculated from mer-
cury/air capillary pressure curves of vertical plugs cut perpendicular to the
sealing surface. The plugs are sealed with epoxy so that mercury can enter
only from the top and base of the plug. The mercury/air capillary pressure
curves are generated using a system that can inject mercury at pressures up
to 60,000 psi [8703 kPa] (equivalent to a hydrocarbon column of >10,000 ft
for 35° API gravity oil and normal saline water).
Depending upon seal type, high-pressure mercury/air injection curves
(HPMIC) of cuttings can be used to approximate those of samples from con-
ventional cores. Injection pressures for cuttings are usually lower than those
from equivalent cores for a particular percent pore volume occupied by mer-
cury. Empirical adjustment factors (EAFs), expressed in psi, for different seal
types are derived from comparisons of HPMIC on epoxy-sealed vertical
conventional core plugs with cuttings or “simulated cuttings” of the same
seal interval. The EAF values are added to the capillary pressure measure-
ment of cuttings to obtain the approximate value of mercury/air capillary

1Now with Poro Technology, Houston, Texas

1
2 Sneider et al.

pressure of a vertical plug. The EAF vary from ~1900 psi (mercury/air) for
type “A” seals to ~25 psi for type “D” seals, using 7.5% mercury pore vol-
ume saturation as the reference saturation.
Careful sample preparation and accurate closure corrections are critical to
obtaining accurate HPMIC measurements and corresponding EAF values for
HCH calculations.

INTRODUCTION We have estimated the HCHs in more than 200


reservoirs where low-permeability lithologies (not
This paper addresses the question: “How useful are faults) are the seals. We observe that the estimated
cuttings of seals to estimate the hydrocarbon column HCHs of hydrocarbon-bearing reservoirs correspond
heights?” This is an important question because cut- to the mercury/air capillary pressure between ~5%
tings of seal lithologies are very common, but cores of and 10% nonwetting phase saturation. We have used
seals are rare. Pc at 7.5% nonwetting phase saturation to indicate the
During the past 25 yr, we have been coring seals and saturation at which the seal actually leaks hydrocar-
collecting cuttings while coring or crushing seal cores bons through its pore network. This saturation is
to produce “simulated cuttings.” Figure 1 is an exam- called “breakthrough” or “leakage” saturation.
ple of a cored seal–reservoir interface. A comparison of Figure 3 shows capillary pressure vs. nonwetting
high-pressure mercury/air injection capillary pressure phase curves for the mercury/air system for seals “A”
curves (HPMICs) of both the cores and associated cut- and “D.” Using the assumed densities of oil, gas, and
tings or simulated cuttings shows that it is possible to water given and interfacial tensions (in dynes/cm) of
estimate the capillary pressure equivalent to a vertical 30, 70, and 480 for oil/water, gas/water, and mer-
seal lithology from HPMICs of cuttings by adding an cury/air systems, mercury capillary pressures are con-
empirically derived adjustment factor (EAF) in psi. verted to oil/water and gas/water systems. If the seal
This paper reviews some principles of hydrocarbon capillary pressure (Pc) for leakage is taken at 7.5% non-
entrapment, discusses sample preparation, and pre- wetting phase saturation, the “A” seal will hold ~1200
sents examples of empirical relationships between ft of oil or 520 ft of gas before leakage through the seals
HPMICs of core samples and associated cuttings. occurs. The “D” seal will hold ~120 ft of oil or ~50 ft of
gas before leakage through the seal. The equations to
convert mercury/air capillary pressure to oil/water
HYDROCARBON ENTRAPMENT and gas/water capillary pressure and to HCH are out-
Schowalter’s 1979 paper on seals reviews the princi- lined in Appendix 1.
ples of hydrocarbon entrapment/accumulation. A It is important to remember that when mercury
brief review of the key principles is presented to injection capillary pressure (MICP) data are converted
explain the entrapment/accumulation processes and to reservoir conditions, the values of oil/gas/water
how mercury/air measurements relate to hydrocar- densities and interfacial tensions must be corrected for
bon/water capillary pressure and, in turn, to hydrocar- reservoir temperature and pressure.
bon column height (HCH) trapped against seals.
Figure 2 is a schematic of a stratigraphically SEAL TYPES
trapped reservoir with its adjacent seals. Hydrocar-
bons entering the reservoir are driven by the buoyancy Sneider et al. (1991) studied several hundred seal litholo-
force or pressure (Pb), which is the difference in den- gies and presented an arbitrary classification of seals based
sity of the hydrocarbons and formation water × the on the hydrocarbon column held. The seal types and
hydrocarbon column height (h) × 0.433, the gradient hydrocarbon columns held are shown in Table 1.
of fresh water. Hydrocarbons entering the reservoir The study used mercury/air capillary pressure
must enter the pores and displace the pore water. The curves, which were converted to an oil/water capillary
equation of this resistive force (Pc, a rock’s capillary system assuming 35° API gravity oil and normal saline
pressure) is shown in Figure 2. Hydrocarbons will water. The oil/water capillary system was converted
continue to fill the reservoir and be trapped against the into the hydrocarbon column height (HCH) held
seals until the buoyancy pressure due to the hydrocar- before leakage through the seal. Leakage is assumed to
bon–water system in the reservoir exceeds the capil- be where Pc equals 7.5% nonwetting phase saturation
lary entry pressure (Pe) of the weakest seal rock. In based on a comparison of HCH observed in the field
water-wet or mostly water-wet systems, hydrocarbons and the column height calculated from the
will continue to leak into the seals until a balance or mercury/air capillary pressure curve when the satura-
equilibrium is reached between the seal entry pressure tion is 7.5%. In Figure 3, the mercury/air capillary
and the pressure within the reservoir system. pressure at P c = 7.5% nonwetting phase saturation
Seal Capacity from Cores vs. Cuttings 3

RESERVOIR–SEAL COUPLET Figure 1. Core of an


anhydrite top seal on a
LITHOLOGY SEAL TYPE Pc @ 7.5% Hg dolomite reservoir, San
(OIL) Saturation (psi) Andres Formation, New
Mexico.
ANHYDRITE A >50,000

“CHICKEN WIRE”
B 809 – 2395
ANHYDRITE

TIGHT DOLOMITE
PARTIALLY
D–E 65 – 130
REPLACED BY
ANHYDRITE

RESERVOIR
DOLOMITE 10 – 80
ROCK

converts to an oil/water capillary pressure for “A” and SAMPLE PREPARATION AND
“D” seals of 113 psi and 11.4 psi, respectively, which in CAPILLARY PRESSURE
turn converts to a HCH held of ~1200 and 120 ft,
respectively, for the “A” and “D” seals. In other words, MEASUREMENTS
the “A” seal will hold a HCH of ~1200 ft if the oil is 35°
API and the pore water is normal saline water. If more Conventional core samples are cut perpendicular to
oil enters the reservoir, hydrocarbons will leak through bedding or perpendicular to the potential sealing sur-
the seal. The hydrocarbon column held by a seal is a face (Figure 4). Core samples and cuttings are
function of its capillary pressure curve (i.e., pore throat extracted to remove all hydrocarbons and are dried at
size distribution) and the density of the hydrocarbons approximately 60°C for at least 24 hr; the dry weight of
and pore water. For other oil, gas, and water densities, the samples is then measured. The sides of the conven-
the hydrocarbon column held can be determined by tional core plug are coated with epoxy so that mercury
the equations in Appendix 1. can enter only at the top and bottom of the plug.

Figure 2. Schematic of a
stratigraphic trap showing
TOP SEAL LATERAL the forces (pressures) control-
Pe
SEAL ling hydrocarbon entrap-
ment. Pb = buoyancy
Pe pressure in psi or the driving
Pe force; Pc = capillary pressure
in psi or the resistive force; Pc
at 7.5% nonwetting satura-
tion is assumed to be the seal
leakage saturation; h =
hydrocarbon column height
BOTTOM in feet; σ = interfacial tension
Pb = h ( w - hc) 0.433, psi SEAL between water and hydrocar-
2 cos bon in dynes; θ = contact or
Pc = dynes/cm2
r wetting angle, in degrees; r =
Pe = seal capillary entry pressure (psi) radius of the capillary tube
(or pore throat) in cm.
4 Sneider et al.

APPROXIMATE Figure 3. Mercury/air


MERCURY-AIR OIL-WATER GAS-WATER
Pc = 7.5% Sat. capillary pressure curves
Pc (psi) h (ft) Pc (psi) h (ft)
3000 of “A” and “D” seals.
o/w = 30 dynes/cm g/w = 70 dynes/cm
A Breakthrough pressure or
D leakage of hydrocarbons
HYDROCARBON through the seal is assumed
113 1,189 232 521 COLUMN HELD
1000 A SEAL to be at 7.5% nonwetting
phase saturation. The
500 approximate oil/water
and gas/water capillary
300 pressures and hydrocarbon
column height for oil and
11.4 120 23.5 53
HYDROCARBON
COLUMN HELD
gas are shown.
D SEAL
100

50 OIL GAS

30

A CLAY-RICH SHALE
D SANDY SILTSTONE

10
60 50 40 30 20 10 0
NONWETTING PHASE SATURATION (%)

The core plug or cuttings are sealed in a glass pen- 60,000 psia. Equilibrium at each high-pressure step is
etrometer (Figure 5a) that is placed into a “low- determined as in the low-pressure stage; however, a
pressure” port of a porosimeter. A vacuum of 10–20µ is 60-sec waiting period is used. Computer hardware/
drawn on the rock sample, and the penetrometer is filled software controls, monitors, and records test results.
with mercury at a “filling pressure” of 1.5 psia. This is Data are corrected for any artifacts by applying cali-
called the “low-pressure” mercury injection stage. bration data obtained by running blanks.
In the low-pressure stage, the injection pressure is More than 100 pressure steps or measurements are
increased incrementally over a series of pressure made during both the low- and high-pressure stages.
steps up to 25.0 psia. Equilibrium condition is estab- At the conclusion of each MICP sample analysis, the
lished at each pressure point (step) when mercury test results are printed out for examination and evalu-
intrusion ceases, as indicated by the lack of any pres- ation. A pressure vs. volume of mercury plot is shown
sure drop at the end of a 20-sec waiting period. At the in Figure 5b. These data are placed on a disk for pro-
end of the low-pressure injection stage, the holder cessing and final data reduction.
with the sample is removed from the porosimeter,
weighed, and then loaded into the high-pressure cell.
The injection pressure is returned to both the same
pressure point and respective amount of mercury VERTICAL SIDES OF VERTICAL
intrusion that existed at the last low-pressure point. PLUG CUT PLUG COATED AND
The pressure is then increased incrementally to SEALED WITH EPOXY

Table 1. Seal Types and Hydrocarbon Columns.

Seal-Flow Barrier 35° API Oil Column Held


Type (m) (ft) 1 INCH
A* ≥1500 ≥5000
A ≥300– <1500 ≥1000– <5000
B ≥150– <300 ≥500– <1000
C ≥30– <150 ≥100– <500 3/4 INCH

D ≥15– <30 ≥50– <100


E <15 <50
F Waste Zone Rocks1
Figure 4. Schematic diagram shows orientation of
1Poor-quality, low-permeability rocks that contain appreciable vertical seal plug that is coated with epoxy. Mercury
hydrocarbons. enters the sample only at the top and bottom.
Seal Capacity from Cores vs. Cuttings 5

AUTOPORE POROSIMETER
(PRESSURE VS. VOLUME MEASURED)

1.0
)

0.8
(

0.6

0.4

0.2

0.0
100 50 10 5 1 0.5 0.10.05 0.01
DIAMETER ( m)

a b
PRESSURE
GENERATOR
Figure 5. (a) Penetrometer with sample surrounded with mercury in the porosimeter pressure chamber. (b) Plot
of cumulative pressure vs. mercury intrusion volume. (After Micromeritics, 1995.)

A key aspect for each HPMIC test is the determina- plug approach one another as the pressure increases.
tion of “closure pressure.” Closure pressure is that The difference in capillary pressure (in psia) at Pc =
pressure at which mercury first enters the sample pore 7.5% mercury (nonwetting phase) saturation is defined
space (i.e., initial pore entry pressure) as opposed to as the empirical adjustment factor (EAF). In this exam-
mercury closing or conforming around the sample to ple, 179 psi – 170 psi = 9 psi; 9 psi would be added to
fill sample surface irregularities that are not part of the the capillary pressure value of the cuttings to approxi-
sample pore volume. The more irregular the sample mate the capillary pressure of the epoxy-sealed vertical
surface(s) (i.e., cuttings), the higher the closure. The plug. The EAF is 9 psi.
closure or apparent mercury intrusion must be deter- Figure 7 illustrates the mercury saturation in a ver-
mined from examination of the plots of pressure vs. tical plug in which mercury can enter only at the ends
mercury volume injected. The closure must be sub- (Figure 7A). At the entry capillary pressure of Pe = 0%
tracted prior to the calculation of capillary pressure vs. mercury (Figure 7B), the nonwetting phase begins to
percent of pore space (pore volume) occupied. These enter or surround the pore system at the end of the
calculations were made for all the HPMIC curves mea- plug. As pressure increases, mercury enters the
sured in our studies. pores. At a mercury saturation of about 7.5%, we
Figure 6 is a portion of two capillary pressure curves visualize that one or more continuous filaments or
that have been corrected for closure. One curve is of an pathways of mercury connect from one end of the
epoxy-sealed vertical plug, and the other is of simu- plug to the other (Figure 7C). We believe that this is
lated cuttings prepared from the rock surrounding at or near the breakthrough saturation (when the
where the plug was cut. The capillary entry pressure nonwetting phase leaks through the seal). It is possi-
(Pe) is 26 psi for the simulated cuttings and 91 psi for ble that the breakthrough saturation may vary for
the plug. The pressure vs. mercury saturation values different seals, but 7.5% mercury saturation is consis-
for the simulated cuttings and epoxy-coated vertical tent with our empirical observations.
6 Sneider et al.

200

Pe10% = 191 psi


Pe10% = 187 psi
180
Pe core = 179 psi
7.5% EAF
Pe cuttings = 170 psi
7.5%
160
Empirical Pe5% =
Adjustment Factor (EAF) = 9 psi
161 psi

140

Pe5% = 130 psi


(psi)

120

100

Pe = 91 psi
0%

80
SIMULATED CUTTING

SEALED VERTICAL PLUG


60

40

Pe0% = 26 psi
20
= 7.9%, ka = 0.02 md

0
30 25 20 15 10 5 0
MERCURY SATURATION (% PORE VOLUME)

Figure 6. A portion of high-pressure capillary pressure curves of a sealed


vertical plug and simulated cuttings of the rock from which the plug was cut.
Note the capillary pressures at Pe = 0% (initial mercury entry) and at 5%, 7.5%,
and 10% nonwetting phase saturation. Ka = air permeability.

TYPICAL HIGH-PRESSURE pore space occupied by the nonwetting phase mer-


MERCURY/AIR CAPILLARY cury, the cuttings curves are below the plug curves.
The cuttings have all sides of the sample fragments
INJECTION CURVES exposed, and mercury will first enter the larger pore
throat sizes and the partial pores exposed on the sur-
Figure 8 shows high-pressure mercury/air injec- face of the sample. Although not illustrated in this
tion capillary pressure curves (HPMIC) for four typi- paper, nonsealed horizontal plugs of seals have cap-
cal seals. In each graph, the curve with black squares illary pressure curves very similar to simulated cut-
is a vertical plug epoxy sealed on its sides, and the tings. The HPMIC curves show the entry pressure
other curve (open squares) is of simulated cuttings and the capillary pressure at 5% and 10% nonwetting
from the same piece of rock from which the core plug phase saturation, expressed as pore space occupied
was cut. At lower values of capillary pressure and by mercury. A 7.5% nonwetting phase saturation
Seal Capacity from Cores vs. Cuttings 7

MERCURY INJECTION AT TOP MERCURY INJECTION AT TOP

X Y X Y

SEALED SAMPLE
WITH EPOXY COATING
ON SIDES

MERCURY INJECTION
A MERCURY INJECTION AT BOTTOM B AT BOTTOM
Pc @ 0% Hg Saturation

MERCURY INJECTION

X Y

C Pc @ 7.5% Hg Saturation

Figure 7. Schematic diagram illustrating the distribution of mercury at Pe = 0% and Pc = 7.5% mer-
cury saturation. Mercury is black. Note the continuous filaments of mercury through the sample.
(A) The mercury injection sample. (B) Pc at ±0% mercury (Hg) saturation. Mercury fills and conforms
with the outer grain surfaces. (C) Pc at 7.5% mercury (Hg) saturation. Mercury fills many pores, and
numerous mercury-filled pathways are continuous from the top to the bottom of the plug. Pc at 7.5%
mercury saturation is assumed to be the breakthrough or leakage pressure of the sample. (After
Micromeritics, 1995.)
8 Sneider et al.

Mercury Injection Capillary Pressure (Pore Volume) Mercury Injection Capillary Pressure (Pore Volume)
100000 100000

10000 Plug 10000

Plug

1000 Cuttings 1000

Cuttings

100 Por. = 3.9% 100 Por. = 9.4%


Perm. = 0.005 md. Perm. = 0.031 md.

10 Pc@ Plug Cuttings 10 Pc@ Plug Cuttings

Entry 130 100 Entry 400 160


5% 1600 700 5% 1000 450
10% 3600 2500 10% 1300 800

1 1
100 90 80 70 60 50 40 30 20 10 0 100 90 80 70 60 50 40 30 20 10 0
Pore Space Occupied (%) Pore Space Occupied (%)

"A" SEAL "B" SEAL


Mercury Injection Capillary Pressure (Pore Volume) Mercury Injection Capillary Pressure (Pore Volume)
100000 100000

10000 10000

Cuttings

1000 1000
Plug
Plug

100 Por. = 13.1% Cuttings 100 Por. = 19.0%


Perm. = 0.43 md. Perm. = 0.43 md.

10 Pc@ Plug Cuttings 10 Pc@ Plug Cuttings

Entry 60 25 Entry 30 22
5% 265 122 5% 105 85
10% 300 225 10% 135 125

1 1
100 90 80 70 60 50 40 30 20 10 0 100 90 80 70 60 50 40 30 20 10 0
Pore Space Occupied (%) Pore Space Occupied (%)

"C" SEAL "D" SEAL

Figure 8. Examples of high-pressure mercury/air injection curves for seal types “A”, “B”, “C”, and “D”. Curves
for the vertical plug are designated with black squares; the cuttings curves (open squares) are of simulated cut-
tings from rock adjacent to the vertical plug.
A B

"A" SEAL "B" SEAL

C D

"C" SEAL 10 µ "D" SEAL


1000 X

Figure 9. SEM photomicrograph of seal types “A,” “B,” “C,” and “D” illustrated in Figure 8.
Note the 10µ scale. “A” Seal—Predominant clay fabric with limited grain support. Clay par-
ticles are compacted and have only a slight recrystallized diagenetic appearance.
Intercrystalline pore volume is low. “B” Seal—Rock fabric shows some grain support and
has common intergranular detrital clay. Clay particles are generally compacted and exhibit
limited diagenetic character. Minor intercrystalline pore space, largely concentrated in inter-
granular areas where clay is less compacted. “C” Seal—Rock fabric shows grain support.
Intergranular clay particles have a more random orientation, which corresponds to an
increase in intercrystalline pore volume. Clay morphology shows more pronounced diage-
netic character. “D” Seal—Grain-supported fabric with development of quartz overgrowth
cement. Clay minerals consist of diagenetic grain-coating/pore-filling chlorite and kaolinite
with common intercrystalline pore space. Partial preservation of intergranular pores, with
apertures restricted by diagenetic phases.
10 Sneider et al.

Table 2. Empirical Adjustment Factors (EAFs) to Estimate Air/Mercury (Air/Hg) Capillary Pressures of
Vertical, Epoxy-Coated Plugs from Air/Mercury Capillary Pressure of Cuttings.

EAF Values Added to Cuttings Air/Hg Capillary


Pressure (psia) at 7.5% Mercury Saturation
Number
Seal Type of Samples Average Min. Max.
A* 6 2315 1402 3120
A 72 1810 923 4009
B 79 455 423 1040
C 48 140 22 363
D 27 30 27 91

correlates best with the hydrocarbon column heights EMPIRICAL ADJUSTMENT FACTORS
(HCH) found in reservoirs.
The air permeabilities of the seals measured under The empirical adjustment factor (EAF) is the differ-
stress are low to very low. Scanning electron micro- ence expressed in pressure (psi) between the capillary
scope (SEM) photomicrographs (Figure 9) show the pressure at a specific mercury saturation measured on a
poor interconnection of pores in the “A” and “B” seals. vertical plug of a seal cut perpendicular to the sealing
For the “C” and “D” seals, pore size and interconnec- surface (i.e., usually a vertical plug cut perpendicular to
tion increase. a bedding surface) and that measured on simulated

1554 Figure 10. Mercury/air and


"D" oil/water capillary pressure
N = 905 SEALS 1564
140 14 curves of reservoir and seal
rocks from the Tar Springs
1574 Formation, Benton Field,
Illinois. The field column
120 12 1584 height is about 90 ft, which
corresponds to the oil
1594 column held by the weakest
7.5% “D” seal. Leakage is
100 10 1604 assumed to be at 7.5% oil
(psi)

(psi)

saturation. IAA =
1614 >1000–5200 md, well-sorted,
lower medium-grained
80 8 1624 sandstone; IA = 200–1000
IC, ID-II md, very well sorted, upper
1634 very fine grained sand-
stone; IA-B = 50–200 md,
60 6 IB 1644 well-sorted, lower and
upper very fine grained
1654 sandstone; IB = 10–50 md,
IA-B moderate to well sorted,
40 4 1664 lower very fine grained
sandstone and siltstone;
IA 1674 and IC, ID-II = 0.1–10 md,
moderately sorted siltstone.
20 2 IAA 1684 OOWC = original oil/water
contact.
1694
FIELD OOWC 1700
0 0 1704
100 80 60 40 20 0
MERCURY AND OIL SATURATION (% PORE VOLUME)
Seal Capacity from Cores vs. Cuttings 11

cuttings of the same seal rock type. The saturation val- reviewer for their editorial comments. We owe special
ues chosen for EAF usually are determined at 5%, 7.5%, thanks to Florence Rollins for drafting assistance and
or 10% nonwetting phase saturation. In this paper, 7.5% to Ramona Sneider for preparing the manuscript.
nonwetting phase (mercury) saturation is used.
From hundreds of pairs of HPMIC curves like those
in Figure 8, EAFs are derived by averaging pressure val- APPENDIX 1
ues between the plug and cuttings values determined at Equations to Convert Mercury/Air Capillary
7.5% nonwetting phase saturation. The EAFs, in psi, are Pressure to Oil/Water and Gas/Water Capillary
the average values that need to be added to an HPMIC Pressure and HCH
curve of cuttings to approximate the capillary pressure
that would be measured on the vertical plug in which The hydrocarbon column held is a function of the
the sides are coated and sealed with epoxy. We have buoyancy pressure, generated by the difference
standardized on deriving the EAFs at 7.5% mercury sat- between the hydrocarbon and water density, necessary
uration. This is based on empirical data that show an to overcome the capillary pressure of the seal rock at a
equivalence between column heights held in reservoirs given nonwetting phase saturation (e.g., 0%, 5%, 7.5%).
and the estimated seal capacity derived from HPMIC The following equation illustrates this relationship:
data of the capping seals.
The most up-to-date data sets on EAFs are shown in
Table 2. These data are based on more than 230 seals. Pc = h( ρ w – ρ h )0.433
h/w (psi)
As we continually add additional pairs of vertical
plugs and cuttings or simulated cuttings, the EAFs Pc (1)
h (ft) = h/w
might be modified, but we expect that future values ( ρ w – ρ h )0.433
will not be significantly changed.
Figure 10 shows mercury/air capillary pressure
curves based on 905 samples of reservoirs and seals The following values are commonly used for the
from the Lower Carboniferous sandstone in the Benton above calculation: water density (ρw) = 1.110 g/cc; oil
Field in Illinois. The trap is a simple, four-way closed density (ρh) = 0.8498 g/cc (35° API); and gas density
anticline. The 24 seals measured are “D” type. The mer- (ρh) = 0.050 g/cc.
cury/air capillary pressure curves of the reservoir and The calculation of seal capacity and hydrocarbon
seals are converted to the oil/water system, and then the column held for a given rock type is based on the
height of hydrocarbon column held by the weakest “D” air/mercury (a/Hg) capillary pressure data and its
seal is calculated. The field HCH of ~90 ft agrees closely conversion to an equivalent hydrocarbon/water
with the hydrocarbon column predicted for the seal (h/w) capillary pressure system using the following
capacity using the capillary pressure value at 7.5% non- equations:
wetting phase saturation.
2σ h/w cos θ h/w
CONCLUSIONS Pc
r σ h/w cos θ h/w
h/w
= =
Pc 2σ a/Hg cos θ a/Hg σ a/Hg cos θ a/Hg
1. Vertical plugs cut perpendicular to sealing sur- a/Hg
r
faces and epoxy-coated on the sides are the most
reliable sample type to obtain high-pressure mer-
cury/air capillary pressure curves on seals. and (2)
2. High-pressure mercury injection curves (HPMIC)
of cuttings from seals at low mercury saturation
values give capillary pressure curves whose pres-  σ h/w cos θ h/w 
Pc = Pc σ 
sure values for seal capacity (i.e., hydrocarbon col- h/w a/Hg
 a/Hg cos θ a/Hg 
umn heights held) are lower than those of the
vertical plugs. where a/Hg = air/mercury/solid/system, h/w =
3. Empirical adjustment factors (EAFs) added to the hydrocarbon/water/solid/system, o/w = oil/water/
capillary pressure curves of the cuttings can be solid system, and g/w = gas/water/solid system.
used to approximate the capillary pressure of The following values are commonly used air/mer-
vertical plugs. cury contact angle (θ ) = 140°; oil/water contact angle
4. The EAF values are picked at a nonwetting phase (θ ) = 0°; gas/water contact angle (θ ) = 0°; air/mer-
saturation of 7.5%. The capillary pressure at 7.5% cury interfacial tension ( σ ) = 480 dynes/cm;
saturation corresponds best with the heights of oil/water interfacial tension (σ) = 30 dynes/cm; and
hydrocarbon columns measured in many fields. gas/water surface tension (σ) = 70 dynes/cm. These
values are for surface or near-surface conditions.
ACKNOWLEDGMENTS When using equations 1 and 2 for reservoir condi-
tions, the values of fluid densities and interfacial ten-
The authors are grateful to Dr. C.L. Vavra of ARCO sions must be corrected for reservoir temperature
E & P Technology, Plano, Texas, and an unknown and pressure.
12 Sneider et al.

Incorporating the values into equation 2 yields the These equations are used to calculate the HCH by
following conversions: substituting the appropriate air/mercury capillary
pressures at 0%, 5%, 7.5%, or 10% mercury saturation.
Pc o/w = 0.082 Pc a/Hg
Pc g/w = 0.190 Pc a/Hg (3)
REFERENCES CITED

By incorporating the values for water, oil, and gas Micromeritics, 1995, Operators manual AutoPore III:
density into equations 1 and 3, the hydrocarbon col- Norcross, Georgia, 258 p.
umn heights (h) for oil and gas yields: Schowalter, T.T., 1979, Mechanics of secondary hydro-
carbon migration and entrapment: AAPG Bulletin,
v. 63, no. 5, p. 723–760.
For oil : h(ft) = 0.728 Pc a/Hg Sneider, R.M., K.K. Stolper, and J.S. Sneider, 1991,
Petrophysical properties of seals (abs.): AAPG Bul-
For gas : h(ft) = 0.414 Pc a/Hg (4)
letin, v. 75, no. 3, p. 673–674.
Noble, R.A., J.G. Kaldi, and C.D. Atkinson, 1997, Oil
saturation in shales: applications in seal evalua-
tion, in R.C. Surdam, ed., Seals, traps, and the
petroleum system: AAPG Memoir 67, p. 13–29.

Chapter 2

Oil Saturation in Shales:


Applications in Seal Evaluation
R.A. Noble
J.G. Kaldi
Atlantic Richfield Indonesia Inc.
Jakarta, Indonesia

C.D. Atkinson
ARCO British Ltd.
Guildford, Surrey, United Kingdom

ABSTRACT
A procedure has been developed to quantify oil saturation in the pore sys-
tem of shales. The technique uses geochemical and rock property measure-
ments of core samples (solvent extract yield, porosity, densities, and kerogen
sorption capacities). The method takes into account the fact that many shales
contain indigenous organic matter (kerogen) and that free hydrocarbons
extracted from the shale may originate either from the sorbed fraction of the
kerogen/mineral matrix or from residual hydrocarbons within the intergran-
ular pore system. A study of the Eagleford Formation from east Texas shows
that mineral surfaces of the shale most likely remain water wet and that
residual oil saturation (So) of the intergranular pore system attains the high-
est values during the intense zone of oil generation (calculated So = 15 to
70%). The saturation values are examined as a function of burial depth and
organic richness to establish typical trends for shales undergoing normal
maturation. Relationships between pore saturations and Rock-Eval S1/TOC
(total organic carbon) ratio are established so that the concepts can be
applied in cases where only Rock-Eval data are available. Samples with
S1/TOC ratios >120 mgHC/gC may contain some nonindigenous hydrocar-
bons, and those with values >200 mgHC/gC almost certainly do. These val-
ues were used to evaluate the residual oil contents and seal performance of
various fine-grained rock facies. A case study from the Talang Akar
Formation, Indonesia, shows that seal rocks with high entry pressures (from
mercury injection capillary pressure [MICP] analysis) have low hydrocarbon
contents in the range expected for in-situ generation. However, shales with
the lowest entry pressures have very high hydrocarbon (HC) contents, indi-
cating impregnation of the pore system with oil from an underlying accumu-
lation. In such samples, the seal rock has most probably attained equilibrium
with the maximum oil column height it was capable of supporting. The

13
14 Noble et al.

method complements existing mercury injection capillary pressure (MICP)


measurements of seal capacity, and provides a rapid means for detecting seal
failure or poor-quality reservoir “waste zones.”

INTRODUCTION EXPERIMENTAL
Laboratory procedures for measuring oil satura- Samples
tion in coarse-grained reservoir rocks are well estab-
lished and have been used for many years in the The two sets of samples investigated in this study
petroleum industry. The most commonly used proce- are from the Eagleford Formation of east Texas and the
dure involves either Dean-Stark extraction of resid- Talang Akar Formation of offshore northwest Java
ual pore fluids or high-temperature vacuum retorting (ONWJ), Indonesia.
(e.g., Koepf, 1978). Neither of these methods is suit- Core samples of Upper Cretaceous Eagleford shale
able for the analysis of shales and other fine-grained from seven wells in east Texas and surrounding areas
sedimentary rocks. Dean-Stark extraction of intact were studied in detail for the purpose of determining
core plugs requires a high-permeability rock matrix oil saturation values. The marine Eagleford Forma-
for solvent access to the pore system, which is typi- tion (Woodbine Group) is an important source rock
cally not the case for shales. Vacuum-retorting meth- in the region of the Sabine Uplift in east Texas and
ods are unsuitable for shales that contain organic Louisiana (Halbouty and Halbouty, 1982; Westcott
matter (kerogen) because this material breaks down and Hood, 1994). The lithologic character of the
to yield hydrocarbons and other products under ther- Eagleford samples is quite variable, ranging from
mal stress. Release of bound water from clay miner- siltstones to limestones to true shales. The cored
als may also complicate the assessment of pore fluid intervals of most interest consisted of dark gray to
saturation. black shales, which were typically well laminated
Despite these analytical difficulties, there are com- and showed no evidence of bioturbation. However,
pelling reasons for acquiring oil saturation data in other facies including siltstones, limestones, and one
nonreservoir rocks. Shales and other fine-grained sample from a sandy reservoir interval (Saum #1,
rocks are the most abundant constituents of sedimen- 9733 ft) were also analyzed. Clay minerals largely
tary basins, often comprising 70–75% of all sediments consisted of kaolinite, chlorite, and, to a lesser
(Gorsline, 1984). Many fine-grained rocks serve as degree, illite (from unpublished XRD data). Visible
source rocks for hydrocarbon accumulations, and oth- porosity was not apparent from thin-section pho-
ers may act as seals and barriers to hydrocarbon flow. tomicrographs of shales.
Understanding saturation profiles in source rocks can Conventional and sidewall cores from wells in the
assist in the interpretation of petroleum expulsion Ardjuna Basin of ONWJ were used to study geochem-
efficiency and timing, which continue to be major ical properties of potential seal facies. The Talang Akar
uncertainties in hydrocarbon charge assessment. Formation consists of a deltaic to marginal marine
Quantifying saturation levels in sealing lithologies sequence of Oligocene age. Samples analyzed in this
can aid in the evaluation of trapping mechanisms and study included delta plain shales, channel abandon-
reserves assessment and can assist in the postmortem ment siltstones, delta front shales, prodelta shales, and
analysis of dry holes. shelf carbonates. Further details of sample type, loca-
In this study, residual oil saturation (So) in shales is tion, and rock properties are given in the accompany-
addressed. The study is divided into two parts. The ing paper by Kaldi and Atkinson (1997).
first examines residual So in the Eagleford Formation
Geochemical Analyses
of east Texas. These organic-rich shales serve as
source rocks for oil and provide an opportunity to A total of 31 Eagleford and 10 Talang Akar samples
examine oil saturation levels in fine-grained rocks were analyzed for geochemical properties. Total
undergoing normal maturation. Relationships organic carbon (TOC) and Rock-Eval data were
between the core-derived saturations and the Rock- acquired for all samples using standard analytical
Eval S1 peak are evaluated. In the second part, a case methods. In addition, Soxhlet extraction and liquid
study from Indonesia is presented. Potential seals chromatography data were collected for the Eagleford
with very different capillary entry pressures are ana- samples. Special consideration was given to acquiring
lyzed, and the amount of free hydrocarbons is mea- very accurate and reproducible results for the
sured in each. The information derived from this extractable organic matter (EOM) using Soxhlet
assessment is used in the evaluation of seal perfor- reflux. The major source of error in this procedure is
mance, which has important applications in explo- associated with obtaining a weight-stabilized fraction
ration and development programs. for gravimetric analysis. The amount of material lost
Oil Saturation in Shales 15

during the solvent “blow-down” stage varies with the Figure 1. An optimum time for evaporation was deter-
composition of the EOM and the duration/tempera- mined to be 80 min (nitrogen stream at 40°C), which
ture used for light-end evaporation. (A discussion of represents a compromise between practicality and a
these variables is presented under “Extractable totally weight-stabilized product. After 80 min. of
Organic Matter.”) evaporation, the measured weight of material was
estimated to be within 5% of the true weight-stabilized
Rock Property Measurements fraction, which was sufficiently accurate for the pur-
poses of this study. This workup procedure was
Core plugs (1 in. [2.5 cm] diameter) were taken at
adopted for all Eagleford shale extracts (Table 2).
sampling points as close as possible to the geochemical
The EOM value determined in this way corre-
samples. The plugs were analyzed for helium porosity
sponds to the C12+ weight fraction of the residual oil
(Boyle’s Law), grain density, and mercury injection (based on gas chromatography analysis). Hence, cor-
capillary pressure (MICP) (0–60,000 psi). The helium rections must be made for loss of light ends (C12– frac-
porosity values were compared with the mercury tion) to give the total oil extract or corrected EOM. The
injection data and were found, more often than not, to amount of light ends in a sample is primarily a func-
be larger, due to greater helium access to smaller tion of its thermal maturity. Organic matter type may
pores. The helium values were used in all calculations exert a secondary influence, although this is relatively
involving porosity. unimportant in geographically extensive marine
The results of geochemical and rock property mea- shales such as the Eagleford samples used in this
surements are presented in Tables 1–3. study. Correlations have been established between the
level of thermal maturity and the density of crude oils
DETERMINATION OF OIL SATURATION (e.g., Engel et al., 1988; Sofer, 1988). Crude oil density
IN SHALES (or API gravity), in turn, is dependent on the amount
of light ends in the sample, as shown in Figure 2. By
Extractable Organic Matter determining the relative amounts of C 12– and C12+ in
crude oils of different gravities, it is possible to correct
Methods developed to measure oil saturation in the EOM yields at various levels of thermal maturity
coarse-grained reservoir rocks are typically not suit- for light-end loss (Table 2). Note that the C12– fraction
able for determining oil saturation in shales. actually corresponds to a carbon number range of
For instance, the widely used Dean-Stark solvent- ~C 5 –C 12 , because the evaporation experiments
extraction procedure for intact core samples (e.g., described above used stock tank oils that had already
Koepf, 1978) is not suitable for low-permeability been separated from their solution gas (C1–C4). The
shales. Similarly, vacuum-retorting methods are inap- corrected EOM values shown in Table 2 correspond to
propriate for shales with indigenous organic matter the yield of extractable liquid hydrocarbons (C5+) at
(kerogen) because the kerogen breaks down under surface conditions.
thermal exposure to yield oil-like hydrocarbons.
It was therefore necessary to use a more appropri-
Occurrence of Residual Oil in Organic-Rich Shales
ate procedure to quantify the residual oil saturation
(S o ) in shale samples. Geochemical techniques are There are three principal locations in which oil may
available to extract the total quantity of residual oil reside in a shale: (1) within the intergranular pore sys-
from the shale. This material is commonly referred to tem; (2) adsorbed and/or absorbed in the kerogen
as bitumen or extractable organic matter (EOM), and is matrix (kerogen sorption); (3) in the case of oil-wet or
obtained by crushing the rock to a fine powder, fol- mixed wettability pore systems, adsorbed on mineral
lowed by continuous solvent extraction under reflux- grain surfaces (mineral sorption). The extraction
ing conditions (Soxhlet technique). The solvent is method does not discriminate between these different
removed by distillation, which also removes light sites of oil retention. When referring to “oil satura-
hydrocarbons from the sample. The precise laboratory tion,” the quantity we are interested in is the amount
conditions used for solvent removal (i.e., temperature of oil remaining in the pore system, so the question
and time) have a significant influence on the carbon then is how to partition the total EOM into its various
number range of residual hydrocarbons, and these genetic fractions. This is achieved by first accounting
conditions may vary from one laboratory to another. for the amount of oil that is retained by sorption, and
Hence, EOM yields are frequently reported either as then subtracting this amount from the total oil extract.
C10+, C12+, or C15+, depending on the methods used. In the following discussion, the term “sorption” will
For our purposes, a standardized solvent removal be used, since it is generally not possible to distinguish
procedure was established from a series of experi- between adsorption and absorption processes in
ments involving crude oils of different gravities (26°, kerogenous rocks.
35°, 44°, and 58° API). In each case, a known weight of There has been much interest in recent years in
oil (200–750 mg) was mixed with dichloromethane (20 studying the hydrocarbon sorption capacity of solid
mL), and the samples were evaporated using a Roto- organic matrices. Much of this work has been driven
vap, followed by nitrogen blow-down at 40°C. The by the gas industry in the context of coal-bed methane
weight of residual material relative to the starting and fractured shale gas projects (Levine, 1993; Yee et
weight of oil is plotted against evaporation time on al., 1993). Langmuir-type sorption isotherms have
16

Table 1. TOC, Rock-Eval, and Solvent Extract Data for Eagleford Samples.

DEPTH TOC* S1 S2 S3 HI OI Tmax EOM Saturates Aromatics NSO Asphaltenes


WELL (ft) (wt %) (mgHC/gRk) (mgHC/gRk) (mgCO2/gRk) (mgHC/gC) (mgCO2/gC) PI (°C) (mgHC/gRk) (wt. %) (wt. %) (wt. %) (wt. %)

HILLTOP RESORT #2 6743 3.03 1.40 11.78 0.42 389 14 0.11 436 3.55 37.1 19.4 29.3 9.7
HILLTOP RESORT #2 6750.5 3.32 1.63 14.25 0.57 429 17 0.10 438 4.23 54.6 18.1 22.1 0.2
HILLTOP RESORT #2 6782 3.77 2.12 15.81 0.69 419 18 0.12 434 3.08 49.7 20.2 20.5 5.0
Noble et al.

HILLTOP RESORT #2 6809 3.75 2.25 14.64 0.65 390 17 0.13 437 4.42 50.0 17.8 22.0 5.6

FEE 183 #1 8944 0.29 nd** nd nd nd nd nd nd 0.23 57.9 26.3 7.9 5.3
FEE 183 #1 8961.1 3.1 1.00 8.69 0.45 280 15 0.10 444 3.10 41.3 26.1 18.5 9.8
FEE 183 #1 8967.5 0.3 nd nd nd nd nd nd nd nd nd nd nd nd
FEE 183 #1 8979 0.2 nd nd nd nd nd nd nd 0.28 41.7 20.8 29.2 4.2

SAUM #1 9760 2.49 2.27 4.19 0.41 168 17 0.35 449 4.41 74.7 12.1 8.4 0.0
SAUM #1 9762 2.62 2.22 4.39 0.47 168 18 0.34 452 4.42 68.3 15.1 7.9 1.9
SAUM #1 9766 1.33 0.73 1.17 0.50 88 38 0.38 449 1.77 67.1 12.8 9.6 6.1
SAUM #1 9773 1.3 4.62 1.40 0.83 108 64 0.77 460 6.94 78.2 9.3 6.0 1.6
SAUM #1 9790 0.67 0.19 0.39 0.56 58 84 0.33 462 0.22 62.8 11.9 16.8 4.0

KIRBY #1 10885 0.43 nd nd nd nd nd nd nd 0.63 66.1 12.5 14.3 2.7


KIRBY #1 10898 0.3 nd nd nd nd nd nd nd 0.07 29.2 62.5 2.1 2.1
KIRBY #1 10907.5 0.09 nd nd nd nd nd nd nd 0.02 45.8 41.7 4.2 4.2
KIRBY #1 10933 0.29 nd nd nd nd nd nd nd 0.38 54.5 28.0 8.0 5.0

SOUTHLAND #2 12375.5 0.37 nd nd nd nd nd nd nd 0.12 46.5 32.6 14.0


2.3
SOUTHLAND #2 12390.5 0.51 0.03 0.16 0.12 31 24 0.16 488 0.10 42.9 19.0 19.0
16.7
SOUTHLAND #2 12434 0.57 0.03 0.14 0.12 25 21 0.18 487 0.05 24.4 14.6 24.4
34.1
SOUTHLAND #2 12439.4 0.51 0.02 0.15 1.32 29 259 0.12 473 0.11 19.4 32.3 11.3
33.9

TODD #1 12967 1.79 2.07 7.79 0.55 435 31 0.21 438 3.36 50.8 19.1 19.7 5.7
TODD #1 12991 2.7 2.12 16.74 0.60 620 22 0.11 437 3.67 46.3 24.4 18.2 6.5
TODD #1 12996 1.43 2.65 5.19 0.63 363 44 0.34 439 3.94 43.8 24.9 22.0 4.8
TODD #1 13004.2 1.22 0.52 4.37 0.62 358 51 0.11 436 0.22 50.3 21.9 21.9 1.1
TODD #1 13013 2.61 0.31 13.15 0.52 504 20 0.02 434 1.34 35.8 31.3 27.4 0.5

MUSSER DAVIS # 1 15911 1.14 0.56 0.33 0.42 29 37 0.63 448 0.98 81.1 8.2 6.0 0.0
MUSSER DAVIS # 1 16591 0.87 0.11 0.26 0.33 30 38 0.30 507 0.43 81.9 5.5 7.9 0.0
MUSSER DAVIS # 1 16592 1.61 0.09 0.31 0.25 19 16 0.22 509 0.38 42.3 20.4 15.5 18.3
MUSSER DAVIS # 1 16605.5 0.87 0.03 0.08 0.20 9 23 0.27 513 0.23 65.0 12.7 11.5 6.4
MUSSER DAVIS # 1 16630 1.01 0.09 0.21 0.19 21 19 0.30 516 0.33 57.1 22.6 7.1 9.5

*TOC = Total organic carbon (gC/gRock); HI = Hydrogen index (S2/TOC); OI = Oxygen index (S3/TOC); HC = Hydrocarbon content (S1/TOC); PI = Production index (S1/S1+S2).
**nd = no data
Table 2. Calculated Oil Saturation Values for Eagleford Samples.*

DEPTH TOC C12- Corr. EOM Sorption Cap. Absorbed HC He Porosity Hg Porosity Rock density Saturation
WELL (ft) (wt %) (Fraction) (mgHC/gRk) (mgHC/gC) (mgHC/gRk) (%) (%) (g/ml) (%So)
HILLTOP RESORT #2 6743 3.03 0.15 4.18 60.0 1.8 4.5 5.0 2.58 16.1
HILLTOP RESORT #2 6750.5 3.32 0.15 4.98 60.0 2.0 6.7 5.1 2.63 13.8
HILLTOP RESORT #2 6782 3.77 0.15 3.62 60.0 2.3 4.6 4.5 2.70 9.4
HILLTOP RESORT #2 6809 3.75 0.15 5.20 60.0 2.3 7.5 5.5 2.65 12.2

FEE 183 #1 8944 0.29 0.35 0.35 60.0 0.2 3.6 0.6 2.73 1.6
FEE 183 #1 8961.1 3.10 0.35 4.77 60.0 1.9 2.9 1.7 2.72 32.7
FEE 183 #1 8967.5 0.30 0.35 nd 60.0 0.2 2.7 0.9 2.67 nd
FEE 183 #1 8979 0.20 0.35 0.43 60.0 0.1 3.6 2.8 2.68 2.7

SAUM #1 9760 2.49 0.65 6.78 60.0 1.5 5.5 3.8 2.65 30.1
SAUM #1 9762 2.62 0.65 6.80 60.0 1.6 3.9 4.1 2.61 41.5
SAUM #1 9766 1.33 0.65 2.72 60.0 0.8 4.6 3.2 2.89 14.1
SAUM #1 9773 1.30 0.65 10.68 60.0 0.8 15.2 12.9 3.28 25.1
SAUM #1 9790 0.67 0.65 0.34 60.0 0.4 4.7 6.0 2.66 0.0

KIRBY #1 10885 0.43 0.50 1.26 60.0 0.3 6.0 20.0 2.85 5.6
KIRBY #1 10898 0.30 0.50 0.14 60.0 0.2 nd 0.3 nd 0.0
KIRBY #1 10907.5 0.09 0.50 0.04 60.0 0.1 4.1 0.4 2.81 0.0
KIRBY #1 10933 0.29 0.50 0.76 60.0 0.2 1.7 2.74 10.8

SOUTHLAND #2 12375.5 0.37 0.50 0.24 60.0 0.2 8.0 4.7 2.76 0.1
SOUTHLAND #2 12390.5 0.51 0.50 0.20 60.0 0.3 9.3 4.7 2.87 0.0
SOUTHLAND #2 12434 0.57 0.50 0.10 60.0 0.3 4.5 5.0 2.70 0.0
SOUTHLAND #2 12439.4 0.51 0.50 0.22 60.0 0.3 0.7 0.2 2.70 0.0

TODD #1 12967 1.79 0.50 6.72 60.0 1.1 2.5 0.6 2.63 69.0
TODD #1 12991 2.70 0.50 7.34 60.0 1.6 nd nd nd nd
TODD #1 12996 1.43 0.50 7.88 60.0 0.9 nd nd nd nd
TODD #1 13004.2 1.22 0.50 0.44 60.0 0.7 nd nd nd nd
TODD #1 13013 2.61 0.50 2.68 60.0 1.6 5.2 4.1 2.71 6.8

MUSSER DAVIS # 1 15911 1.14 0.80 4.90 60.0 0.7 nd nd nd nd


MUSSER DAVIS # 1 16591 0.87 0.80 2.15 60.0 0.5 9.6 9.9 2.69 5.4
MUSSER DAVIS # 1 16592 1.61 0.80 1.90 60.0 1.0 7.8 7.5 2.70 3.8
MUSSER DAVIS # 1 16605.5 0.87 0.80 1.15 60.0 0.5 8.5 9.0 2.73 2.4
MUSSER DAVIS # 1 16630 1.01 0.80 1.65 60.0 0.6 10.4 9.3 2.73 3.2
Oil Saturation in Shales

*Corr. EOM = EOM/(1–C12–fraction); C12–fraction from Figure 2; oil density = 0.85 g/ml; saturation calculated using equation 2 (see text).
17
18 Noble et al.

been determined for various gaseous components in


coals to estimate storage capacity and gas reserves.
Tmax
(°C)

432
430
436
436
433
436
433
434
422
432
The liquid hydrocarbon retention capacity of coals is
much more complex due to the elaborate microstruc-
ture and internal bonding heterogeneity of the coal
0.36
0.40
0.23
0.24
0.23
0.34
0.38
0.17
0.33
0.36

*TOC = Total organic carbon (gC/gRock); HI = Hydrogen index (S2/TOC); OI = Oxygen index (S3/TOC); HC = Hydrocarbon content (S1/TOC); PI = Production index (S1/S1+S2).
(wt %) (mgHC/gRk) (mgHC/gRk) (mgHC/gRk) (mgHC/gC) (mgCO2/gC) (mgHC/gC) PI

matrix. In fact, the extraction procedure and solvent


system used for this purpose have a large influence on
the ultimate yield of liquid hydrocarbons. Brown
HC

(1993) suggested using molar sorption capacities


228
241
45
37
45
82
107
25
41
45
determined from methane isotherms to calculate the
oil sorption capacity of coals. For oils with average
molecular weights of 210–280 g/mol, an oil-prone coal
(50% vitrinite, 50% other macerals, predominantly
liptinite) would retain about 50 to 70 mgHC/gC by sorp-
16
33
61
216
56
18
36
88
212
55
OI

tion processes (~3–5% by weight of coal). Experi-


mentally determined EOM yields for coal samples
depend on the solvent system and extraction method,
and can range from 10 to 100 mgHC/gC for coals of
405
363
151
117
151
160
178
124
82
80

subbituminous rank (Radke et al., 1980; Boudou, 1984;


HI

Durand et al., 1989; Larsen et al., 1991; Littke and


Leythaeuser, 1993).
The capacity of other types of kerogen to retain oil
has also been examined by using chemical and geo-
chemical methods. Sandvik et al. (1992) reported sorp-
0.56
0.55
0.50
1.92
0.99
0.40
1.95
0.77
0.36
0.11
S3

tion levels generally in the range of 3 to 18% by weight


of solid organic mass, with averages ~6–9% for dis-
persed kerogen. Note that these values have been cor-
rected for light-end loss. Different chemical affinities
for absorption were also reported, with polar com-
pounds sorbing more readily than aromatics, which in
14.1
6.14
1.24
1.04
2.67
3.59
9.71
1.09
0.14
0.16
S2

turn are sorbed preferentially to isoparaffins and nor-


mal paraffins (saturates). Detailed quantitative infor-
mation is presently not available in the public domain
to demonstrate the different sorption behavior of
Table 3. TOC and Rock-Eval Data for Talang Akar “Seal” Samples.*

source rock kerogens as a function of kerogen type


7.93
4.07
0.37
0.33
0.79
1.84
5.83
0.22
0.07
0.09

and maturity. However, from coal studies, it is appar-


S1

ent that vitrinitic and inertinitic coals have a greater


capacity for sorption than liptinite-rich coals, and that
sorption capacity increases with coal rank (Levine,
1993). Our present estimates are that oil-prone kero-
TOC

3.48
1.69
0.82
0.89
1.77
2.25
5.45
0.88
0.17
0.20

gens (types I and II) in the oil window will retain on


average ~50–60 mgHC/gC, whereas gas-prone kero-
gens (type III) have the capacity to retain ~90–100
Channel aband. siltst.
Channel aband. siltst.

mgHC/gC. These values will require further refine-


Talang Akar Facies

ment as our understanding of kerogen retention prop-


Shelfal carbonate
Shelfal carbonate
Delta plain shale
Delta plain shale

Delta front shale


Delta front shale

erties evolves.
Prodelta shale
Prodelta shale

The adsorption of organic compounds on mineral


surfaces is well known and is particularly important
for clays (Anderson, 1986). The clays show different
affinities for sorption of organic substrates, depend-
ing on the polarity and functionality of the latter.
Crude oil components, such as asphaltenes and NSO
compounds (polars), are sorbed most readily and are
considered to be the primary agents for altering the
Measured
Depth (ft)

wettability of petroleum reservoirs (Collins and Mel-


7897.0
7898.0
7264.9
7270.3
6970.8
8058.3
7892.2
6980.6
6983.6
8031.0

rose, 1983; Cuiec, 1984). In the case of kerogenous


shales, the wettability is still a matter of conjecture. It
seems reasonable that the kerogen laminae within the
shale are oil wet, since kerogen is a mixture of organic
BZZA-7
BZZA-7

BZZA-7

KLN-1

polymers with significant affinity for oil. However,


BZ-2
BZ-2
BZ-2

BZ-2
BZ-2
Well

BS-1

there is very little field or laboratory evidence that


clearly shows whether the mineral grains change
Oil Saturation in Shales 19

Evaporation Time (min) Figure 1. Plot of weight loss


vs. evaporation time for
0 20 40 60 80 100 120 140 160 180
crude oils of different API
0 gravity.

10
API = 26
20
Weight Loss (%)

30
API = 35
40
API = 44
50

60
API = 58
70

wettability during the maturation process. On the adsorbed, but clearly does not prevent the adsorption
other hand, data have been presented for the Kim- of polar organic molecules. Collins and Melrose (1983)
meridge shale, U.K., which was interpreted as clear found that pre-adsorption of 6 wt. % water resulted in
evidence of a water-wet intergranular pore system 12 mg/g organic adsorption on kaolinite. Once
throughout the zone of intense oil generation adsorbed, the asphaltenes form a stable complex with
(Mackenzie et al., 1987). In a water-wet pore system, the clays. Only extraction with powerful organic sol-
the amount of petroleum sorbed onto mineral surfaces vents (such as the Soxhlet technique used in this work)
is minimal and can be ignored in the calculations of So. can remove the asphaltenes from mineral surfaces,
However, if the minerals are involved in oil retention, and even then incomplete recovery is reported
the amount of hydrocarbons sorbed per unit weight of (Clementz, 1976).
rock must be taken into account. We present some The EOM data for the Eagleford shales were stud-
quantitative estimates to address this issue. ied in the context of evaluating the role of mineral
Laboratory adsorption studies using pure minerals adsorption. An assumption is made that the mineral
show that clays have the capacity to retain significant surfaces become either completely or fractionally oil
amounts of “heavy-end” petroleum compounds (i.e., wet if sufficient resins and asphaltenes are adsorbed.
resins and asphaltenes). Under anhydrous conditions, Using the experimental data of Collins and Melrose
minerals such as kaolinite, chlorite, montmorillonite, (1983), this threshold amount is estimated to be about
and illite typically adsorb 20–30 mg/g (Clementz, 10 mg/g clay in an aqueous environment. Consider
1976; Collins and Melrose, 1983). The presence of the Saum #1 samples, which are presently in the oil
water interferes, to some extent, with the amount window and afford some of the highest EOM yields

100 Figure 2. Plot of API


90 gravity and corresponding
80 maturity level vs. weight
of C12– in crude oils. The
Weight Percent C12-

70
relationship of oil and
60 gravity and maturity was
50 estimated from data
40 presented by Engel et al.
(1988) and Sofer (1988).
30

20
10

0
API Gravity: 10 20 30 40 50 60 70

Maturity Level: Low Moderate Peak Late


20 Noble et al.

(Table 1). Our unpublished XRD data show average Oil Saturation (%)
clay contents of about 20%. The dominant clay miner-
als are chlorite and kaolinite, both of which have simi- 0 10 20 30 40 50 60 70 80 90 100
lar surface adsorption properties. Using the estimated
adsorption threshold of 10 mg polar organics per gram
of clay and 20% clay/rock, then 2 mg/g rock is poten- TOC below 1 %
tially required to create an oil-wet clay surface. Other 4000 TOC above 1 %
minerals in the rock, such as calcite and quartz, have
lower adsorption capacities but will still retain some
material, perhaps another 1 mg/g. Out of a total of 3 6000
mg polars per gram of rock, solvent extraction would
recover, as a conservative estimate, at least two-thirds
of the material, or 2 mg/g. This implies that Soxhlet
8000
extraction of a shale with oil-wet mineral surfaces

Depth (ft)
should afford an extract that contains a minimum of
~2 mg resins and asphaltenes per gram of rock. In con-
trast, measured resin and asphaltene yields in the 10000
Saum #1 well range from 0.05 to 0.5 mg/g, which is far
below the amount needed to satisfy the mineral sorp-
tion capacity. Hence, based on this information, it 12000
appears that the mineral surfaces of these samples
remain largely water wet in the depth range corre-
sponding to the zone of intense oil generation. 14000

Calculation of Oil Saturation


The variables needed to calculate oil saturation are 16000
porosity, fraction of EOM residing in the pore system,
and rock and oil densities. The total oil extract (cor-
rected EOM) is potentially comprised of three genetic 18000
components: total oil extract = residual oil in pores +
oil sorption by kerogen + oil sorption extract by miner- Figure 3. Oil saturation vs. depth relationship for
als. In shales with mineral surfaces that are predomi- Eagleford samples. The error bars for So correspond
nantly water wet, the mineral sorption term can be to calculated ranges using +25% error in porosity
disregarded, so the EOM is partitioned into two frac- and +33% error in kerogen sorption capacity (Ksc)
tions using the simplified equation (i.e., 40–80 mg/g). For display clarity, not all error
bars are shown.

 S ×φ × ρ 
EOMcorr =  o oil  + ( Ksc × TOC )
 Table 2, the degree to which depends on in-situ pres-
 ρ rock × 103  (1) sure, temperature, and composition. In the absence of
any significant oil loss during core recovery (a reason-
where EOMcorr = EOM corrected for light-end loss able assumption for shales), the calculated saturations
(mgHC/gRock); So = oil saturation (fraction); φ = porosity represent the volume of available pore space occupied
(fraction); ρoil and ρrock = bulk densities of oil and rock at sur- by oil that has shrunk to its equilibrated surface volume
face conditions (g/mL); Ksc = kerogen sorption capacity due to exsolution of gas. Shrinkage factors of 1.1–1.5 can
(mgHC/gC); and TOC = total organic carbon (fraction). be expected for samples in this study.
Rearrangement of equation 1 gives the expression There is additional uncertainty associated with each
for oil saturation: of the input variables, such as kerogen sorption capac-
ity, measured porosity, and TOC. Porosity, in particu-
ρ rock
So = [ EOM corr – (Ksc × TOC)] ×
1 lar, is likely to be a major contributor to the overall
× (2)
φ ρoil × 10–3 variance of calculated saturations. Upper and lower
limits have been estimated for the key variables, which
Equation 2 was used to calculate oil saturation in the were used in a Monte Carlo simulation to calculate an
Eagleford samples. The results are presented in Table 2, error range for oil saturation. The resulting ranges
along with values for variables used in the computa- (10/90 percentiles) are plotted as error bars in Figure 3.
tions. Note that surface density of oil is used in the cal- Many factors can influence the amount of resid-
culation. This value is typically higher than the density ual oil saturation in shales. When considering geo-
of subsurface petroleum, which includes a component chemical variables, the most important of these are
of solution gas (England et al., 1987). The actual subsur- the level of maturation and the organic richness. In
face petroleum saturations (oil with solution gas) will Figure 3, the depth axis is used to approximate the
therefore be somewhat higher than those shown in level of maturation. Maturation modeling studies
Oil Saturation in Shales 21

5.00 Figure 4. Relationship


TOC below 1 % between EOM and
Rock-Eval S1 peak.
4.00 TOC above 1 % Rk = rock.
Rock-Eval S1
(mgHC/gRk)

3.00

2.00

1.00

0.00
0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00

Extractable Organic Matter


(mgHC/gRk)

calibrated to measure vitrinite reflectance and subsur- range typically used to model primary migration from
face temperatures have shown this to be a reasonable source rocks via two-phase Darcy flow (England et al.,
assumption (Westcott and Hood, 1994). Nevertheless, 1987; Durand, 1988; Okui and Waples, 1992).
a few anomalies are apparent. For instance, heat flow
in the vicinity of the Todd #1 well from West Feliciana Relationship of Oil Saturation
Parish, Louisiana, is lower than in the east Texas coun- and Rock-Eval Parameters
ties, resulting in lower maturation of Todd #1 samples
(12,967–13,013 ft) than at equivalent depths in other In order to arrive at oil saturation values by using
wells. Taking these variations in thermal regime into the above procedure, a good deal of high-quality core
account, the zone of oil generation for Eagleford kero- data are required. However, it may not always be pos-
gens generally spans the depth range from ~7500 to sible to acquire this amount of data, particularly when
13,000 ft. working with small amounts of samples such as side-
Organic richness variability is illustrated in Figure 3 wall cores. Rock-Eval pyrolysis is a readily available
using different symbols for samples with TOC above screening method that is inexpensive and routinely
and below 1%. Most of the low-TOC samples have oil applied to small quantities of shaly samples (~100
saturations below ~5%, and only one sample has a mg). Correlation of typical Rock-Eval parameters and
slightly higher value at ~11% So. These low oil satura- oil saturation was therefore investigated for the pur-
tions are due to insufficient reactive kerogen for trans- pose of establishing a rapid means of quantifying
formation into free oil. The richer samples show more residual oil saturation in shales.
variability with depth. At depths <7500 ft, the shales The Rock-Eval parameter designed to measure the
with high TOC (>1%) have low saturations (<20%). amount of free hydrocarbons by thermal “distillation”
This is interpreted as being due to insufficient thermal is the S1 peak (Tissot and Welte, 1984). Figure 4 shows
maturity for significant oil generation. Samples within a cross-plot of S1 vs. EOM for the Eagleford shale sam-
the oil window (7500–13,000 ft) have saturations rang- ples. As expected, a good linear correlation is observed
ing from 15 to 70%, with most samples having values between the amount of distillable (S1) and extractable
in the 25–40% range. At greater depths, oil saturations (EOM) hydrocarbons. Note that the line of best fit does
are again low (<10%), presumably due to overmatu- not have a slope of 1, since the two parameters quan-
rity and initiation of oil-to-gas cracking. Total in-situ tify different components of the residual oil (Cooles et
hydrocarbon saturations (oil + gas) are expected to al., 1986; Wilhelms et al., 1991). The S1 parameter typi-
either remain the same or increase with depth as the cally measures volatile hydrocarbons in the C 5–C 25
fluid becomes increasingly gas rich. However, it was carbon number range, whereas the EOM measures
not possible to demonstrate this experimentally due to C12+ components, including nonvolatile entities such
loss of gas from the cores at the surface. The residual as asphaltenes and high-molecular-weight polar
liquid hydrocarbon data shown in Figure 3 are in gen- (NSO) compounds (Larter, 1988). In addition, hydro-
eral agreement with intuitive trends; that is, low oil carbon sorption is expected to have a differential effect
saturations in the immature zone, building to higher on the two analytical methods, with the S1 pyrolysis
saturations during active oil generation and expulsion, measurement being affected to a greater extent than
and finally decreasing again at elevated maturity due EOM solvent extraction (Tarafa et al., 1983).
to thermal destruction of residual oil. The So values for As with extractable organic matter (EOM), the S1
samples in the oil window (~25–40%) are also in the yield is not a direct measure of So , since it includes
22 Noble et al.

100 Figure 5. Plot of oil saturation


vs. hydrocarbon content
90 TOC below 1 % (S1/TOC).
80 TOC above 1 %
70
Oil Saturation

60
So (%)

50
40
30
20
10
0
0 50 100 150 200 250 300 350 400
Hydrocarbon Content
S1/TOC (mgHC/gC)

both sorbed hydrocarbons and hydrocarbons residing kerogenous shales undergoing natural maturation.
within the intergranular pore system. The magnitude The one outlying data point in Figure 5 is from a sandy
of the S1 peak is primarily a function of organic rich- reservoir interval in the Saum #1 well. This well dis-
ness and thermal maturity. In order to take into covered about 12 ft of oil column in poor- to moderate-
account the effects of organic richness, the S1 peak is quality reservoir rock. The hydrocarbon content of the
typically normalized to TOC, the resulting parameter reservoir sample is extremely high (~355 mgHC/gC)
being referred to at ARCO as the “hydrocarbon con- and is well above the indigenous hydrocarbon trend.
tent” (the bitumen index; Wilhelms et al., 1991). The It provides a clear example of the effects of nonindige-
hydrocarbon content, expressed in units of mgHC/gC, nous (migrated) oil saturation on the magnitude of the
is a particularly attractive parameter for estimating hydrocarbon content value.
residual So, since it implicitly accounts for the variable Unusually large S1 peaks have been used for many
amounts of sorbed hydrocarbons in rocks with differing years as a means to detect nonindigenous hydrocar-
kerogen concentrations. bons (Clementz, 1979; Peters, 1986). The latter may
Figure 5 shows a plot of oil saturation (So), calcu- result from both natural processes (i.e. impregnation
lated from equation 2, vs. Rock-Eval hydrocarbon con- with migrating oil) or drilling-induced staining (e.g.,
tent (S1/TOC). A general relationship of increasing oil diesel additives). The empirically determined cutoff
saturation with S 1/TOC is apparent for all but one for the S1/TOC parameter is typically in the range of
sample. The relationship is interpreted to represent the 150–180 mgHC/gC (Lundell, 1991). The Eagleford data
normal maturation trend that occurs in these shales presented here show that this cutoff does, in fact, cor-
during their passage through the oil and gas windows. respond very well with the value expected at 100% So
Note the observed So ranges from 0 to ~70%, and the in these shales.
corresponding hydrocarbon content ranges from 0 to
~120 mgHC/gC. If this trend is extrapolated to 100%
So , within the limits of the data, the resulting hydro- APPLICATIONS IN SEAL EVALUATION
carbon content may vary from ~130 to 200 mgHC/gC. Seal Potential
This range of values is expected to correspond to the
upper limit that one might expect to find in shales The evaluation of seal potential is an important part
within the oil window. Expressing this in a different of any exploration project. Kaldi and Atkinson (1997)
way, the S1/TOC value that corresponds to 100% So have discussed three key elements that require consid-
represents the maximum indigenous hydrocarbon eration in seal evaluation: seal capacity, seal integrity,
content that arises from in-situ hydrocarbon genera- and seal geometry. Seal capacity deals with the capil-
tion and expulsion. lary properties of seal rocks and their capacity to sup-
The precise value for the cutoff between indigenous port a hydrocarbon column in an adjacent reservoir.
and nonindigenous hydrocarbons is dependent on a Seal integrity and seal geometry are concerned with
number of factors and is expected to show some differ- rock mechanics (brittle vs. ductile deformation) and the
ences from one source rock formation to another. The spatial distribution of seal facies, respectively. All three
interpretation of the data shown in Figure 5 implies elements are of equal importance in seal evaluation; in
that samples with hydrocarbon contents above ~160 terms of using the geochemical concepts discussed in
mgHC/gC (midpoint of range) contain “nonindige- this report, seal capacity is of primary concern.
nous” hydrocarbons; that is, they contain more free When drilling an exploration well, information that
hydrocarbons than one would expect to find in is typically sought is whether the target reservoir is
Oil Saturation in Shales 23

Figure 6. Map of Ardjuna


Basin in the vicinity of
the BZZ oil field.

present and whether there is a hydrocarbon column expected relative to structural spill point, then analysis
worthy of further testing. In many instances, the ques- of the saturation profile through the seal may show
tion about the seal is not given adequate attention, that the size of the column is at its maximum level rel-
because many people consider it to be implicitly ative to seal capacity. There may also be other
addressed in the evaluation of the hydrocarbon col- instances where the so-called “seal” is, in fact, a poor-
umn; that is, if a column is present, there must be a quality reservoir (waste zone), and that the actual seal
seal. If hydrocarbons are not present, a number of fac- occurs higher in the section.
tors could be responsible (e.g., charge, seal, and/or
trapping geometry). Indigenous vs. Nonindigenous Hydrocarbons
A great deal of valuable information can, in fact, be in Seals
derived from analyzing samples of the seal, in the case
of both dry holes and hydrocarbon discoveries. The Many fine-grained lithofacies that serve as seals for
use of mercury injection capillary pressure (MICP) hydrocarbon accumulations are known to contain
data in seal evaluation has been known for many appreciable amounts of kerogen. For example, the
years, and the geological applications of this technique HRZ shale (North Slope of Alaska), which seals the
have been reviewed by Vavra et al. (1992). Specifically, main Prudhoe Bay Reservoir, is an organic-rich shale
MICP provides information about pore size distribu- with TOC values often >3–4%. Similarly, the kerogen-
tions, pore entry pressures, and maximum hydrocar- rich Devonian Woodford shale seals many fields in the
bon column heights that may exist due to capillary Permian basin of west Texas. Shales such as these will
contrast between reservoir and seal facies. However, develop their own saturation profile as a function of
the technique does not address the type and amount of depth (maturation), regardless of whether they seal a
fluids originally present in the pore system of the sam- hydrocarbon accumulation. The petroleum that is gen-
ple, and whether hydrocarbons, if present, occur in erated from the kerogen or inherited directly from pre-
appreciable concentration. This information can be cursor organisms is referred to as “indigenous
obtained in a relatively straightforward manner from hydrocarbon saturation.” On the other hand, any
geochemical studies, and can provide some key hydrocarbon moving into the rock as a result of buoy-
insights in the evaluation of drilling results. For ancy forces exceeding pore entry pressure is referred
instance, if a well is commercially “dry,” and it can be to as “nonindigenous hydrocarbon saturation.” Dis-
shown that the seal rock is saturated with nonindige- tinguishing the two is essential for proper evaluation
nous hydrocarbons, the reason for failure is probably of seal performance.
not charge limitations, but more likely due to a poor Two techniques are used to recognize nonindige-
seal that is incapable of supporting the required nous hydrocarbons in seals. The first involves quanti-
amount of hydrocarbons. Alternatively, if a discovery fying the amount of free hydrocarbons in the sample
is made, but the hydrocarbon column is smaller than and then comparing this amount to the quantity that
24 Noble et al.

100000 Figure 7. Calculated range


of hydrocarbon column
Hydrocarbon Column Height (ft)

Shelf heights that could be


Carbonates supported by Talang Akar
10000 seals (after Kaldi and
Atkinson, 1997).
Channel
1000 Abandonment
Silts Delta Front
Delta Plain Shales
Pro-Delta
Shales Shales
100

Increasing Seal Capacity


10

would be expected from normal maturation of the kero- It is important to note that even though the shelf
gen. If the observed quantity exceeds the expected carbonates have the highest sealing capacity, they do
amount, nonindigenous hydrocarbons are likely to be not necessarily make the best seals. In fact, these car-
present. The second method involves analysis of the bonates make rather poor seals, since they are quite
molecular composition of the hydrocarbons. Biomarker brittle and prone to fracture (i.e., seal integrity is low).
and molecular fingerprinting of seal extracts can be Nevertheless, in terms of their capacity to prevent
used to determine if the hydrocarbons were generated hydrocarbons from entering the matrix pore system,
in situ or if they are migrated from a remote source. they are extremely tight and require very high pres-
A case study from the Talang Akar Formation, off- sures in the intruding fluid.
shore northwest Java, Indonesia, is presented in which
the application of both techniques is demonstrated. Hydrocarbon Content vs. Seal Capacity
The residual hydrocarbon saturation in the Talang
TALANG AKAR SEAL STUDY Akar seal samples was quantified using geochemical
Seal Analysis of Various Facies methods. Small pieces of conventional and sidewall
core were available for analysis. There was insufficient
The Talang Akar Formation consists of a sequence material to perform a detailed EOM/So study by using
of fluvio-deltaic to shallow-marine sands, shales, Soxhlet extraction, so Rock-Eval was used as the pri-
coals, and limestones of upper Oligocene age. Much mary analytical tool. The hand samples were carefully
has been published on the petroleum geology of this examined and showed no evidence of fractures or
region, and the reader is referred to the literature for external contamination with drilling additives. Hence,
background information (Ponto et al., 1988; Butter- it was assumed that any free hydrocarbons quantified
worth and Atkinson, 1993; Young and Atkinson, 1993; during the analytical procedure originated either from
Suria et al., 1994). A generalized location map showing the matrix pore system or by desorption from the solid
all the major producing fields in the region is pre- kerogen/mineral matrix.
sented in Figure 6. An evaluation of intraformational Table 3 shows the TOC and Rock-Eval data for vari-
seals from the BZZ field of the Ardjuna subbasin has ous Talang Akar facies. Some of the samples have quite
been conducted by Kaldi and Atkinson (1997). In that high TOC values and S2 pyrolysis yields and are consid-
study, the seal potentials of five Talang Akar facies ered good potential source rocks. However, at these
were evaluated: delta plain shales, channel abandon- well locations, all samples are insufficiently mature for
ment silts, delta front shales, prodelta shales, and shelf significant hydrocarbon generation, as shown by the
carbonates. Kaldi and Atkinson (1997) used mercury Tmax values (≤436°C). The samples were taken at true
injection capillary pressure (MICP) data for represen- vertical depths shallower than about 6000 ft, which, on
tative samples to calculate hydrocarbon column a regional basis, places them outside the zone of intense
heights that could be supported by each sealing facies, oil generation (Noble et al., 1991). From a maturation
as shown in Figure 7. The shelf carbonate facies was standpoint, it is unlikely that the oil accumulations at
found to have the highest seal capacity, with maxi- BZZ were generated locally by these organic-rich rocks.
mum supportable columns in excess of 1000 ft. At the On the other hand, these fine-grained facies provide
other end of the scale, delta plain shales were found to intraformational seals for sandstone reservoirs.
have the lowest entry pressures, with maximum col- The hydrocarbon content (S1/TOC) provides a mea-
umn heights ~100 ft. sure of the amount of free hydrocarbons relative to the
Oil Saturation in Shales 25

Nonindigenous
250

HCs
Hydrocarbon Content

200
S1/TOC (mgHC/gC)

150

100

Indigenous
HCs
50

0
Delta Plain Channel Abandonment Pro-Delta Delta Front Shelf Carbonate
Shale Siltstone Shale Shale

Figure 8. Hydrocarbon content (S1/TOC) of Talang Akar seals. Values above 150 mg/g are indicative of non-
indigenous hydrocarbons, as is seen for the delta plain facies.

total organic pool. As shown in the study of Eagleford 10,000


shale samples, this parameter varies systematically • Delta Plain Shales
with calculated values of hydrocarbon saturation in the Delta Front Shales
Pro-Delta Shales
0–70% saturation range (Figure 5). The corresponding Ch. Aband. Silts
S1/TOC values range from 0 to 120 mgHC/gC, repre-
senting the amount of indigenous hydrocarbon that 1000
Hydrocarbon Column

was generated directly from the host kerogen. Values


above the 130–200-mgHC/gC range are considered to
Height (ft)

be the result of nonindigenous hydrocarbons moving


into the pore/kerogen network from an external
source. The precise cutoff to detect nonindigenous 100

hydrocarbons may vary from one formation to


another, but is expected to fall within the general range
of 150–180 mgHC/gC. In the case of the Talang Akar
samples, S1/TOC values above ~160 were assumed to
10
contain some nonindigenous hydrocarbons.
10 100 1000
Figure 8 shows the hydrocarbon contents of the var-
Hydrocarbon Content
ious Talang Akar seal facies. The values range from 25 S1/TOC (mgHC/gRk)
to 241 mgHC/gC, with three samples recording yields
>100 and two >160 (Table 3). The two samples with the Figure 9. Plot of hydrocarbon content vs. column
highest S1/TOC values are delta plain shales. Their height from MICP. Rk = rock.
S1/TOC values of 228 and 241 are strongly indicative
of nonindigenous hydrocarbons. A plot of hydrocar-
bon content vs. maximum possible column height hydrocarbons in the pore/kerogen matrix could be
(from MICP) shows that the same two samples have due to oil columns smaller than the potential holding
the lowest sealing capacity, capable of supporting oil capacity of the seal, or to seal failure occurring via
columns of only ~100 ft (Figure 9). At the other end of some other pathway with lower entry pressures (such
the scale, samples with low hydrocarbon contents that as fractures).
clearly fall in the expected range for indigenous
hydrocarbons have much higher seal capacities Molecular Fingerprinting of Hydrocarbons
(potential oil columns of 250 to >1000 ft). These results
indicate that the poorest sealing facies (delta plain Additional evidence for distinguishing indigenous
shales) have probably been impregnated with oil from from nonindigenous hydrocarbons was obtained from
an underlying hydrocarbon column that exceeded the the molecular signature of the extractable organic mat-
maximum seal capacity. In contrast, all other facies ter. The technique used was thermal extract gas chro-
show no evidence of oil saturation over and above that matography (TEGC), in which the “distillable”
expected from natural in-situ generation from kero- hydrocarbons (equivalent of S1 peak) are liberated
gen. In these samples, the absence of nonindigenous from the rock and analyzed by gas chromatography.
26 Noble et al.

Figure 10. Gas chromatograms


for Talang Akar samples: (a)
thermal extract of “good” seal;
(b) thermal extract of “poor”
seal; (c) whole oil gas chro-
matogram of produced BZZ oil.
pr = pristane; n17 = normal C17
alkane.
Oil Saturation in Shales 27

Figure 10 shows the TEGC traces for two Talang Akar in the oil window (10–80%), and then decreasing
seal samples: one which is described as a “good” seal again at higher maturities. Mass balance suggests
(high entry pressure, low hydrocarbon content) and that the mineral grains in our Eagleford shales
the other, as a “poor” seal (low entry pressure, high (≤3.8% TOC) remain water wet during hydrocar-
hydrocarbon content). Also shown is the gas chro- bon generation and expulsion. The kerogen
matogram of a Talang Akar crude oil from a produc- matrix, on the other hand, is considered to be oil
ing well in the BZZ field. It is clear from Figure 10 that wet at all times.
the good seal yields very small amounts of hydrocar- 3. Oil saturation values correlate with the “hydro-
bons that do not have the same characteristics as the carbon content” parameter (S1/TOC) from Rock-
reservoir oil. In contrast, the TEGC trace for the poor Eval. Values for this parameter <120 mgHC/gC
seal affords a molecular pattern that is visually much correspond to oil saturations below 100% in
more akin to the BZZ oil. Note that this comparison is water-wet shales. Values >120 mg/g may indi-
based on the C 12+ carbon number range, since the cate the presence of some nonindigenous hydro-
gasoline and kerosene range hydrocarbons (<C12) are carbons, whereas values >200 mg/g almost
not present in the rock extracts due to sample handling certainly do. Nonindigenous hydrocarbons (i.e.,
and preparation. those that were not generated by the host kero-
The largest peak in the good seal chromatogram is gen and hence are from another source) result
pristane (pr), which occurs in much higher abundance either from contamination with drilling fluids
than the adjacent nC17 alkane (pr/nC17 = 8.8). The ratio (e.g., diesel) or by impregnation of oil during
of pristane to nC17 generally decreases with maturation, migration and entrapment.
and values this high typically indicate immaturity. This 4. A case study from the Talang Akar Formation,
observation agrees with Rock-Eval Tmax and regional Indonesia, shows that seal rocks with high entry
maturity trends. The sample lacks the full suite of pressures (from MICP analysis) have low hydro-
n-alkanes that are normally found in non-biodegraded carbon contents in the range expected for in-situ
oil-bearing rocks. The overall signature is consistent generation. However, shales with the lowest
with indigenous hydrocarbons derived from a rela- entry pressures have very high hydrocarbon con-
tively poor quality, immature source rock. In contrast, tents, indicating impregnation of the pore system
the hydrocarbon signature of the poor seal is much with oil from an underlying accumulation. In
more oil-like. Pristane is still the largest component, but such samples, the seal rock has most probably
the n-alkanes are present in greater quantities (pr/nC17 attained equilibrium with the maximum oil col-
= 3.7). A full suite of n-alkanes is clearly apparent, par- umn height it was capable of supporting. This
ticularly in the waxy region (>nC25). Comparing the information on seal performance is of value in
pr/nC17 ratio of the BZZ oil (pr/nC17 = 1.9) with the seal economic evaluations as well as in examining
extracts shows that the poor seal has intermediate val- exploration opportunities in shallower horizons.
ues between the good seal (indigenous hydrocarbons)
and the reservoired oil (migrated hydrocarbons). This is ACKNOWLEDGMENTS
interpreted to be the result of mixing between the
indigenous hydrocarbon fraction of the shale with oil The authors express their appreciation to the ARCO
that has entered the rock as a result of the relatively low Geochemistry lab staff in Plano, Texas, for analytical
entry pressures. The pr/nC17 ratio was chosen to illus- data, including Luc Vo for his painstaking quantitative
trate this point, although several other parameters that extract and oil evaporation work. The advice of Steve
are based on more complicated biomarker structures Franks on the locations and geological setting of Eagle-
could have been used for the same purpose. The molec- ford shale cores is appreciated. Helpful reviews of the
ular fingerprinting technique provides evidence to sup- manuscript were provided by Alton Brown, Steve
port the S1/TOC pyrolysis data that indicate seals with Franks, and Keith Katahara (ARCO), and AAPG
low entry pressures contain some nonindigenous reviewers Brian Fowler and Leta Smith. This article is
hydrocarbons, whereas those with higher entry pres- published with permission of the managements of
sures do not. ARCO Exploration and Production Technology and
Atlantic Richfield Indonesia Inc., whose support for
this project is appreciated.
SUMMARY AND CONCLUSIONS
1. The amount of free oil in the pore system of REFERENCES CITED
shales, referred to as residual oil saturation (So),
can be estimated from geochemical data and Anderson, W.G., 1986, Wettability literature survey.
porosity measurements. The key geochemical Part 1: rock/oil/brine interactions and the effects of
variables needed to calculate oil saturation are core handling on wettability: Journal of Petroleum
extractable organic matter (EOM) corrected for Technology, p. 1125–1144.
light-end loss, total organic carbon (TOC), and Boudou, J.P., 1984, Chloroform extracts of a series of
kerogen sorption capacity. coals from Mahakam delta: Organic Geochemistry,
2. Oil saturation values calculated in this way are v. 6, p. 431–437.
low for immature samples (So <10%), increasing Brown, A.A., 1993, Expulsion efficiency of methane
28 Noble et al.

from vitrinitic coal: ARCO unpublished research p. 194–204.


memorandum, RM 93–0012, p. 1–29. Levine, J.R., 1993, Coalification: The evolution of coal
Butterworth, P., and C.D. Atkinson, 1993, Syn-rift as a source rock and reservoir rock for oil and gas,
deposits of the Northwest Java basin: Fluvial sand- in B.E. Law and D.D. Rice, eds., Hydrocarbons from
stone reservoirs and lacustrine source rocks, in C. D. coal: Tulsa, AAPG Studies in Geology, v. 38,
Atkinson, J. Scott, and R. Young, eds., Clastic rocks p. 39–77.
and reservoirs of Indonesia: Indonesian Petroleum Littke, R., and D. Leythaeuser, 1993, Migration of oil
Association Core Workshop, p. 211–219. and gas in coal, in B.E. Law and D.D. Rice, eds.
Clementz, D.M., 1976, Interaction of petroleum heavy Hydrocarbons from coal: Tulsa, AAPG Studies in
ends with montmorillonite: Clays and Clay Miner- Geology, v. 38, p. 219–236.
als, v. 24, p. 312–319. Lundell, L., 1991, Topics in source rocks II: Recogni-
Clementz, D. M., 1979, Effects of oil and bitumen satu- tion of source rocks and migrated hydrocarbons
ration on source rock pyrolysis: AAPG Bulletin, based on Rock-Eval analyses: ARCO unpublished
v. 63, p. 2227–2232. Research Memorandum, RM 91–0005, p. 27.
Collins, S.H., and J.C. Melrose, 1983, Adsorption of Mackenzie, A.S., I. Price, D. Leythaeuser, P. Muller,
asphaltenes and water on reservoir rock minerals: M. Radke, and R. Schaefer, 1987, The expulsion
Society of Petroleum Engineers, v. 1180, p. 249–256. behavior of petroleum from Kimmeridge clay
Cooles, G.P., A.S. MacKenzie, and T.M. Quigley, 1986, source rocks in the area of the Brae oil field, UK con-
Calculation of petroleum masses generated and tinental shelf, in J. Brooks and K. Glennie, eds.,
expelled from source rocks: Organic Geochemistry, Petroleum geology of north-west Europe: London,
v. 10, p. 235–245. Graham & Trotman, p. 865–877.
Cuiec, L., 1984, Rock/crude oil interactions and wetta- Noble, R.A., C.H. Wu, and C.D. Atkinson,1991,
bility: an attempt to understand their interrelation: Petroleum generation and migration from Talang
Society of Petroleum Engineers, v. 13211, p. 1–14. Akar coals and shales, offshore NW Java: Organic
Durand, B., 1988, Understanding of HC migration in Geochemistry, v. 17, p. 363–374.
sedimentary basins (present state of knowledge): Okui, A., and D.W. Waples, 1992, Relative permeabili-
Organic Geochemistry, v. 13, p. 445–459. ties and hydrocarbon expulsion from source rocks,
Durand, B., A.Y. Huc, and J.L. Oudin, 1989, Oil satura- in A.G. Dore et al., eds., Basin modeling: advances
tion and primary migration: observations in shales and applications: Amsterdam, Elsevier, p. 293–301.
and coals from the Keribu wells, Mahakam Delta, Peters, K.E., 1986, Guidelines for evaluating petroleum
Indonesia, in B. Doligez, ed., Migration of hydrocar- source rock using programmed pyrolysis: AAPG
bons in sedimentary basins: Paris, Editions Technip, Bulletin, v. 70, p. 318–329.
p. 173–195. Ponto, C.V., C.H. Wu, A. Pranoto, and W.H. Stinson,
Engel, M.H., S.W. Imbus, and J.E. Zumberge, 1988, 1988, Improved interpretation of the Talang Akar
Organic geochemical correlation of Oklahoma depositional environment as an aid to hydrocarbon
crude oils using R- and Q-mode factor analysis: exploration in the ARII offshore northwest Java
Organic Geochemistry, v. 12, p. 157–180. contract area, in Proceedings of Indonesian
England, W.A., A.S. MacKenzie, D.M. Mann, and T.M. Petroleum Association 17th Conference, Jakarta,
Quigley, 1987, The movement and entrapment of p. 397–422.
petroleum fluids in the subsurface: Journal of the Radke, M., R.G. Schaefer, and D. Leythaeuser, 1980,
Geological Society of London, v. 144, p. 327–347. Composition of soluble organic matter in coals: Rela-
Gorsline, D.S., 1984, A review of fine-grained sediment tion to rank and liptinite fluorescence: Geochemica
origins, characteristics, transport and deposition, in and Cosmochimica Acta, v. 44, p. 1787–1800.
D.A.V. Stow and D.J.W. Piper, eds., Fine-grained Sandvik, E.I., W.A. Young, and D.J. Curry, 1992,
sediments: The Geological Society (Blackwell Scien- Expulsion from hydrocarbon sources: The role of
tific) Special Publication 15, p. 17–34. organic absorption: Organic Geochemistry, v. 19,
Halbouty, M.T., and J.J. Halbouty, 1982, Relationships p. 77–87.
between East Texas field region and Sabine uplift in Sofer, Z., 1988, Biomarkers and carbon isotopes of oils
Texas: AAPG Bulletin, v. 66, p. 1042–1054. in the Jurassic Smackover Trend of the Gulf Coast
Kaldi, J.G., and C.D. Atkinson, 1997, Evaluating seal States, USA: Organic Geochemistry, v. 12, p. 421–432.
potential: example from the Talang Akar Forma- Suria, C., C.D. Atkinson, S.W. Sinclair, M.J. Gresko,
tion, offshore NW Java, Indonesia: AAPG Memoir and B. Plahaperdana, 1994, Application of inte-
67, p. 85–101. grated sequence stratigraphic techniques in non-
Koepf, E.H., 1978, Core handling—core analysis meth- marine/marginal sediments; an example from the
ods, in D.C. Bond, ed., Determination of residual oil Upper Talang Akar Formation, Offshore Northwest
saturation: Oklahoma City, IOCC, p. 36–71. Java: Proceedings of the 23rd Indonesian Petroleum
Larsen, J.W., J.C. Cheng, and C.S. Pan, 1991, Solvent Association, v. I, p. 145–160.
extraction of coals during analytical solvent swelling. Tarafa, M.E., J.M. Hunt, and I. Ericson, 1983, Effect of
A potential source of error: Fuel, v. 5, p. 57–59. hydrocarbon volatility and adsorption on source
Larter, S., 1988, Some pragmatic perspectives in source rock pyrolysis: Journal of Geochemical Exploration,
rock geochemistry: Marine Petroleum Geology, v. 5, v. 18, p. 75–85.
Oil Saturation in Shales 29

Tissot, B.P., and D.H. Welte, 1984, Petroleum forma- Organic Geochemistry, v. 17, p. 351–354.
tion and occurrence 2nd. ed.: Berlin, Springer- Yee, D., J.P. Seidle, and W.B. Hanson, 1993, Gas sorp-
Verlag, 699 p. tion on coal and measurement of gas content, in B.E.
Vavra, C.L., J.G. Kaldi, and R.M. Sneider, 1992, Geo- Law and D.D. Rice, eds., Hydrocarbons from Coal:
logical applications of capillary pressure: A review: Tulsa, AAPG Studies in Geology, v. 38, p. 203–218.
AAPG Bulletin, v. 76, p. 840–850. Young, R., and C.D. Atkinson, 1993, A review of
Westcott, W.A., and W.C. Hood, 1994, Hydrocarbon Talang Akar Formation (Oligo-Miocene) reservoirs
generation and migration routes in the East Texas in the offshore areas of SE Sumatra and NW Java, in
basin: AAPG Bulletin, v. 78, p. 287–307. C.D. Atkinson, J. Scott, and R. Young, eds., Clastic
Wilhelms, A., S.R. Larter, and D. Leythaeuser, 1991, rocks and reservoirs of Indonesia: Indonesian
Influence of bitumen-2 on Rock-Eval pyrolysis: Petroleum Association Core Workshop.
Krushin, J.T., 1997, Seal capacity of nonsmectite shale,
in R.C. Surdam, ed., Seals, traps, and the petroleum
system: AAPG Memoir 67, p. 31–47.

Chapter 3

Seal Capacity of Nonsmectite Shale


James T. Krushin1
Amoco Exploration and Production Company, U.S.A.

ABSTRACT
Laboratory-derived petrophysical measurements confirm that the sealing
shale’s largest interconnected pore throats can limit the size of hydrocarbon
columns. These largest interconnected pore throats define the seal capacity
of the shale. Interpreting displacement pressure from high-pressure mercury
injection porosimetry (MIP) data permits calculating seal capacity, the
hydrocarbon column-limiting capillary property of the rock. Displacement
pressure is the pressure at which the nonwetting phase (i.e., mercury in the
laboratory tests) begins to displace the wetting phase from the largest inter-
connected pore throats. The 12 well-indurated nonsmectite shales studied
range in age from Precambrian to Jurassic and vary in mineralogy, porosity,
permeability, cation exchange capacity, organic content, and stratification.
The shales are treated as two distinct groups with respect to interpreting dis-
placement pore throat size: nonorganic shales and organic shales. Estimation
of mineral percentages by X-ray diffraction analysis, and classifying the
shales according to silt/clay ratios, laminations, and major nonclay/nonsilt
mineralogy, permit petrographic prediction of seal capacity for nonorganic
shales. Quartz content of the matrix is the best predictor of the displacement
pore throat size for nonorganic shales. Sandy mudstones have the largest
measured tabular displacement pore throats for nonorganic shales and are in
the 30–40 nm range. This pore throat size range can limit the size of very
large gas columns. Clay-rich and calcareous shales have such small
displacement pore throats (<15 nm) that they are excellent capillary seals. The
organic shales studied have large displacement pore throats relative to their
low porosity when compared to nonorganic shales. Volume reduction of the
matrix associated with hydrocarbon generation contributes to the largest
pore throats in organic shales.

INTRODUCTION Seal assessment has equal importance with the other


geologic parameters needed for a hydrocarbon accumu-
Prospect evaluation involves properly assigning lation. Whereas reservoirs are often cored, described,
technical risk to reservoir, structure, charge, seal, etc. and documented, structures are mapped and further
refined; seals, specifically shale seals, are not studied
1Currently consultant, 13102 Fallsview Lane, #4904-F, Houston, Texas as often. The maximum hydrocarbon column an

31
32 Krushin

undeformed shale can seal is determined by its capil- Seal capacity is breached when the hydrocarbon
lary properties. There is little published quantifiable column generates enough buoyancy pressure to begin
data on the capillary sealing ability of shales. This is saturating the seal. The laboratory technique of mer-
surprising because shales are often forecast to seal cury injection porosimetry (MIP) identifies the pres-
hydrocarbons and they comprise large quantities of sure required to begin saturating the seal. In this
the drillable crust where oil and gas exploration occur. paper, displacement pressures, visually estimated by
Potter et al. (1980) estimate shale forms more than 60% backward extrapolation of log Pc vs. Sw on MIP pro-
of sediments worldwide, and Jones and Wang (1981) files, define seal capacity for different shale types (Fig-
state shales comprise over 75% of the clastic fill in sed- ure 1). Mercury injection porosimetry is a laboratory
imentary basins. technique that has been in use for about 50 yr to simu-
Zieglar (1992) studied the hydrocarbon column late migrating hydrocarbons (Jennings, 1987). MIP
heights for unfaulted accumulations in the Rocky data are obtained by injecting mercury, which acts as
Mountain area with respect to shale seal efficiency. the nonwetting phase, into the sample in a stepwise
He used limited and tenuous shale seal capillary fashion. The percentage pore volume, saturated by
measurements, and concluded that the longest 10% mercury at each pressure increment, is recorded. The
hydrocarbon columns were being limited by shale results are usually plotted as mercury saturation vs.
capillary sealing capacity. Therefore, this work, injection pressure (Jennings, 1987; Vavra et al., 1992).
which uses laboratory-derived measurements, was In practice, there is mercury saturation recorded prior
undertaken to determine the relationship between to reaching the displacement pressure, but this mer-
shale compositions and the maximum hydrocarbon cury saturation at low pressure (beginning with entry
column that they can seal. pressure) is attributed to mercury infilling surface
This study examines and describes some basic shale irregularities, such as nicks, gouges, and vugs on the
types, and shows how petrographic and petrophysical sample. (Vavra et al., 1992). These surface irregulari-
measurements predict their seal capacity. The paper ties are attributed to sampling techniques and are not
begins by introducing previously published studies on believed to be present in in-situ conditions.
seals, and then continues by describing and classifying Displacement pressure should not be confused with
selected shales petrographically. Next, petrophysical breakthrough pressure. Breakthrough pressure is the
measurements on vertically oriented shale plugs are minimum pressure required to cause a continuous
described, and these results are interpreted with nonwetting phase (i.e., hydrocarbon rivulet in the
regard to seal capacity. seal). Establishing a continuous nonwetting phase
causes permeability to hydrocarbons, and Schowalter
(1979) estimates that about one-half of the hydrocar-
BACKGROUND bon column could be lost before seal capacity is
regained. For unfractured, fine-grained siltstones,
Proper techniques for evaluating seals are dis-
Rudd and Pandey (1973) estimate that the nonwetting
cussed in Schowalter (1979), Downey (1984), Watts
phase needs to saturate the pore volume to 20–30%
(1987), Ulmishek (1988), and Kaldi and Atkinson
before breakthrough would occur.
(1994, 1997). These authors state that proper capil-
This study uses displacement pressure to define
lary evaluation of top seals involves quantifying not
seal capacity. Once the displacement pressure of the
only the properties of the seal (i.e., pore throat dis-
top seal is reached, any hydrocarbons migrating into
tribution) but also the properties of the fluids (e.g.,
the underlying reservoir would not go to extending
wettability, contact angles, and hydrodynamic con-
the column in the reservoir but to saturating the crestal
dition), if present. This paper does not address many
portion of the seal. At this point, the reservoir column
of the other important sealing scenarios, such as
would be at its limit. This happens because the crest is
mechanical properties and geometry (both thickness
exposed to the longest hydrocarbon column and,
and lateral continuity); fault-related features,
therefore, the highest buoyancy pressures. The equa-
including gouge, deformation bands, and juxtaposi-
tion for calculating this maximum hydrocarbon col-
tion (Smith, 1966); pore pressure effects, or the prop-
umn height is
erties affecting diffusion of hydrocarbons (Krooss,
1987).
Kaldi and Atkinson (1994) classify top-seal poten- Pds – Pdr
tial based primarily on three criteria: (1) seal capacity H max = (1)
(the size of the hydrocarbon column a lithology can 0.433(ρ w – ρhc )
support based on both capillary and fluid properties),
(2) seal geometry (the structural position, thickness, where Hmax = maximum hydrocarbon column in feet,
and areal extent of the lithology), and (3) seal integrity and Pds and Pdr = the displacement pressures of the
(which includes the mechanical properties such as seal and reservoir, respectively, for the brine-hydro-
ductility and the propensity to fracture). Therefore, carbon system in question. For a good-quality reser-
since this paper attempts to describe the capillary seal- voir, the displacement pressure is assumed to be 0.
ing ability of nonsmectite shale seals, it will address The denominator, 0.433( ρ w – ρ hc ), = the buoyancy-
Kaldi and Atkinson’s “seal capacity” part of top-seal derived pressure generated from the density differ-
classification. ence between the brine ( ρ w) and the hydrocarbons
Seal Capacity of Nonsmectite Shale 33

(ρhc), respectively. The 0.433 is for unit conversions


and takes into account the gravitational constant
(Berg, 1975; Downey, 1984; Watts, 1987; Vavra et al.,
1992).

METHODS
Sample Selection
Shales for this study were chosen on the basis of
core availability. Considerations in choosing and
working with shale core include:

• They are rarely cored. Reservoir rocks are the pri-


mary targets of a coring program. When a shale is
cored, it is often by mistake.
• Shales are often fissile and/or fractured. If shales are
cored, they may be too fractured for proper tests.
• Shales are improperly stored. Shales that are not
very well indurated and those that are smectite
rich may disaggregate, desiccate, or swell. Also,
dissolution of soluble minerals may occur.

Shales were sampled from conventional cores.


Table 1 lists the cored wells. Well-indurated, nonsmec-
tite shales were chosen because the largest pore
throats measured in smectite-rich shales are regarded
to be an artifact of drying. Drying smectite removes Figure 1. Schematic representation of a typical shale
the interlayer water, causing the interlayer pores to mercury injection porosimetry (MIP) profile (adapt-
shrink and form artificial large pores in the sample. ed from Schowalter, 1979; Jennings, 1987).
Therefore, MIP is not an appropriate method of analy- Displacement pressure is determined by extending
sis for smectite-rich shales because of the prerequisite the slope of the plateau to “0” mercury saturation.
drying (Murray and Quirk, 1992). Although shales in Displacement pressure is the pressure required to
this study are classified as nonsmectite, some contain begin saturating the shale with the nonwetting
trace amounts of this clay mineral. phase. In practice, there is mercury saturation prior
The 12 samples in this study range in age from Pre- to reaching the displacement pressure, but this satu-
cambrian to Jurassic (Table 1). These samples cover a ration is attributed to mercury infilling surface
range of porosity, cation exchange capacity (CEC), irregularities induced from sampling. This mercury
organic content, mineralogy, and primary deposi- displacement pressure is converted to in-situ reser-
tional structures (i.e., laminated and nonlaminated). voir hydrocarbon column height or pore throat size
Table 2 shows the mineralogy and organic content of using equations 1 and 2, respectively. Also note the
each sample. similarities between this schematic mercury injec-
tion curve and the actual curves shown in Figure 8.
Shale Classification
The use and meaning of the term “shale” are worth (1980), shown in Figure 2, using the percentage of silt
a brief discussion to alleviate the confusion that has and the presence of laminations to classify shale types.
surrounded it. “Shale,” when used as a separate word This method of classifying shales applies when over
such as in the title, describes a general classification of 50% of the grains are less than sand-size (62 µ), and 67%
very fine grained, predominantly terrigenous, clastic of the grains are less than silt fraction (5–62 µ).
sedimentary rocks and has no specific reference to fis- Shales in this study are further classified based on
sility or laminations. This meaning is so ingrained in fabric and the fraction of silt. Silt fraction divides the
the literature (Tourtelot, 1960) and in common “oil- shales into two major categories; less than one-third or
patch” jargon that this use is justified. In this respect, between one-third and two-thirds, with correspond-
shale has equal weight with the general terms “sand- ing prefix of “clay” or “mud,” respectively. Lamina-
stone” and “limestone.” Shale also has the fairly pre- tions, which are parallel stratifications less than 10 mm
cise double meaning of a “laminated clayey rock,” and thick, subclassify the fabric (Lundegard and Samuels,
this is the source of much confusion (Tourtelot, 1960). 1980). Therefore, based on silt fraction and the pres-
In this report, “shale” is also used as a root word to ence or absence of laminations, a shale is placed in one
further classify the samples with laminations. of four categories (Figure 2). This classification proved
In this paper, shales are subdivided according to the useful because of the broad and readily identifiable
classification described by Lundegard and Samuels categories.
34 Krushin

Table 1. Well Name, Location, Formation, Age, Depth, and Shale Type of the Conventional Cores Used in This
Study.

Well Depth
Well Name Formation/Age Location Abbreviation (ft) Shale Classification
Amoco 1— Morrison, Greeley County, RB 1511 Calcareous,
Rebecca Bounds Whitehorse Kansas lithic, mudstone
and Day Creek RB 1532 Sandy mudstone
Formations/ RB 1713 Sandy mudstone
Permian–Jurassic RB 1761 Feldspathic
mudstone
RB 1820 Sandy mudstone
RB 1856 Sandy claystone

Amoco Fayetteville Shale/ Rogers County, CAT 1200 Calcareous claystone


Catoosa-12 Carboniferous Oklahoma CAT 1202 Calcareous claystone

Amoco Oman-1 Upper Abu Oman M 9649 Dolomitic mudshale


Maimum Maharah
Formation/
Precambrian

Amoco— Antrim Shale/ Ostego County, S 1433 Organic clayshale


1–28 Swift Late Devonian Michigan S 1461 Organic clayshale
to Early S 1537 Organic clayshale
Carboniferous

Shale Composition Determination have found (Figure 3). Additionally, Shaw and
Weaver (1965), by using X-ray diffraction-absorption
Estimation of clay mineral compositions and techniques on 300 Paleozoic through Tertiary shale
amounts are based primarily on X-ray diffraction samples, concluded that the average major minerals’
analysis, similar to the shale composition work of composition is 31% quartz and 61% clay.
O’Brien and Slatt (1990). X-ray diffraction analysis used The silt-size and larger grain populations were esti-
CuKα radiation. Integrating peak areas and using a set mated by optical microscopy and standard compari-
of external standards identified the relative mineral son charts (Terry and Chilingar, 1955). In addition to
abundances for the bulk sample (Cullity, 1978). Clay silt/clay fraction and laminations used to classify the
minerals were specifically identified in the clay size shale, additional modifiers such as “sandy” and “cal-
fraction that was separated using Stoke’s Law method careous” are used to further describe the sample. The
(Head, 1984). This fraction was air-dried and glyco- modifiers are based on visual estimation for grain size
lated prior to analysis (Eslinger and Pevear, 1988). Min- and amount using petrographic microscopy. The
eral percents reported for X-ray analysis are considered modifying term is nonsilt and nonclay and generally
semiquantitative because of the following limitations. >10%. “Organic” is an additional modifying term. The
The X-ray diffraction data are a weight percent organic shales in this study have at least 7% total
measurement dependent on such factors as orientation organic carbon (TOC). Organic content was deter-
of the grains, thickness of the mount, and evenness of mined using a Leco Carbon Determinator.
the spread of mixed minerals, and are accurate within Scanning electron microscopy and the energy-
5 to 10% (Carroll, 1970). Even though X-ray analysis is dispersive X-ray analysis assisted in depositional fab-
a weight percent, it is treated here as a semiquantita- ric and bulk mineral determination, respectively. The
tive volume percent because the weight-to-volume descriptions that follow are based on composition of
conversions are roughly 1:1 for the common suite of the solid matrix. Note that the mineral percentages
minerals found in this study. Also, the layer silicates listed for each sample in Table 2 are for percent of the
identified by X-ray diffraction are inferred to be the matrix; the term “matrix” describes the nonorganic
only clay-size minerals. The histograms illustrated in solid material.
Figure 3 show the amount of fine-grained material that
other researchers have found in shales. Just as there is Petrographic Description
confusion about the meaning of the term “shale,” there
is also confusion concerning the amount of clay in Clayshale
shales. Clearly, the amount of clay identified in the sam- Three of the studied samples are in this category.
ples in this study is within the range of what others These samples have less than one-third of the grains
Seal Capacity of Nonsmectite Shale 35

larger than 5µ and are thinly to thickly laminated.


Quartz content ranges from 30 to 35%. Clay content
varies from 55 to 65%, with illite being the most abun-
dant and chlorite second in abundance. Considering
Chlorite Kaolinite TOC the major modifier, these samples are further

8
7
7
described as organic clayshales. These organic
clayshales are from the Antrim shale of the Michigan
Basin. The Antrim is Upper Devonian to Lower Mis-
sissippian and is correlative to the widespread Wood-

10
10
5
ford and Devonian shales of the central and eastern
Layer Silicates

United States, respectively. The organic clayshales are


samples S 1433, S 1461, and S 1537. These organic
clayshales are well indurated and contain 7–8% TOC.
10
15
10
15

10
15
10
20
5

5
5
Pyrolysis indicates the organic matter to be predomi-
nately type 1, oil prone, in the initial stages of oil gen-
eration. Figure 4 shows sample S 1461, which is typical
Illite

of the organic clayshales in this category.


40
40
35
30
30
40
25
25
40
40
45
45
Organic matter is commonly amorphous or com-
pressed planktonic algae palynomorphs (i.e., Tasman-
ities). Figure 4C shows the typical “deflated
Total
40
45
45
45
40
60
40
40
50
55
55
65

basketball” structure of the Tasmanities or Tasmanities-


like body commonly found in these shales (O’Brien
and Slatt, 1990).
Minor minerals include pyrite, dolomite, and pla-
Anhydrite

gioclase. Porosities of the samples are low; that is, less


than 4%. Silt-size quartz is generally less than 10%,
10

based on optical microscopy.


Claystone
This category contains shales that have less than
Hematite-
pyrite

one-third silt-size clasts and show no laminations.


Table 2. Mineral Composition and Organic Content of the Shale Samples.*

Two of the three samples in this category contain more


5
2

5
5

than 10% secondary calcite. The two carbonate-rich


samples are from the Fayetteville shale, Rogers
County, Oklahoma. Because of the high secondary car-
Sample ID Quartz Plagioclase K-feldspar Calcite Dolomite
Mineralogy X-ray (wt. %)

bonate content of 35% (Figure 5), these samples (CAT


10
10

20
5

5
5
5
5

1200 and CAT 1202) have the modifier “calcareous.”


The other sample in this group, RB 1856, is from the
Whitehorse Formation in the Rebecca Bounds-1 well,
Greeley County, Kansas. Sample RB 1856 has approxi-
mately 15% sand-size quartz; hence, the modifier
35

35
35
2

“sandy” is used. Clay amounts for these three clay-


*Mineralogy was determined from X-ray diffraction analysis.

stones range between 40 and 60%, with illite compris-


ing the largest portion of the clay.
Porosities are higher for the samples in this cate-
1

5
5

5
5

gory than those in the clayshale category. The porosi-


ties for the claystone samples range from 6.6 to 19.4%.
The samples with the lowest porosities have the high-
est amount of secondary calcite. Each of the three clay-
stones have 5% dolomite.
25
5

5
5

5
5
5

Mudstone
This category contains the most varied mineralogy
due to the differing amounts of feldspar, calcite, and
20
40
35
30
35
25
20
20
20
30
35
35

anhydrite. Samples RB 1511, RB 1532, RB 1713, RB


1761, and RB 1820 are mudstones based on silt-size
fraction, ranging between one-third and two-thirds,
CAT 1200
CAT 1202

and the lack of laminations. (Figure 6 illustrates sam-


RB 1511
RB 1532
RB 1713
RB 1761
RB 1820
RB 1856

M 9649
S 1433
S 1461
S 1537

ple RB 1713, which is typical of the sandy mudstones


in this category.) These samples are from the Morri-
son, Whitehorse, and Day Creek intervals, taken from
the Amoco Rebecca Bounds-1 well in Greeley County,
36 Krushin

Figure 2. Shale
classification used in
this study. The term
“shale” is only applicable
to the four textural/ silt
fraction groups containing
less than 2/3 silt fraction.
This figure is taken from
Lundegard and Samuels
(1980) and is used with
permission of the
publishers.

Kansas, and are predominantly red beds deposited in a mudstone with 30% feldspar. Sample RB 1511 contains
sabkha-like environment. Support for this interpretation sedimentary rock fragments and 35% calcite. This sam-
are the interbedded aeolian sands and evaporites, and ple is classified as a calcareous, lithic mudstone.
lack of marine fauna. These samples contain between 40 Additional nonclay minerals commonly comprising
and 45% clays. Samples RB 1532, RB 1713, and RB 1820 10% or less for this class of mudstones include calcite,
have about 15% very fine grained sand-size quartz dolomite, pyrite, and anhydrite. All the samples have
grains based on visual estimation and are modified disrupted bedding (see Figure 6A). The matrix of sam-
with the term “sandy.” Sample RB 1761 is a feldspathic ple RB 1820 has halite, which is recognized from the Cl

50
Patchett (1975)

40
O'Brien and Slatt (1990)

30
Number

20

10

0
0 10 20 30 40 50 60 70 80 90 100
Clay Percentage
Figure 3. Histograms of clay percentage found in 221 shale samples. The percentage is based on the solid bulk
material only (e.g., matrix). The data for Patchett (1975) represent 135 samples that range in age from
Ordovician to Tertiary. The data from O’Brien and Slatt (1990) are 86 samples ranging in age from Ordovician
through Quaternary. The percent clay for Patchett (1975) represents a weight percentage of <2 µm using
Stoke’s Law separation. The percent clay for the data from O’Brien and Slatt (1990) is layer silicate percentage
and is based on X-ray diffraction analysis.
Seal Capacity of Nonsmectite Shale 37

Figure 4. Sample S1461


of the Antrim Shale is
typical of the organic
clayshales studied.
(A) Thin-section slide
photograph shows thin
laminations and dark
color due to organic
content. Sample has 7%
TOC. (B) Thin-section
photomicrograph shows
fine-grained texture with
pyrite at top and no
visible porosity.
Crossed nicols. (C) SEM
photograph has flattened
Tasmanities (T) and
oriented clays (C).
Energy-dispersive X-ray
pattern shows Si, Mg,
Al, and Fe indicative of
quartz and chlorite. K
peak indicates illite.

peak from the energy-dispersive X-ray spectrum. and is visually estimated from log Pc vs. Sw plots (Fig-
These samples as a group have the highest porosity ure 1). Mercury injection porosimetry to 30,000 psia
values, ranging from 16 to 22.2%. (206,850 kPa) injection pressures were measured by K
and A Laboratories, Inc., Tulsa, Oklahoma, on the 12
Mudshale well-indurated shales by using unjacketed, oven-dried
Sample M 9649 is a mudshale from the Precambrian samples cut from core perpendicular to bedding (Figure
Upper Abu Mahara Formation, Oman. This thickly to 8). Maintaining sample microfabric and microstructure
lenticular-laminated sample contains 50% clays, with during the procedure is a concern. Two other drying
illite being the most abundant. This sample illustrates an techniques have been applied to weakly consolidated,
obvious problem with using sheet silicates as deter- fine-grained material to address this concern, namely
mined by X-ray diffraction as percent clays. Some of the critical-point drying and freeze-drying (Garcia-
illite is actually muscovite, larger than clay-grain size Bengochea, 1978; Delange and Lefebvre, 1984; Griffiths
(Figure 7). Less than 10% of the sample is sand-size and Joshi, 1991). Garcia-Bengochea (1978) concluded
quartz, based on visual estimation. Therefore, the shale that one of the major factors affecting the success of the
is solely modified by the term “dolomitic” because it is drying technique is the strength of the sample. There-
the only nonsilt, nonclay mineral >10%. Porosity is ~1%. fore, because these shales are well indurated, the oven-
drying technique is believed to not affect the pore throat
sizes. Additionally, the good relationships between
Mercury Injection Porosimetry
porosity measured by different oven-drying techniques
High-pressure mercury injection porosimetry (MIP) and between different permeability measurements sug-
determines the sealing shale displacement pressure, gest the samples retained their integrity on a microscale.
38 Krushin

Figure 5. Sample CAT 1202 is a calcareous claystone taken from the Fayetteville Shale, Rogers County,
Oklahoma. (A) Thin-section slide photograph shows lack of laminations. (B) Thin-section photomicrograph
illustrates fine-grained texture, lack of laminations, and no visible porosity. Black mineral is pyrite (P); white
mineral is sparry calcite. Crossed nicols. (C) SEM photograph shows uniform clay matrix and abundant rhom-
bohedral calcite (C). Energy-dispersive X-ray pattern shows large Ca peak indicative of calcite. Also, Si, Al,
Mg, and K indicative of quartz, illite, and chlorite. Fe peak may also indicate chlorite and/or pyrite.

Because the morphology of the clay minerals is the Hg displacement pressures converted to pore
sheetlike, the MIP-derived displacement pressure was throat size for the shale samples using a standard con-
converted to the distance between two parallel sheets tact angle of 140° and an interfacial tension of 485
and is referred to as tabular-shaped pore throats (Ayl- dynes/cm (Vavra et al., 1992). Diamond (1970) mea-
more and Quirk, 1967). The equation for conversion of sured contact angles of 147° for kaolinite and illite, but
mercury (Hg) displacement pressure to tabular- the standard value of 140° was used because these
shaped pore throats is (using the Washburn, 1921, shales are not pure kaolinite/illite samples.
equation modified from Schowalter, 1979): Although shale seal capacity is mainly a function of
the capillary properties of the rock, additional mea-
–2Cγ cos θ surements (e.g., porosity and permeability) were used
D tp = (2)
Pd to illustrate relationships among the various petro-
physical properties and to give the reader additional
confidence in the manner of analysis and interpreta-
where D tp = diameter of tabular-shaped pore (in tion. Both porosity and permeability were measured
nanometers); C = unit conversion constant of 1.45 × by two methods. Table 3 summarizes these data.
10 2; γ = interfacial tension of the Hg-air system (in Petrophysical measurements also include cation
dynes/cm); θ = contact angle expressed in degrees; and exchange capacity. The low values help confirm the
Pd = displacement pressure of Hg (psia). Table 3 shows nonsmectite composition of these shales.
Seal Capacity of Nonsmectite Shale 39

Figure 6. Sample RB 1713 of the Whitehorse Formation is typical of the sandy mudstones studied. (A) Thin-sec-
tion slide photograph shows poor sorting, disruptive bedding, and red color due to hematite staining. (B) Thin-
section photomicrograph shows very fine-grained sand and silt. Blue areas of epoxy are porosity. This sample has
20.2% Boyle’s Law derived porosity. Plane light. (C) SEM photograph shows cavities of plucked silt clasts (arrows)
and no orientation to clay fabric. Energy-dispersive X-ray pattern of Si, Al, and K indicates quartz and illite. Fe
peak represents hematite.

Porosity helium porosity, a value using the regression shown in


Figure 9 is calculated.
Boyle’s Law helium porosity was measured on
humidity-dried and vacuum-dried, vertically oriented
Permeability
plugs cut from conventional core. The humidity-dried
samples were dried for 48 hr at 130°F (54°C) and 35% rel- Permeabilities were measured for both air and
ative humidity. Samples were humidity dried because of water. K and A Laboratories, Inc., Tulsa, Oklahoma,
concern that complete drying would cause the samples measured air permeability as well as the Boyle’s Law
to fracture from desiccation. According to Bush and porosity on the humidity-dried samples. Core Labora-
Jenkins (1970), humidity drying should leave about two tories Inc., Bakersfield, California, measured the water
molecular layers of water on the clay minerals and thus permeability along with the Boyle’s Law porosity for
prevent complete desiccation, although during the vacuum-dried samples.
study dessication was not a problem. Additionally, Permeabilities to both air and water were measured
Boyle’s Law helium porosity for 9 of the 12 samples was on 1-in. (2.54-cm) diameter, vertically oriented samples
measured on companion samples that were dried under cut from conventional core. Each test used companion
vacuum and at 122°F (50°C) (Table 3). Predictably, this samples cut from the same depth. The air permeability
results in higher porosity values than those porosities was measured on the humidity-dried samples, using
measured by humidity drying (Figure 9). Interpretation humidified air as the permeant. The resulting air per-
of these data uses the vacuum-derived porosity. For the meabilities were not corrected for gas slippage. These
three samples that do not have measured vacuum-dried tests were done with a hydrostatic confining pressure
40 Krushin

Figure 7. Sample M 9649 of the sandy, dolomitic mudshale of the Upper Abu Maharah Formation is the only
mudshale studied. (A) Thin-section slide photograph shows excellent thin to thick laminations. (B) Thin-sec-
tion photomicrograph shows interval of disrupted bedding, very fine-grained sand and silt and mica (M).
Plane light. (C) SEM photograph shows mica (M). Energy dispersive X-ray pattern shows peaks of Si, Al, Mg,
K, and Fe. These indicate quartz, illite, and biotite.

of 400 psia (2758 kPa). The resulting air permeabilities using a 50 psia (345 kPa) pressure pulse. The two per-
range from 0.3 to 5.9 µd (Table 3). meabilities relate extremely well (Figure 10), with the
The pressure-pulse, transient method was also water permeability having the higher value. Jones
used because of documented reproducibility in this and Owens (1979) document the opposite relation-
permeability range (Amaefule et al., 1986). This ship for low-permeability sandstones. The relation-
method uses a small water-pressure pulse upstream ship documented here for shales is in spite of the
of the plug. The pressure decay characteristics are water permeability being measured at a higher con-
dependent on the permeability. The samples were fining pressure and the air permeabilities not being
vacuum-pressure saturated using a standard labora- corrected for gas slippage. The effects seen here are
tory synthetic solution of 50,000 ppm NaCl and 5000 attributed to two causes: (1) the salinity of the perme-
ppm CaCl2. This solution was also used as the perme- ating fluid decreasing the distance of the diffuse dou-
ant. The measured water permeabilities range in val- ble layer on the clay surfaces, thereby allowing for
ues of 4.2–8.3 µd. The brine pressure decay more open fluid pathways (Jose et al., 1989), or (2) a
permeabilities were run with an effective hydrostatic relative permeability effect. The humidity drying
confining pressure of 1000 psia (6895 kPa) and by introduced a water phase and may have caused a
Seal Capacity of Nonsmectite Shale 41

100000

10000
Mercury Injection Pressure (psia)

(B) Claystones
(A) Mudstones
1000
RB 1856 CAT 1200 CAT 1202
RB 1511 RB 1532 RB 1713
RB 1761 RB 1820

100

10

1
70 60 50 40 30 20 10 0 70 60 50 40 30 20 10 0
Mercury Saturation (% Pore Volume) Mercury Saturation (% Pore Volume)

(C) Mudshale and Clayshales

M 9649 S 1433
S 1461 S 1537

70 60 50 40 30 20 10 0
Mercury Saturation (% Pore Volume)

Figure 8. High-pressure mercury injection porosimetry (MIP) profiles used for calculating seal capacity. The
pressures are converted to pore throat sizes using equation 2 in the text. (A) contains the MIP profiles for the
mudstones, (B) contains the MIP profiles for the claystones, and (C) contains the MIP profiles for the one
mudshale sample and the clayshale samples. Mudshale sample M 9649 has a relatively large part of its pore
volume attributed to surface effects. This is not surprising because the sample only has 1.2% porosity.

decrease in the absolute air permeability (J.G. Patch- Nonorganic Shales


ett, personal communication, 1991).
Table 4 shows correlation coefficients between the
shale displacement throat size and many of the petro-
DISCUSSION graphic and petrophysical measurements. The correla-
tion coefficients in Table 4 illustrate the usefulness of
As mentioned earlier and illustrated by equation 1, interpreting the factors that control seal capacity sepa-
the shale displacement pressure greatly controls the rately for nonorganic shales. The best correlation for
seal capacity. The Hg displacement pressure was visu- displacement pore throat size, for nonorganic shales, is
ally estimated from the MIP profiles (Figure 8). Dis- with quartz content of the matrix. Surprisingly, quartz
cussion of the organic clayshales is separate from the content by itself relates better to displacement pore
nonorganic shales because hydrocarbon generation throat than either bulk volume quartz (used here as
from the organic component is interpreted to have an quartz percent × [1–porosity]) or bulk volume quartz
effect on the displacement pore throat size. and porosity. Figure 11 is a plot of displacement pore
42 Krushin

which indicates halite. This anomalously large dis-

Displacement
placement pore throat size is attributed to partial dis-

Pore Throat
Size (nm)
solution of halite in the core, which results in large
pore throats. Therefore, the discussion and predictive

11
37
31
15
60
14
7
8
7
72
24
15
regression (Figure 11) do not include this sample.
The displacement pore throat sizes for the nonor-
ganic shales in this study range from 7 to 37 nm. Kat-
sube et al. (1991) document pore throat sizes for
deeply buried shales with less than 9% porosity from
Hg-Air System
Entry Pressure

the Ventura Gas Field, offshore Nova Scotia, Canada.


Plotting the MIP data from Katsube et al. (1991) with
(psia)
9500
2900
3500
7000
1800
7500
15,000
13,000
15,000
1500
4500
7000
mercury saturations vs. injection pressure does not
yield obvious displacement pressures (there is no
plateau that can be extended to 0 Hg saturation). Also,
displacement pore throat sizes were not reported for
their data. The geometric mean of the pore throat sizes
Permeability

for their ten shale samples range between 9 and 16 nm.


7.9**

6.0**

8.9**

5.5**
Water

The pore throat sizes presented in Katsube et al. (1991)


(µd)

7.3
5.7

8.3

5.5
4.2

4.8
4.9
5.3 certainly support the smaller displacement pore
Petrophysical Measurements

throats measured here. Figure 11 shows that the


nonorganic shales with the smallest displacement pore
throats are calcareous or dolomitic mudshales, mud-
stones, and claystones. Also, Figure 11 shows that
Vacuum Dried Permeability

sandy mudstones contain the largest pore throats,


4.7
3.5
3.2
2.5
4.7
5.9
2.3
0.3
1.8
1.2
0.9
1.0
(µd)

with sandy claystones and feldspathic mudstones hav-


Air

ing intermediate values.


Heling (1970) studied both pore throat size distribu-
Table 3. Petrophysical Measurements of the 12 Shale Samples.

tions and fabrics of Tertiary shales, as a function of


depth, from the Rhinegraben of southwestern Ger-
Porosity—

many. He found that the coarser grained components


19.0**

22.4**

19.4**

of shale (e.g., quartz) shield the larger pore throats and


(%)

12.6
20.2

20.8

6.6
7.9
1.2
3.7
3.6
3.9

inhibit compaction-generated fabric orientation,


thereby preserving the larger pore throats as the shales
are compacted. This supports the findings that show
the largest displacement pore throats are found in
shales that contain the coarsest grained components,
Humidity Dried

for example, sandy mudstones, although it is not


**Calculated from regressions; see text and figures.
Porosity—

known at what depth this shielding begins to override


(%)
13.5
10.1
14.4
16.0
13.7
13.8
5.0
3.9
0.8
1.2
1.4
3.4

the effects of compaction. Because the largest pore


throats in this study are associated with the coarsest
grained shales, it appears that quartz shielding is a
dominant factor. Not represented in this study are
poorly compacted clay-rich shales.
Exchange
Sample ID Capacity

Additional effects of compaction on shale seal


Cation

capacity are inferred from MIP performed on clayey


6.7
6.1
3.8
6.4
2.0
4.8
4.4
7.7
1.1
0.1
1.5
*

soils and clays that were subjected to increasing uni-


axial overburden strain in the laboratory. These stud-
ies provide useful analogies for the effects of
CAT 1200
CAT 1202

*Not measured.
RB 1511
RB 1532
RB 1713
RB 1761
RB 1820
RB 1856

compaction on shales. The compaction studies of


M 9649
S 1433
S 1461
S 1537

Sridharan et al. (1971), Delange and Lefebvre (1984),


Griffiths and Joshi (1989), and Lapierre et al. (1990)
show that the loss of pore volume with increasing
states of compaction is at the expense of the largest
pore throats, and this loss is nonrecoverable and
inelastic. Therefore, clay-rich shales should attain
throat size vs. the amount of quartz in the matrix. The their highest degree of seal capacity due to the great-
sample having 35% quartz and 60 nm displacement est depth of burial that they have been exposed to.
pore throat size does not appear to follow the trend This degree of seal capacity will not be decreased
with the other samples. An examination of the miner- upon uplift, since unloading involves the elastically
alogy of this sample, RB 1820, shows why it is differ- recovered deformation (Mitchell, 1976). Disregarding
ent. The energy-dispersive X-ray pattern shows Cl, the effects of secondary cements, the shales with the
Seal Capacity of Nonsmectite Shale 43

Figure 9. This chart shows


the relationship between
porosity measurements on
samples that were dried two
different ways. Boyle’s Law
helium porosity for shale
samples that were humidity
dried for 48 hr at 130°F
(54°C) and 35% relative
humidity are compared
to companion samples that
were dried at 122°F (50°C)
under vacuum. The
vacuum-derived porosity
is used for the discussion
in the text.

highest compaction-derived seal capacity would be shales. These are in the range of 15–72 nm, with
those with the least coarse-grained fabric; that is, the porosities of ~4%. The large displacement pore throats
clayshales and claystones. for organic shales are clearly in contrast to nonorganic,
Quantifying seal capacity involves both the proper- calcareous claystones, with porosities ranging from 6.6
ties of the rock (e.g., displacement pore throat size) and to 7.9%. These nonorganic calcareous claystones have
the properties of the fluids, buoyancy pressure, contact displacement pore throat sizes of <10 nm. Luffel and
angle, and interfacial tension. Zieglar (1992) studied Guidy (1989) report a maximum displacement pore
hydrocarbon column sizes to determine if seal capacity throat size of 100 nm for organic Devonian shales,
could be a limiting factor in the size of the columns. He with porosities ranging from 2 to 7%. Considering
concluded, based on tenuous shale capillary data, that SEM photomicrographs, thin-section microscopy, and
only the largest 10% of unfaulted gas columns in the the low porosity values, these large displacement pore
Rocky Mountain area have generated enough buoyancy throats do not appear to be associated with quartz
pressure to overcome the shale seal capacity. His 10% shielding or low levels of compaction. The relatively
largest percentile gas columns are about 1000 ft (305 m), large pore throats are attributed to hydrocarbon gen-
and the largest gas column is 1800 ft (549 m). Using eration effects.
Zieglar’s fluid properties, these gas columns convert to Scanning electron microscope microfabric studies
displacement pore throat size. His average conversion of organic shales by O’Brien et al. (1994) document
factor for gas buoyancy gradient of 0.35 psia/ft (7.91 microchannels in organic-rich shales, some possibly
kPa/m) results in 350 psia (2413 kPa) and 630 psia (4344 with oil globules in them. Therefore, organic shales
kPa), respectively, for buoyancy-generated gas pres- that are generating hydrocarbons are actually creating
sures. Substituting into equation 2 the hydrocarbon/ larger pore throats, associated with microchannel,
brine contact angle of 0° for a water-wet system, gas- micotunnel, or microfracture formation (O’Brien, 1989;
water interfacial tension of 49 dynes/cm, and the afore- O’Brien et al., 1992). These large pore throats are
mentioned pressures results in a displacement formed by the volume reduction associated with
tabular-pore throat size range of 23–41 nm. This is very hydrocarbon generation. No attempt is made to relate
close to the largest pore throats measured in this study size of displacement pore throat to column height
and supports Zieglar’s (1992) conclusions. Shorter gas because of the uncertainty of what contact angle to
columns would not generate enough pressure to over- use because of the possibility for nonwater-wet in-situ
come the shale’s displacement pressure. conditions.

Organic Shales
CONCLUSIONS
A review of Table 3 shows that organic shales have
large displacement pore throats relative to their very Nonsmectite shales were studied by using laboratory-
low porosity values, when compared to nonorganic derived measurements to determine how well they
44 Krushin

9 Figure 10. This chart shows


the relationship between
8 air permeability and water
Shale Permeability permeability. The air
Water Permeability (microdarcies)

7 permeability was measured


using humidified air and
6 is not corrected for gas
slippage. The water
5 Y = 0.83X + 3.98 permeability was measured
2
R = 0.88 using the pressure-pulse
4 technique with a standard
laboratory solution as the
3 permeant.

0
0 1 2 3 4 5

Air Permeability (microdarcies)

can seal hydrocarbons. The maximum size of the nonorganic shales that have abundant calcare-
hydrocarbon column a shale can seal is limited by the ous or dolomitic cement, or are clay rich, have
shale’s largest interconnected pore throat. This capil- displacement pore throats of <15 nm and would
lary sealing property of nonsmectite shales, referred to be excellent capillary seals. Nonorganic shales
as seal capacity, is determined by identifying the dis- with abundant coarse-grained material (e.g.,
placement pressure on an MIP plot. Plotting mercury sandy mudstones) contain the largest measured
saturations vs. mercury injection pressure on a linear pore throats. These are in the range of 30–40 nm
log-scale plot easily identifies the displacement pres- and could control the size of long gas columns
sure. Maximum hydrocarbon columns are estimated (~1000 ft; 305 m).
by converting MIP-derived displacement pressure to 3. Seal capacity can be predicted by quantifying
reservoir conditions. The major conclusions of this the amount of quartz in the shale matrix. Quartz
work are: content measured by using X-ray diffraction
analysis predicts displacement pore throat size
1. Nonsmectite shales are divided into two major by the following equation: displacement pore
categories with regard to factors controlling dis- throat size (nm) = 1.4 (% quartz in matrix) – 20.5.
placement pore throat size. These categories are 4. Compaction also can affect shale seal capacity.
nonorganic and organic shales. Maximum seal capacity for clay-rich shales is a
2. Petrology is a major control of displacement result of the maximum depth of burial and is
pore throat size in nonorganic shales. Classify- maintained upon uplift. Compaction was not
ing nonorganic shales with regard to amount of studied as part of this work, but compaction
silt/clay ratios, laminations, and major nonclay, effects on shale seal capacity are inferred from
nonsilt minerals allows for a preliminary esti- compaction studies performed on clays and
mation of seal capacity. The low-porosity, clay-rich soils. These compaction studies on

Table 4. Correlation Coefficients Between Seal Capacity* and Selected Petrophysical Measurements.**

Porosity +
Kw Porosity Quartz in Bulk Volume Bulk Volume
Seal Capacity (µd) (%) Matrix (%) Quartz (%) Quartz (%)
Nonorganic Shales Only 0.47 0.47 0.82 0.75 0.75
All Samples 0.10 0.07 0.57 0.46 0.33

*i.e., displacement pore throat size (nm).


**The nonorganic shales relate better (have higher correlation coefficients) to the petrophysical/petrographic properties than the group con-
taining all the shale samples.
Seal Capacity of Nonsmectite Shale 45

Figure 11. This chart illustrates the relationship between quartz content of the matrix, obtained from X-ray
diffraction analysis, and the displacement pore throat size, from MIP. Data are for nonorganic shales only.
The sample with the largest displacement pore throat size contains halite interpreted to have partially dis-
solved prior to analysis, perhaps in storage. Thus, the large-displacement pore throat is an artifact, not repre-
sentative of in-situ conditions, and is not included in the final interpretation or the regression.

fine-grained material show that compaction is at of Amoco for X-ray diffraction analysis and interpreta-
the expense of the largest pore throats, thereby tion. Thanks go to J. Martin-Elliott for assistance in
increasing seal capacity with increasing degree of drafting some of the figures and to F.M. Krushin for
compaction, although it is not known at what typing the manuscript. The reviews of Leta K. Smith
depth the clay-rich shale has such small pore and Norman R. Morrow greatly improved this manu-
throats that seal capacity is no longer a reasonable script. The coordinating efforts of Katherine I.
exploration concern. This question remains unan- Kirkaldie are also appreciated. I would also like to
swered because shallowly buried shales in active thank the management of Amoco Exploration and
depositional settings are smectite rich, and MIP is Production for granting me permission to publish this
not an appropriate analytical technique. Shales paper.
with abundant coarse-grained components (e.g.,
sandy mudstones) have quartz grains shielding
the larger pore throats and inhibiting compaction- REFERENCES CITED
generated fabrics. Thus, increasing compaction
would not increase seal capacity for a sandy mud- Amaefule, J.O., K. Wolfe, J.D. Walls, A.O. Ajufo, and E.
stone, as it would for a claystone. Questions con- Peterson, 1986, Laboratory determination of effec-
cerning the complete relationship among tive liquid permeability in low quality reservoir
displacement pore throat size, quartz content, and rocks by the pulse decay technique: Society of
degree of compaction remain unanswered. Petroleum Engineers 56th Annual California
5. Matrix volume reduction associated with hydro- Regional Meeting, p. 493–499.
carbon generation causes relatively large dis- Aylmore, L.A., and J.P. Quirk, 1967, The micropore
placement pore throats in organic shales. size distribution of clay mineral systems: Journal of
Soil Science, v. 18, p. 1–17.
Berg, R.R., 1975, Capillary pressure in stratigraphic
ACKNOWLEDGMENTS traps: AAPG Bulletin, v. 59, p. 939–956.
Bush, D.C., and R.E. Jenkins, 1970, Proper hydration of
This paper resulted from a Amoco Petrophysics clays for rock property determinations: Journal of
Class XXIV project. I would like to thank G.R. Powers Petroleum Technology, v. 22, p. 80–84.
46 Krushin

Carroll, D., 1970, Clay minerals: A guide to their X-ray Canadian Geotechnical Journal, v. 27, p. 761–773.
identification: U.S. Geological Survey, Special Paper Luffel, D.L., and F.K. Guidy, 1989, Reservoir rock
26, 79 p. properties of Devonian shale from core and log
Cullity, B.D., 1978, Elements of X-ray diffraction, 2nd analysis: The Society of Core Analysts Annual
ed.: Reading, Addison-Wesley Publishing Co., 553 p. Technical Conference, Paper No. 8910, 15 p.
Delange, P., and G. Lefebvre, 1984, Study of the struc- Lundegard, P.D., and N.O. Samuels, 1980, Field classi-
ture of a sensitive Champlain clay and of its evolu- fication of fine-grained sedimentary rocks: Journal
tion during consolidation: Canadian Geotechnical of Sedimentary Petrology, v. 50, p. 781–786.
Journal, v. 21, p. 21–35. Mitchell, J.K., 1976, Fundamentals of soil behavior:
Diamond, S., 1970, Pore size distribution in clays: Clay New York, John Wiley and Sons, 422 p.
and Clay Mineralogy, v. 18, p. 7–23. Murray, R.S., and J.P. Quirk, 1992, Comment on “Clay
Downey, M.W., 1984, Evaluating seals for hydro- fabric response to consolidation” by F.J. Griffiths
carbon accumulations: AAPG Bulletin, v. 68, and R.C. Joshi: Applied Clay Science, v. 6,
p. 1752–1763. p. 411–413.
Eslinger, E., and D. Pevear, 1988, Clay minerals for O’Brien, N., 1989, Origin of lamination in Middle and
petroleum geologists and engineers: SEPM Short Upper Devonian Black Shales, New York State:
Course 22, 311 p. Northeastern Geology, v. 11, p. 159–165.
Garcia-Bengochea, I., 1978, The relationship between O’Brien, N.R., and R.M. Slatt, 1990, Argillaceous rock
permeability and pore size distribution of com- atlas: New York, Springer-Verlag, 141 p.
pacted clayey silts—interim report: Joint Highway O’Brien, N.R., R.M. Slatt, and J. Senftle, 1992, The sig-
Research Project 78–4, 179 p. nificance of oil shale fabric and primary hydrocar-
Griffiths, F.J., and R.C. Joshi, 1989, Change in pore size bon migration: Eastern Oil Shale Symposium,
distribution due to consolidation of clays: Géotech- November 17–20, Lexington, Kentucky, p. 294–297.
nique, v. 39, p. 159–167. O’Brien, N.R., G.D. Thyne, and R.M. Slatt, 1994,
Griffiths, F.J., and R.C. Joshi, 1991, Change in pore size “Micro-channels” in organic-rich shales and
distribution owing to secondary consolidation of mudrocks as primary hydrocarbon migration path-
clays: Canadian Geotechnical Journal, v. 28, ways: Possible example from the Monterey Forma-
p. 20–24. tion, California (abs.): AAPG Annual Convention,
Head, K.H., 1984, Manual of soil laboratory testing, p. 227.
volume 1: Soil classification and compaction tests: Patchett, J.G., 1975, An investigation of shale conduc-
London, ELE International Limited, 339 p. tivity: SPWLA 16th Logging Symposium, Paper U,
Heling, D., 1970, Micro-fabrics of shales and their 41 p.
rearrangement by compaction: Sedimentology, Potter, P.E., J.B. Maynard, and W.A. Pryor, 1980, Sedi-
v. 15, p. 247–260. mentology of shale: New York, Springer-Verlag,
Jennings, J.B., 1987, Capillary pressure techniques: 306 p.
Application to exploration and development geol- Rudd, N., and G.N. Pandey, 1973, Threshold pressure
ogy: AAPG Bulletin, v. 71, p. 1196–1209. profiling by continuous injection: AIME-SPE Paper
Jones, E.A., and H.F. Wang, 1981, Ultrasonic velocities 4597, 7 p.
in Cretaceous shales from the Williston Basin: Geo- Schowalter, T.T., 1979, Secondary hydrocarbon migra-
physics, v. 46, p. 288–297. tion and entrapment: AAPG Bulletin, v. 63,
Jones, F.O., and W.W. Owens, 1979, A laboratory p. 723–760.
study of low permeability gas sands: Society of Shaw, D.B., and C.E. Weaver, 1965, The mineralogical
Petroleum Engineers Symposium on Low Perme- composition of shales: Journal of Sedimentary
ability Gas Reservoirs, Paper No. 7551, 10 p. Petrology, v. 35, p. 213–222.
Jose, U.V., S.T. Bhat, and B.U. Nayak, 1989, Influence Smith, D.A., 1966, Theoretical considerations of seal-
of salinity on permeability characteristics of marine ing and non-sealing faults: AAPG Bulletin, v. 50,
sediments: Marine Geology, v. 8, p. 249–258. p. 363–374.
Kaldi, J.D., and C.D. Atkinson, 1994, Seal potential of Sridharan, A., A. Hschaeffl, and S. Diamond, 1971,
the Talang Akar Formation, BZZ area, offshore Pore size distribution studies: Proceedings of ASCE
northwest Java, Indonesia (abs.): AAPG Annual Journal of Soil Mechanics and Foundation Division,
Convention, p. 183. May, p. 771–787.
Katsube, T.J., B.S. Mudford, and M.E. Best, 1991, Petro- Terry, R.D., and G.V. Chilingar, 1955, Summary of
physical characteristics of shales from the Scotian “Concerning some additional aids in studying sedi-
shelf: Geophysics, v. 56, p. 1681–1689. mentary formations” by M.S. Shvetson: Journal of
Krooss, B.M., 1987, Experimental investigation of the Sedimentary Petrology, v. 25, p. 229–234.
diffusion of low molecular weight hydrocarbons in Tourtelot, H.A., 1960, Original and use of the word
sedimentary rocks, in B. Doligez, ed., Migration of “shale”: American Journal of Science (Bradley Vol-
hydrocarbons in sedimentary basins: 2nd IFP ume), v. 258–A, p. 335–343.
Exploration Research Conference, 681 p. Ulmishek, G.F., 1988, Types of seals related to migra-
Lapierre, C., S. Leroueil, and J. Locat, 1990, Mercury tion and entrapment of hydrocarbons, in L.B.
intrusion and permeability of Louisville clay: Magoon, ed., Petroleum systems of the United
Seal Capacity of Nonsmectite Shale 47

States: U.S. Geological Survey Bulletin, v. 1870, 68 p. Watts, N.L., 1987, Theoretical aspects of cap-rock and
Vavra, C.L., J.G. Kaldi, and R.M. Sneider, 1992, Geo- fault seals for single- and two-phase hydrocarbon
logical applications of capillary pressure: A review: columns: Marine and Petroleum Geology, v. 4,
AAPG Bulletin, v. 76, p. 840–850. p. 274–307.
Wasburn, E.W., 1921, Note on a method of determining Zieglar, D.E., 1992, Hydrocarbon columns, buoyancy
the distribution of pore sizes in a porous material: pressures, and seal efficiency: Comparisons of oil
Proceedings, National Academy of Sciences, v. 7, and gas accumulations in California and the Rocky
p. 115–116. Mountain area: AAPG Bulletin, v. 76, p. 501–508.
Niemann, J.C., and M.R. Krolow, 1997, Delineation of a
pressure fault seal from shale resistivities, in R.C.
Surdam, ed., Seals, traps, and the petroleum sys-
tem: AAPG Memoir 67, p. 49–55.

Chapter 4

Delineation of a Pressure Fault Seal


From Shale Resistivities
James C. Niemann
Mark R. Krolow
Chevron USA Production Co.
New Orleans, Louisiana, U.S.A.

ABSTRACT
Pressure differentials between fault blocks are definitive evidence of a fluid
seal across faults. A method is proposed for using log-derived and calculated
pressure gradient values to correlate and map pressures on cross sections to
delineate pressure differentials across fault zones. These values were calculated
and mapped along a known hydrocarbon sealing fault inferred to be a pressure
seal in a field from the Gulf of Mexico. The results indicate that this method can
generate a detailed picture of the fluid and pressure history to aid in the docu-
mentation and prediction of sealing or leaking relationships along fault zones.

INTRODUCTION However, where abnormal pressure primarily is


caused by compaction disequilibrium, pressures can
Fluid pressure differentials have been documented be estimated from wireline log-derived shale porosity
along many hydrocarbon sealing fault closures in the indicators, such as resistivity or conductivity, and
Gulf Coast and Gulf of Mexico Tertiary fields (Myers, interval transit time or velocity measurements. These
1968; Fowler, 1970; McCullough, 1990). Pressure dif- values can then be correlated between wells along
ferences between fault blocks are not the only indica- structural and stratigraphic datums to evaluate pres-
tions that a fault is sealing. The presence of tilted fluid sure differences along fault zones with unknown seal-
contacts or freshwater zones indicates hydrodynamic ing or leaking relationships. An offshore Gulf of
flow across a leaking fault (still capable of supporting Mexico gas field with an inferred hydrocarbon sealing
hydrocarbon columns) such as in the Bastian Bay Field fault was chosen to determine if log-derived shale
in southeast Louisiana (Myers, 1968). However, in the properties can be used to calculate pressures and
absence of such criteria, pressure differences between delineate pressure differences across faults; this would
fault blocks are strong evidence for a fluid sealing fault indicate if a fault was sealing or leaking.
that retards the equilibration of pressures and
supports hydrocarbon accumulations. Ideally, pres-
sure differences are determined from downhole SEALING FAULT DESCRIPTION
pressure measurements in wells either with a wireline
formation testing tool or a downhole pressure bomb. Figure 1 is a structure map on a Miocene age gas
Unfortunately, in some parts of the world, downhole reservoir sand in a field located in the Gulf of Mexico,
pressure data are not commonly available and are not offshore Louisiana. The Miocene section is highly
consistently measured by many oil and gas operators. expanded downthrown to the “A” fault, creating a
49
50 Niemann and Krolow

A' 7

0'
-1450

00'
0
-1450

5
-14
"A" FAULT

0'
00
-14

'

-14
000

00
0'
-15

'
00
40
'

-1
000
-16 5

0' "B" FAULT


00
4 -13

N
3

GAS RESERVOIR
1
A
OFFSHORE GULF OF MEXICO GAS FIELD
0 1 2 3 KM CHEVRON USA PROD. CO.
13300' SD STRUCTURE
0 1 2 MILES

Figure 1. Structure map at the 13,300 ft sand horizon with >200 m (700 ft) of
gas column trapped upthrown to the “B” fault. On N-S cross section A–A′,
note that wells 8, 9, and 10 shown on Figure 2 are out of the map view to the
north. SD = sand.

large rollover anticline typical of many Tertiary Gulf Sealing faults are defined by differences in hydro-
of Mexico producing fields. The structure has been carbon-water contacts in reservoirs juxtaposed across
bisected by another expansion fault (“B”), creating a the fault zones that cannot be accounted for by errors
steep asymmetric anticline with the southeast limb in depth measurement or by capillary pressure varia-
collapsed by fault expansion as shown by cross section tions between the reservoirs. Reservoir juxtaposition is
A–A′ in Figure 2. Gas is trapped primarily in several best documented by construction of a fault plane pro-
stacked sandstones on the upthrown side of the “B” file or “Allan Map” (Allan, 1989). Because of variable
fault, but a few gas reservoirs are present on the changes in the fault displacement, bifurcation of the
downthrown side of the “B” fault as well. Gas column fault (not shown on the structure map) and the large
heights range up to 800 ft (250 m), which is atypically expansion of the section downthrown, a reliable Allan
large for fault traps in the Gulf of Mexico. Map projection, could not be constructed on the “B”
Pressure Fault Seal 51

A A' A"
1 2 3 4 5 6 7 8 9 10

2.5 KM-

TSNP

LT
3 KM -
-10000'
FAU
"B"

3.5 KM-

2000
TSNP 4 KM -
13

T
30

UL
0'

FA
SD

"
"A
2000 0 FEET 4.5 KM-
MA

-15000'
PH
OR

OFFSHORE GULF OF MEXICO GAS FIELD


IZO

CHEVRON USA PRODUCTION CO.


N

N-S STRUCTURAL CROSS-SECTION


SANDSTONE
GAS RESERVOIR

Figure 2. N-S structural cross section A–A′ showing stacked gas reservoirs trapped upthrown and downthrown
to the “B” fault. SD = sand, TSNP = top of supernormal pressure.

fault. However, the cross section shows some water- the fault. Therefore, to map out the pressure differen-
filled sandstone to gas-filled sandstone juxtapositions tial across the entire fault surface in two dimensions,
across the “B” fault, and the changing displacement pressures were calculated from well log–derived shale
along the fault strike suggests that the upthrown gas porosity properties.
reservoirs are juxtaposed against water-filled sand-
stones at some or many points along the fault. Further-
more, the top of the transition zone from a normal, DETERMINATION OF PRESSURE
hydrostatic pressure environment to a supernormal FROM WELL LOGS
pressure environment, determined from well logs and
drilling mudweights, is displaced ~2000–3000 ft The determination of pressures from well logs is
(600–900 m) downthrown across the “B” fault, as based on the work of Athy (1930), Terzaghi and Peck
shown by the line labeled TSNP (top of supernormal (1948), and Hubbert and Rubey (1959), who established
pressure) on Figure 2. The apparent juxtaposition of a that porosity is an inverse exponential function of the
lower pressure section against a higher pressure sec- effective stress, which is the difference between the
tion and the possible juxtaposition of gas reservoirs overburden or lithostatic stress and the fluid pressure.
against water-filled sandstones, plus the anomalously If the change in fluid pressure with depth remains con-
high gas columns in some of the reservoirs, suggests stant (hydrostatic gradient), the effective stress will
that the fault itself is sealing. The displacement of the increase, because the overburden stress increases at a
transition to supernormal pressure indicates a pres- greater rate than the hydrostatic pressure change. Con-
sure differential across the “B” fault, but it does not sequently, porosity will decrease exponentially with
document the dynamic changes that may be occurring depth as the effective stress increases. Under supernor-
on either side of the fault and whether the pressure mal conditions where the fluid pressure gradient is
differential continues with depth. There are only two much higher than hydrostatic, the porosity will
available initial bottom-hole pressure measurements decrease at a lower rate or may actually increase locally
from wireline formation testing tools, and both of with depth. Therefore, by knowing or estimating the
those measurements came from the upthrown side of overburden and hydrostatic pressure gradients, the
52 Niemann and Krolow

Figure 3. Logarithmic shale


DEPTH RESISTIVITY (OHM-M) resistivity vs. depth plot for
0.5 1.0 5.0 10.0 the #4 well overlain by a
series of isopressure
gradient lines in terms of
PPG (lb/gal) mudweight.
The gradient overlay was
2.5 KM calculated from a modified
form of Eaton’s equation:
PRESSURE FPG = 19.2 – (19.2 – 9.0)
GRADIENT (Ro/RN)1.5, where Ro =
ppg (kPa/m) observed shale resistivity
and RN = shale resistivity
11-12 (13.0-14.2) along the normal
3.0
-10000' 12-13 (14.2-15.3) compaction trend line.
Along the right side of the
11-12 (13.0-14.2)
diagram is a partial log of
NOR

10-11 (11.8-13.0) the pressure zones in the


well as determined from the
MAL

9-10 (10.6-11.8)
10-11 (11.8-13.0) shale resistivities.
3.5
COM

11-12 (13.0-14.2)
12-13 (14.2-15.3)
PAC

"B"
13-14 (15.3-16.5)
TION

14-15 (16.5-17)
TRE

4.0 MAP HORIZON


ND L
10 p
INE
15.0

pg (
- 9.0
ppg

11.8

4.5
ppg
(17.6

-15000'
kPa/
(10.6
m)
kPa/

k
Pa/m
m)

fluid pressure can be back-calculated by deriving shale pressure gradient; Ro = observed shale porosity indica-
porosities from well log electric or acoustic measure- tor (resistivity/conductivity or interval transit
ments. Many variations on the calculation of pore pres- time/velocity) at some depth; RN = extrapolated shale
sure from log data have been published (Hottman and porosity indicator at that depth if the section were nor-
Johnson, 1965; MacGregor, 1965; Hamouz and Mueller, mally compacting; and X = an empirically derived
1984; Martinez and King, 1987; Weakley, 1989, 1990). The exponential value. The technique works only in areas
method of Eaton (1975) was chosen for this study where supernormal pressure primarily is caused by
because it is generally accepted for use in the industry, the inability of shales to compact normally due to sed-
can be used in a variety of geologic settings, and is adapt- iment loading and formation of vertical fluid seals.
able to studies with either an abundance or a sparsity of Where overpressuring is caused by other factors (for
geologic data. A general form of the Eaton equation is: example, hydrocarbon generation or by tectonic
stresses), the calculations will be in error. When using
X
shale resistivity or conductivity data, the results may
FPG = OBG – (OBG – HPG ) o R 
R be affected adversely by changes in formation water
N
(1) resistivity, borehole rugosity, shale anisotropy result-
ing from high angles between the borehole and bed-
ding, and pervasively silty or diagenetically tight,
where FPG = formation pressure gradient; OBG = calcareous shale sections. Interval transit time data
overburden pressure gradient; HPG = hydrostatic has the advantage of not being affected by changes in
A A' A"
1 2 3 4 5 6 7 8 9 10

2.5 KM-

LT
FAU
TSNP

"B"
3 KM -
-10000'-

3.5 KM-

PRESSURE GRADIENT
4 KM -
<9.0 PPG (<10.6 kPa/m)
ULT 9.0-10.0 (10.6-11.8)
FA 10.0-11.0 (11.8-13.0)
" 11.0-12.0 (13.0-14.2)
"A
12.0-13.0 (14.2-15.3)
4.5 KM-
13.0-14.0 (15.3-16.5)
-15000'-
14.0-15.0 (16.5-17.7)
15.0-16.0 (17.7-18.9)
16.0-17.0 (18.9- 20.1)
17.0-18.0 (20.1- 21.2)

2000 OFFSHORE GULF OF MEXICO GAS FIELD


CHEVRON USA PRODUCTION CO.
A-A' N-S STRUCTURAL CROSS SECTION
WITH PRESSURE GRADIENTS
0 FEET
2000 0 FEET
Pressure Fault Seal

Figure 4. Pore pressure zones in terms of PPG mudweight are overlain on the cross section from Figure 2. This overlay is derived from the pore
pressure logs that are calculated from each well’s resistivity profile. These pressure logs are then correlated along structural and stratigraphic
datums to yield a pore pressure cross section.
53
54 Niemann and Krolow

formation water resistivity. For most areas of the Gulf of these values between wells along interpreted struc-
Mexico, however, away from the flanks of salt domes, tural and stratigraphic datums, a pressure gradient
by using straight-hole well logs on moderate dip struc- cross section, in terms of mudweight, was generated.
tures, negligible errors should result from using either This pressure gradient cross section, shown in Figure
type of data. For the purposes of establishing fluid seals 4, dramatically displays the pressure differential along
across faults, the relative differences in pore pressure the fault surface and its wide variation in magnitude
are required rather than the absolute, accurate determi- with depth. Over most of the fault, the upthrown side
nation of pressure at any one point; thus, the choice of is of higher pressure than is the downthrown side. At
method or data to be used for pressure determination is the intersection of “B” fault and the map horizon,
less important than the consistent application of that however, the differential is slightly reversed, with
method, understanding the assumptions used, to all the lower pore pressure upthrown and higher pressure
log data included in the study. downthrown. This higher pressure on the juxtaposed
The overburden pressure gradient can be obtained downthrown side of the fault may add to the fault
by using a composite graphical function derived by zone seal capacity, accounting for larger gas columns
Eaton (1975) or by integrating a bulk density log from seen in the upthrown sandstones at or near the map
a well in the study area with corrections for water horizon. The pressure gradients between fault blocks
depth where applicable. The OBG was assumed to be a along some sections of the fault are equivalent, which
constant 1.0 psi/ft (22.6 kPa/m) in this study to sim- may be coincidental or could mean that portions of the
plify the calculations. The hydrostatic pressure gradi- fault zone are leaking and that the fault blocks are in
ent was assumed for this study to be 0.468 psi/ft (10.6 pressure communication. Arching of the pressure gra-
kPa/m). For this area of the Gulf of Mexico, 1.5 has dients upward across stratigraphic boundaries at the
been used previously as the X exponent, and that middle level of the cross section on the downthrown
value yields results that are consistent with the limited side of the fault suggests there has been some limited
downhole formation pressure data from the field. Ro is pressure communication up or across the fault to the
obtained by plotting the shale porosity indicator val- downthrown side. The mapping of pressure gradients
ues vs. depth, as shown in Figure 3 for the #4 well in from the shale resistivity gives a very complex and
the cross section. If plotted on a semilogarithmic scale, dynamic picture of seal and possible leakage along the
the trend of the shale porosity indicator data, in this fault, which is not possible from other discrete point
case shale resistivity, is approximately linear in the data such as downhole formation pressures, even if
normal or hydrostatic pressure section. This trend is abundant pressure test data were available.
called the normal compaction trend line (NCTL). Val-
ues along the NCTL, which is equivalent to RN in the CONCLUSION
Eaton equation, extrapolate what the shale resistivity
would measure if the entire section were normally Mapping of pressure differentials across fault
compacting under hydrostatic pressure conditions. blocks can aid in the prediction of fault seal and the
The NCTL can also be considered the hydrostatic iso- determination of the height of potential trapped
pressure gradient line equivalent to 0.468 psi/ft or 9.0 hydrocarbon columns. Formation pressure data for
PPG mudweight (10.6 kPa/m). “Mudweight” is an documenting pressure differentials are often scarce or
industry term for pressure gradient that refers to the unavailable, especially at horizons and faults of inter-
density of drilling fluid necessary to balance down- est. Pressures can be calculated from log-derived shale
hole formation pressure safely during well drilling porosity data, such as commonly available resistiv-
and is equivalent to kPa/m by multiplying by 1.18 (or ity/conductivity or interval transit time logs. Approxi-
psi/ft by multiplying by 0.052). Using the Eaton equa- mate determinations of pressure might be made from
tion and the NCTL, a series of isopressure gradient seismic velocity in untested areas where well log data
lines can be calculated and superimposed over the do not yet exist. By applying this pressure data to cross
shale resistivity data in Figure 3. The pressure gradient sections, pressure gradients can be correlated along
environment can then be logged (shown for a portion geologic boundaries and extrapolated to fault zones
of the well on the right-hand side of Figure 3) based on where differences between fault blocks can be
where the shale resistivity data fall within the isopres- observed and evaluated. The areally continuous
sure gradient overlay. This is a tedious process to per- nature of shale porosity–derived pressure data as con-
form manually, especially when a nonconstant OBG trasted with discrete, single-point, formation pressure
function is desired; therefore, computer calculation of data gives a more complete and detailed, although
pressure gradients from digital well log data is recom- approximate, picture of the fluid movement and pres-
mended. sure history in the shales; this enhances our under-
standing and prediction of hydrocarbon entrapment
PRESSURE GRADIENT CROSS-SECTION along fault zones.
MAPPING
REFERENCES CITED
Pressure gradients vs. depth were calculated for
additional logs in the cross section, and the values Allan, U.S., 1989, Model for hydrocarbon migration
were then plotted on the cross section. By correlating and entrapment within faulted structures: AAPG
Pressure Fault Seal 55

Bulletin, v. 73, p. 803–811. Martinez, R.D., and G.A. King, 1987, Formation pres-
Athy, L.F., 1930, Density, porosity and compaction of sure prediction using seismic data from the Gulf of
sedimentary rocks: AAPG Bulletin, v. 15, p. 1–24. Mexico: OTC Paper 5513, presented at the 19th
Eaton, B.A., 1975, The equation for geopressured pre- Annual Offshore Technology Conference, Houston,
diction from well logs: Society of Petroleum Engi- Texas, April 27–30, p. 259–288.
neers Paper 5544, presented at the 50th Annual Fall McCullough, J.D., Jr., 1990, MacPac Field (Matagorda
SPE-AIME Meeting, Dallas, Texas, September Island Block 487-L), new gas reserves in a mature
28–October 1, 5 p. area: Gulf Coast Association of Geological Societies
Fowler, W.A., 1970, Pressure, hydrocarbon accumula- Transactions, v. 40, p. 585–596.
tion, and salinities—Chocolate Bayou Field, Myers, J.D., 1968, Differential pressures, a trapping
Brazoria County, Texas: Journal of Petroleum Tech- mechanism in Gulf Coast oil and gas fields: Gulf
nology, April, p. 411–423. Coast Association of Geological Societies Transac-
Hamouz, M.A., and S.L. Mueller, 1984, Some new tions, v. 18, p. 56–80.
ideas for well log pore-pressure prediction: Society Terzaghi, K., and R.B. Peck, 1948, Soil mechanics in
of Petroleum Engineers Paper 13204, presented at engineering practice: New York, John Wiley and
the 59th Annual Fall SPE-AIME Meeting, Houston, Sons, Inc., 566 p.
Texas, September 16–19, p. 1–11. Weakley, R.R., 1989, Recalibration techniques for accu-
Hottman, C.E., and R.K. Johnson, 1965, Estimates of rate determination of pore pressure from shale
formation pressures from log-derived shale proper- resistivity: Society of Petroleum Engineers Paper
ties: Journal of Petroleum Technology, June, 19563, presented at the 64th Annual Fall SPE-AIME
p. 717–722. Meeting, San Antonio, Texas, October 8–11,
Hubbert, M.K., and W.W. Rubey, 1959, Role of fluid p. 457–466.
pressure in mechanics of overthrust faulting, part 1: Weakley, R.R., 1990, Plotting sonic logs to determine
Mechanics of fluid-filled porous solids and its formation pore pressures and creating overlays to
application to overthrust faulting: AAPG Bulletin, do so worldwide: International Association of
v. 70, p. 115–165. Drilling Contractors/Society of Petroleum Engi-
MacGregor, J.R., 1965, Quantitative determination of neers Paper 19995, presented at the IADC/SPE
reservoir pressures from conductivity log: AAPG Drilling Conference, Houston, Texas, February
Bulletin, v. 49, p. 1509–1511. 27–March 2, p. 703–712.
Sales, J.K., 1997, Seal strength vs. trap closure—a fun-
damental control on the distribution of oil and gas,
in R.C. Surdam, ed., Seals, traps, and the petroleum
system: AAPG Memoir 67, p. 57–83.

Chapter 5

Seal Strength vs. Trap Closure—


A Fundamental Control on the Distribution
of Oil and Gas
John K. Sales1
Mobil Oil Corporation (retired), U.S.A.

ABSTRACT
Prediction of hydrocarbons ahead of the drill is our highest goal. Every
factor relating to it has been studied extensively, except the role that traps
play. I suggest that this role is important, and in some cases pivotal. Traps
act as “valves,” controlling what they retain, leak, and spill. Closure (vertical
distance between spillpoint and top of the trap) vs. seal strength (height of
the hydrocarbon column the seal can retain before leaking) controls this.
Three classes of trap are essential to cover the possibilities. Two key hydro-
carbon columns are used as thresholds for the classification: the highest pos-
sible gas column and the highest possible total column the seal of the trap
will allow. The spillpoint may be above (class 1), below (class 3), or between
(class 2) these two thresholds. The three classes distribute hydrocarbons dif-
ferently. Class 1 traps spill rather than leak gas, finally spilling oil from the
trap. Class 3 traps spill neither fluid, but accumulate oil until it balances seal
strength; thereafter, gas plus excess oil leaks. Class 2 traps leak gas but spill
oil, and have gas/oil contacts suspended in midtrap. The argument that oil
has a lower interfacial angle and entry pressure applies only until leakage
starts. Then gas’s higher position, smaller molecules, and lower viscosity
make it more mobile. Traps equilibrate based on this principle: full of gas
(class 1), partly full of oil (class 3), or filled with a mixed charge (class 2).
To the extent that the best provinces are exposed to an excess of both flu-
ids, this principle controls world distribution of oil and gas. More practically,
in an economy-favoring oil, higher-class traps leak gas, preserving oil in
themselves and traps updip. The trap class of a discovery has implications
that may modify drilling sequence.
During uplift, class 1 traps may be gas-flushed and thus degrade. Class 3
traps should change minimally. Class 2 traps should vent additional, now
more buoyant gas, and make more space for oil, improving their economics.
A slightly different scenario applies if seal strength decreases during uplift.
In the North Sea, Gullfaks, Snorre, and most of the Ekofisk group of chalk
anticlines are class 3, Oseberg is class 2, and Troll East, is class 1. Turner Valley
in Alberta, which has a two-phase fill, perpetual gas flare, great closure, and
1Currently at RR-1, Box 197, Cabot, Vermont, 05647

57
58 Sales

is surrounded by gas fields with less closure, is class 2. The Mahakam


Delta in Indonesia has seemingly random distribution of oil and gas fields,
but on closer inspection, the class 1 gas fields may have the less-damaged
seals and the class 3 oil fields the more-damaged seals. Despite historic
interpretation as a class 1 trap, the Little Lost Soldier Field in Wyoming
appears to be class 3.

INTRODUCTION well, is most often used to determine pressure at which


a hydraulic seal starts to fail by development of
Definitions microfractures. As a rule of thumb, these new frac-
tures start to open at ~80% of lithostatic pressure,
The following paragraphs define terms and concepts defined as the weight of the overburden. This 80%
as used in this paper. “Closure” is the vertical distance pressure is called “fracture pressure.” On a pres-
from the crest of trap to spillpoint. “Spillpoint” is the sure/depth diagram, it is represented as a fracture gra-
highest point in the periphery of the trap—where the dient at 80% of the lithostatic gradient.
growing hydrocarbon column first has egress from it— Seal failure does not axiomatically lead to trap fail-
and the geometric feature limiting the hydrocarbon col- ure; it may instead lead to a steady-state, in which
umn. “Cryptic” spillpoints are those that cannot be seen there is matched flowage into the base of the hydrocar-
on seismic or reconstructed from the structural/strati- bon column and leakage from the top of the column,
graphic framework of the trap; points of smear-gauge with an economically viable volume of hydrocarbons
failure on trap-bounding faults are characteristically retained. Spilling indicates bed-parallel flow into and
cryptic. “Filled-to-spill” indicates a full trap filled to the out of the base of the hydrocarbon column. Leaking
spillpoint; such a trap may also simply be called “full.” indicates flow through the seal in upper parts of the
“Seal strength” refers either to pressure in excess trap. Stratigraphically higher traps are those sealed by
of hydrostatic or to the height of hydrocarbon col- a shallower horizon than the trap in question and nor-
umn that can be retained before leakage. Such a mally positioned directly above it (both inclined axial
column may be thoroughly described as the “maxi- planes and fault cutoffs can modify this vertical stack-
mum possible, oil-dominated column” or the “ ing). Structurally higher traps are those that updip at
maximum possible, most buoyant gas column.” Alter- shallower levels but are sealed with the same horizon
natively, it may simply be called the “trappable col- that seals the trap in question.
umn,” as context requires. “Reserve seal strength” Unless otherwise stated, all discussion in this paper
refers to strength in excess of that needed to balance assumes traps have been exposed to a minimum of one
the most-buoyant (gas) column that closure allows. trap-volume each of both oil and gas. The terms
Conversely, “reserve-closure” refers to closure in “charge” and “fill” are used interchangeably to indicate
excess of the longest possible (oil-dominated) column type and volume of hydrocarbons in the trap. With one
that seal strength allows. exception, condensates are not discussed; hydrocarbons
A pressure-enhanced seal has excess pressure in are considered to be oil or gas only, oil is considered to
and above the seal driving fluids downward into the be less dense than water, and gas less dense than oil.
reservoir. It is characterized by a pressure overhang on These fluid phases are separated by fluid contacts: the
the well’s pressure curve. A pressure-degraded seal gas/oil contact above and oil/water contact below. A
has excess pressure in the reservoir driving fluids volume-controlled fluid contact is dictated by the
upward through the seal. It is characterized by a flat- amount of fluid that has entered the trap. A spillpoint-
tened pressure step in the pressure curve. Seal failure controlled fluid contact is dictated by the spillpoint. A
indicates that the seal has started leaking hydrocar- pressure-controlled fluid contact, also called an equilib-
bons. Trap failure indicates that the seal is damaged to rium fluid contact, is dictated by pressure constraints
the extent that no hydrocarbon column is possible. A determined by seal strength. The latter type of contact
trap is said to have a membrane seal if failure is remains stationary, even as fluids enter and exit the trap.
through preexisting sedimentologically created pores These relationships show most clearly on a pres-
that become the failure path; it is said to have a sure/depth diagram—a cross-plot showing increasing
hydraulic seal if the failure path is through fractures depth downward and increasing pressure to the right.
newly opened by pressures in the reservoir. On such a diagram, water has a more gentle hydro-
The mercury injection test is most ofte used to deter- static gradient, reflecting its greater pressure buildup
mine capillary entry pressure, the pressure at which a with increasing depth. Gas has the steepest gas gradi-
hydrocarbon-wet failure path is established, in a sample ent, reflecting the least pressure buildup with increas-
from a membrane seal. The leak-off test, done in the ing depth. Because density of oil is intermediate, the
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 59

oil gradient is intermediate. These diagrams may to evolve, the trap is now steady-state, with fluid con-
extend from surface to basement, or crest of trap to tacts nearly stationary and dictated by pressure, rather
oil/water contact, as needed. Aspects of them, such as than volume of oil and gas entering the trap.
pressure in a certain horizon, can be contoured basin- The closure vs. seal strength principle leads to three
wide. Trap classes are often discussed in the order 1, 3, classes of trap, defined on the relation between closure
2 because the classification represents a continuum of and the maximum possible gas column and the longest
seal strength relative to closure; it is clearer to explain possible (oil-dominated) column. After capillary break-
the end-members and then the intermediate case. The through and establishment of a hydrocarbon-wet “fail-
terms “seal” and “sealing horizon” are used in prefer- ure” path through the seal, the longest column is
ence to “caprock.“ The closure vs. seal strength princi- almost surely a nearly pure oil column.
ple refers to interrelation of those two factors, the basis The resulting three-class trap classification can be
for the trap classification presented here. Trap process- visualized most easily in a simplified cross section
ing refers to the manner in which traps retain, leak, and showing one of each class of trap (Figure 1). This
spill fluids, based on this principle. Class 1 reasoning cross section contains the three principal reference
(Gussow’s differential entrapment) refers to spilling of lines used in the classification: (1) top of trap, (2) max-
oil from a trap by gas, and the implications this has for imum possible gas column, and (3) the deeper
fluid distribution in successive updip traps. maximum possible, oil-dominated column. Spillpoints
may be above the maximum possible gas column
Overview (class 1); below the maximum possible, oil-dominated
column (class 3); or between these two reference col-
Most papers on trapping and migration assume umn heights (class 2). The following paragraphs pro-
traps have enough seal strength to prevent leakage. If vide a brief comparison of trap classes (Table 1 gives a
so, because gas is more buoyant and overlies oil, and summary).
because oil is normally generated first and at lower lev- Class 1, gas-prone, traps have seal strength exceeding
els of maturation, gas might force oil to be spilled from closure: reserve-seal strength, even for a gas-full column.
traps in a manner that can be predicted. A trap should Gas can flush oil and, once it does, oil “underflows,”
sequentially be charged with water, oil/water, oil, rather than entering, the trap. Such traps spill both flu-
gas/oil, gas. A migration route would show the transi- ids, leak neither, and preferentially accumulate gas.
tion gas, gas/oil, oil, oil/water, water in successively Class 3, oil-prone, traps are at the other extreme,
updip traps. Once traps in such a route have been with spillpoints deeper than longest possible, oil-
flushed with gas, additional oil would underspill them dominated column that their seal’s strength allows.
and continue updip, eventually to be lost to the surface. Rather than having reserve-seal strength, as class 1
At the extreme, migration chains, basins, or whole traps have, they have reserve-closure. Some combina-
provinces might become gas-flushed, even though oil tion of great closure and/or diminished seal strength
had preceded gas in the traps. This is sometimes forces a class 3 trap to leak from its crest before it fills
referred to as “Gussow’s principle” (Gussow, 1954). to its spillpoint. Because gas overlies oil, oil accumu-
Such analysis assumes, however, that seal strength lates until it uses all seal strength, with none left to
exceeds closure (i.e., that the seal can retain even a gas retain gas. Gas flows through such a trap, with accu-
column, the most buoyant column, filling it). The sce- mulation of a minimal free gas cap. Gas leakage must
nario is very different if closure exceeds seal strength; balance the rate that gas is entering the trap (i.e., be
that is, if the seal leaks before the trap fills. Such a trap steady-state). Oil also leaks steady-state, but does
must preferentially leak gas: gas is more buoyant and accumulate. Only oil in excess of that required to bal-
has a position above oil; it has both its buoyancy and ance seal strength is leaked. Such traps have failed in
that of the total oil column below it forcing it through the classical sense that capillary entry pressure has
the seal; and gas is made up of smaller, more mobile been exceeded and a hydrocarbon-wet failure path has
molecules and has less viscosity. In such traps, because been established. They are, however, still viable traps.
oil can’t spill, it accumulates until the buoyancy of the In fact, from an economic perspective, they are some of
oil column matches seal strength, with no seal strength our best traps, because their hydrocarbon columns are
remaining to support an additional gas column. With dominated by oil. Thus, class 3 traps perform nearly
the exception of gas in solution in the oil, gas must now opposite to that predicted for class 1.
leak as fast as it enters the trap, with minimal free gas Class 2, balanced or two-phase-prone traps, are
in the trap, regardless of the quantity of gas entering transitional between class 1 and class 3 and have
the trap. Oil in excess of that required to use all of the intermediate seal strength relative to closure. They
trap’s seal strength must also leak from the upper part have enough seal strength to support a full oil col-
of the column as fast as it enters the trap. umn, but not enough to support a full gas column.
During filling, trap charge evolves, reflecting vol- The height of their free gas cap is limited by gas leak-
umes of oil and gas entering the trap. After one trap age, but the spillpoint intervenes before all seal
volume each of both oil and gas have entered the trap, it strength for oil is utilized. With the gas gradient con-
comes to an equilibrium forced by its particular combi- strained by the strength of the seal for gas, and the oil
nation of closure relative to seal strength. Once at equi- gradient constrained to the spillpoint, only one posi-
librium, additional oil and gas entering the trap will not tion of the gas/oil contact, somewhere in midtrap, is
significantly change the charge. Rather than continuing viable. Like the oil/water contact in class 3 traps, the
60 Sales

Figure 1. Diagrammatic
G G cross section showing one
O O of each class of trap, and its
CREST OF TRAP
characteristic fill. Bubbles
(G = gas, O = oil) indicate
G ?
G the type of fluid that is
GAS G/O spilled or leaked from each
class of trap.
W
MAX GAS
COLUMN
OIL
O
OIL
WATER
MAX (OIL-DOMINATED)
COLUMN WATER

CLASS 1 CLASS 2 CLASS 3

gas/oil contact in class 2 traps is pressure-controlled, These relationships show most clearly in pressure/
rather than volume-controlled. This is called the depth graphs (Figure 2). Gussow (1954) correctly inter-
“equilibrium gas/oil contact.” Regardless of how preted the mechanics and implications of what are here
much gas leaks from the crest of class 2 traps, and how termed class 1 traps. He, however, considered all traps
much oil spills from their spillpoint, their gas/oil con- to be class 1 and thus, conceptually, only considered
tact will remain suspended in midtrap. one-third of the possible spectrum of solutions.
During filling, fluid contacts are volumetrically Implications for exploration and field development
controlled: they evolve as more oil and gas are are significant. Traps function as “valves,” distributing
added. Thereafter, fluid contacts are either spillpoint oil and gas in a manner that can be predicted if the
or pressure-controlled. Gas/water contacts in class 1 basic data are available. The two key pieces of data are
traps and oil/water contacts in class 2 traps are spill- whether the trap is filled to its spillpoint and whether
point-controlled. Gas/oil contacts in class 2 traps and it is actively leaking. These factors are discussed later.
both fluid contacts in class 3 traps are pressure-con- Thus, if seal strength exceeds closure (class 1), the
trolled. The gas/oil contact in class 3 traps, however, trap is likely to be full of gas, and traps stratigraphi-
may be forced right against the top of the trap because cally above it are likely to be water-wet. Traps updip
oil is using all seal strength. Pressure-controlled con- are likely to be gas-charged, although oil charges may
tacts maintain position as fluids flow through the trap. occur farther up the migration route. If seal strength is

Table 1. Characteristics of the Trap Classes.

Characteristic Class 1 Class 2 Class 3


Dominant Fluid Gas Both Oil
Spills Both Oil Neither
Leaks Neither Gas Both
At Seal Cap No Yes—Gas Yes—Both
Filled-to-Spill Yes Yes No
Controls GOC* Charge Volume Pressure Pressure
Controls OWC** Spillpoint or Charge Spillpoint Pressure
Trap Updip Gas Oil Water-Wet
Trap Stratigraphical Above Water-Wet Gas Both
Uplift Gas Flushes Improves Variable

*GOC = gas/oil contact.


**OWC = oil/water contact.
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 61

PRESSURE
RESERVE SEAL STRENGTH
➟ Figure 2. Pressure-depth diagrams
characterizing the three classes of
SEAL STRENGTH TOP OF TRAP
traps. Note that they have the same

GA S
W
CLASS 1 TRAP three fundamental horizontal lines as
Figure 1: crest of trap, maximum

GR A
AT BOTH OIL AND GAS SPILL
ER
GR (NOTHING LEAKS) possible gas column, and the deeper,

DIEN
AD O
IE
NT maximum possible total (oil-

T
SPILL POINT
G GRADIENT OF MAXIMUM dominated) column.
POSSIBLE GAS COLUMN
DEPTH

MAXIMUM POSSIBLE GAS COLUMN


MAXIMUM POSSIBLE (OIL-DOMINATED) COLUMN

GAS OIL FLUID TYPE

G PRESSURE
RESERVE SEAL STRENGTH (OIL)

SEAL STRENGTH TOP OF TRAP

W
AT EQUILIBRIUM
CLASS 2 TRAP
ER GAS/OIL CONTACT
GR
AD
GRAD

OI

IE
L

NT
GAS T
GR
AD
IEN

IE
NT
DEPTH

MAXIMUM POSSIBLE GAS COLUMN

O

SPILL POINT

MAXIMUM POSSIBLE (OIL-DOMINATED) COLUMN

PRESSURE ➟
GAS AND OIL LEAK
SEAL STRENGTH TOP OF TRAP

W
AT
CLASS 3 TRAP
ER
GAS

GR
AD
OI

IE
L

NT
GRA

GR
AD
DIE

IE
NT
NT
DEPTH

MAXIMUM POSSIBLE GAS COLUMN


MAXIMUM POSSIBLE (OIL-DOMINATED) COLUMN

OIL/WATER CONTACT
NOTHING
RESERVE
SPILLS
CLOSURE

SPILL POINT
62 Sales

exceeded by closure (class 3), the trap may contain maturation boundaries migrate through the rock mass
oil, leak gas and oil to stratigraphically higher traps, of the basin.
and not spill either fluid updip to structurally higher Variations in rate of fluid input are assumed to be
traps. If the trap has a relationship between closure subordinate to the relationship between seal strength
and seal strength that is intermediate (class 2), it is and closure in determining fluid contact levels. Vari-
likely to leak gas to traps that are stratigraphically ations in rate of input to traps at seal strength (class 2
higher and spill oil to traps that are structurally and class 3) will be compensated for by adjustments
higher, while retaining a two-phase fill. Both class 2 in rates of leakage and spillage. Thus, while a tempo-
and class 3 traps prevent gas from continuing updip rary surge might disturb this equilibrium, the shift of
to structurally higher traps, allowing traps updip to fluid contacts should be minimal and short-lived
retain oil, even if these traps updip are class 1 and because of the downward increase in perimeter
capable of being flushed with gas. If several traps in a (“flare”) of most traps’ seals. If, for example, a trap is
basin act this way, the basin may contain significant forced to leak more gas because of a surge of gas into
oil, even when exposed to excess gas. If these higher- the trap, the area of seal exposed to gas leakage can be
class traps are the deeper traps in the basin, located doubled with only a nominal lowering of the gas/oil
near the lower ends of long migration routes, their contact because of this strong downward increase in
oil-enhancing effect may be multiplied, because the perimeter. If the surge lowers the gas/oil contact,
many traps updip may retain oil. The entire line of not only does the surface area exposed to gas leakage
reasoning can be reversed, and stratigraphically and increase, but pressure increases at every point, acceler-
structurally shallow traps can be used to decipher the ating leakage at each point. These two factors com-
probable fill of deeper traps. bined should create pronounced stability in the
position of the gas/oil contact and force reestablish-
ment of equilibrium rapidly; in a geologic time frame,
Assumptions the trap would be steady-state.
The following assumptions are generally adhered An approximate quantification can be done by con-
to in this study: sidering a trap as a cone with low-dipping sides.
Hydrocarbons migrate as separate phases of differ- Indeed, the cone representing the “average” trap
ing density, controlled dominantly by buoyancy (Eng- would probably have dips in the 2˚–5˚ range, although
land et al., 1987; Tissot, 1987; Durand, 1988; Ungerer et faulted boundaries are often steeper. This almost
al., 1990). Seal failure may be by capillary break- degenerates to a circle. Doubling the height of the gas
through (Berg, 1975) or hydraulic fracturing (Watts, column is approximated by doubling the radius of the
1987). Both are covered by the trapping mechanics circle. Thus, the area of the cone exposed to gas is
suggested here. Condensates are also included, but are about squared when the free gas cap height is dou-
limited to class 1 or class 3 traps. bled. It is more difficult to quantify the increased rate
Leaking and spilling cease when charging ceases, of leakage, but it would be significant. As a rule of
with minimal shortening of columns from what they thumb for testing, I suggest that leakage is about
were during maximum charging. The alternative, cubed when the gas cap height is doubled on average
that seals become hydrocarbon-wet, the trap drains, higher-class traps. If we invert this line of reasoning,
and the seal does not repair itself until the column is doubled inflow of gas would only require a small
dissipated and the seal again becomes water-wet, cube-root increase in height of the gas cap to reestab-
seems unlikely; there are too many high and oil- lish equilibrium leakage.
dominated hydrocarbon columns in areas requiring A simple experiment illustrates this: hold a cone-
long-term retention for that alternative to be the gen- shaped coffee filter under a faucet and increase flow
eral rule. Class 2 and class 3 traps have to have had until the water level equilibrates in midfilter. First
hydrocarbon-wet seals that were actively leaking increase and then decrease flow from this equilibrium
when they were forming. They are dynamic rather level; note how large changes in flow rate cause rela-
than static traps, that cause traps in areas of active tively small changes in the equilibrium water level.
charging simply because a column “backs up” The average trap should have even more stability
behind them due to excessive inflow relative to leak- because the cone representing it is relatively lower and
age. Somehow, perhaps because of pore clogging flatter (has a greater flare factor) than the cone of the
(Leith et al., 1993), they become statically competent coffee filter.
at nearly the same column heights as when actively It is assumed that traps either leak or spill a particu-
charging. lar fluid, but seldom do both in subequal proportions.
In class 2 and class 3 traps, it is assumed that fluid If seals were perfect, no pore would leak until all
contacts and rates of spilling and leaking continually started leaking at the same entry pressure, at any one
equilibrate. This seems to be substantiated by rapid height in the trap. To the extent that seals are imper-
rates of migration relative to geological processes fect, there may be a gray zone that allows some leak-
(Dembicki and Anderson, 1989; Schowalter, 1991; age in class 1 traps that still dominantly spill their
Xiaowen et al., 1992). Rates of fluid flow and, there- fluids. This is not to be confused with the situation in
fore, equilibration seem almost always to exceed tec- class 2 traps that leak gas but spill oil because of pres-
tonic rates and, more importantly, rates at which the sure constraints.
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 63

PRESSURE ➟
SEAL AT TOP OF PRESSURE CELL
PRESSURE CELL
OVER PRESSURE THAT HYDROCARBON COLUMN
CURVED

CHARACTERIZES CELL
SEAL FOR TRAP
O
THIS TRAP MUST BE LEAKING
GAS

HYDROCARBONS – GAS IF 2-PHASE


G
DEPTH

BOTH IF OIL-CHARGED
TOP OF OIL COLUMN
WA

LIMITED BY FRACTURE
OIL

GRADIENT – THIS TRAP


TE
GRA

RG

MUST BE CLASS 2 OR CLASS 3


GRAD

DIE

RA
2 PHASES

OIL GRADIENT FOR TRAP


DIE
NT

LI
IENT

NT

FR TH
A OS
C TA
TU
R TI
E C
DEEP PRESSURE G GR
CONDEN-

R AD
SATES

BUILDS TOWARD OIL/WATER A


D IE
FRACTURE IE NT
CONTACT N
T
GRADIENT
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
Figure 3. Idealized pressure/depth diagram showing an overpressured compartment (lighter shaded area) with
a trap (darker shaded area) superimposed on it. The trap is limited by the fracture gradient. Solid lines repre-
sent gradients; the dashed line is a pressure curve, such as a mud weight curve from the well. Because gas is
compressible, its gradient is curved rather than straight. The depth where it becomes tangential with the oil
gradient approximates the top of the condensate realm. Bottom scale is density.

SINGLE VS. DUAL SEAL STRENGTH 2 and class 3), the column will be limited first by how
MODELS much hydrocarbon has reached the trap, or later by
transmissibility of the hydrocarbon-wet failure path
There is room for interpretation as to what seal relative to input to the trap. Only for a theoretical
strength and seal failure mean. Most classic treat- instant, and only in a trap with excess closure relative
ments (e.g., Berg, 1975) emphasize capillary entry to seal strength, will the hydrocarbon column be lim-
pressure, which is the pressure differential at which ited by the initial capillary breakthrough pressure in
hydrocarbons are initially forced through pores of a water-wet seal.
the previously water-wet seal, as the key factor. At the point that leakage is initiated, at least six
Semantics become important. This factor places the things can happen: (1) traps can drain; (2) fluid contacts
theoretical upper limit on column height, but traps can come to equilibrium; (3) column heights might
do not catastrophically “fail” when it is exceeded. reduce to a lower “dynamic” seal strength that con-
Rather, they become “supply and demand” (class 2 trasts with the initial “static” seal strength that existed
and class 3) type traps that “back up” hydrocarbon prior to capillary breakthrough; (4) seals might readjust
columns while equilibrating steady-state to the inter- and reduce pore-throat sizes, as at first larger and then
play between active input to, and leakage from, the successively smaller molecules begin to clog failure
trap. If the trap has reserve-seal strength (class 1), the paths; (5) traps might respond to supply variations,
column will be limited first by how much of each retaining higher columns during active inflow and
fluid has reached it, or later by the spillpoint. If the lower columns as inflow wanes; or (6) traps might
trap has excess closure relative to seal strength (class drain to a lower column height, hydrocarbon filaments
64 Sales

PRESSURE ➟ Figure 4. Pressure-depth


diagram showing the
SEAL STRENGTH FOR OIL SEAL STRENGTH FOR GAS IN
IN DUAL-STRENGTH CASE DUAL-STRENGTH CASE, FOR
difference between the
BOTH IN SINGLE-STRENGTH CASE single (solid line) and dual
(dashed line) seal strength
TOP OF TRAP, ALSO EQUILIBRIUM models. Note that the
GAS/OIL CONTACT FOR dual-strength model leads
SINGLE-STRENGTH CASE
to a significant gas cap
OIL GRADIENT, when at equilibrium; the
MAXIMUM COLUMN, single-strength model
DUAL-STRENGTH CASE
does not.
OIL GRADIENT,
DEPTH

MAXIMUM COLUMN,
MAXIMUM POSSIBLE SINGLE-STRENGTH CASE
(GAS DOMINATED) COLUMN,
W
IF ONLY GAS AVAILABLE TO TRAP AT
ER

GR
AD
IE
NT
OIL/WATER CONTACT (DUAL-STRENGTH CASE)

BOTH ARE "EQUILIBRIUM" CONTACTS

MAXIMUM POSSIBLE (OIL-DOMINATED) COLUMN,


ALSO OIL/WATER CONTACT (SINGLE-STRENGTH CASE)

may “snap off,” repairing failure paths, and traps may with a column made up almost entirely of oil, although
refill to pre-initial-leakage column heights. gas may be contained in solution (Figure 4, solid-line
It is not apparent which, or what combination, or in case). If seals have a higher seal strength for gas than
what sequence, or what quantification of each possi- for oil, traps should start to leak oil steady-state before
bility applies. One thing is apparent: it is not axiomatic the gas cap diminishes to 0. Class 3 traps should equi-
that traps “fail” (drain) when capillary entry pressure librate with one-third to one-fourth gas by column
is exceeded; there are commercial traps with active height (Figure 4, dashed-line case). This paper advo-
leakage above them, with seals resaturated with cates that neither of the first two alternatives are cor-
hydrocarbons from the reservoir, and traps not filled- rect and that, after leakage of both fluids has been
to-spill in areas of overexposure to overample hydro- established, it is gas’s higher position and lower vis-
carbons. There are traps not filled-to-spill whose cosity that allow oil to drive it out of the trap.
upper hydrocarbon column pressure as deduced by The correct interpretation should be represented
drill stem tests is in balance with fracture pressure by a maximum in frequency of occurrence (cluster) in
defined by leak-off tests (Figure 3). statistical data on trap fill (Figure 5). To date, the only
Schowalter (1979) suggests that, under most condi- test seems to be of the Norwegian data (Sales, 1993),
tions, gas has higher entry pressure than oil, due to a as exemplified in the cross-plot of column height vs.
higher interfacial angle; oil is more “wettable.” Watts percent of oil in the column-by-column height (Fig-
(1987) follows this lead and shows that the relation ure 6). Note that this cross-plot also identifies traps
between column height can be variable for the two flu- that appear filled and not filled, and that there is con-
ids, with the trappable oil column sometimes being siderable consistency: gas traps are almost always
less and sometimes more than the trappable gas col- filled, oil traps are not. A data cluster at nearly 100%
umn. Although not a major emphasis of his paper, oil by column height (similar to Figure 4, single-
Watts also suggested the concept of a “gas veneer,” a strength case), but the presence of no such data clus-
thin veneer of gas clogging the critical pore throats ter at 60–80% oil by column height (similar to Figure
and adding its entry pressure characteristics to the oil- 4, dual-strength case), is strong evidence for either,
dominated column that it is holding back. He pointed (1) the gas-veneer and single-seal-strength interpreta-
out that such a column should utilize gas’s higher tion or (2) the relative viscosity argument that
entry pressure and oil’s lower buoyancy stress to increases gas’s mobility to greater than that of oil in
allow the longest of all possible columns. hydrocarbon-wet leakage paths. The data do not
While these uncertainties do not affect the viability indicate that leaking traps have higher entry pressure
of the closure vs. seal strength principle, they do affect for gas than for oil, leading to a significant free gas
the configuration of traps at equilibrium. If the gas- cap in traps at equilibrium. Tentatively, the most gen-
veneer model is appropriate, the gas cap degenerates eral case is the simpler single-seal-strength model
to a thin film and class 3 traps should tend to stabilize (solid-line case in Figures 4, 5).
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 65

% OIL IN COLUMN
0 10 20 30 40 50 60 70 80 90 100
600
DATA CLUSTER IF SINGLE-STRENGTH

DATA CLUSTER IF
500 DUAL-STRENGTH
COLUMN HEIGHT (m) CASE PREVAILS

DOMINANTLY FILLED,
GAS DOMINATES
400

DOMINANTLY NOT FILLED,


300
OIL DOMINATES

200

"SADDLE" REFLECTS
LESS-FREQUENT
100 OCCURRENCE OF
CLASS 2 TRAPS

CLASS 1 CLASS 2 CLASS 3 TRAPS


0
Figure 5. Idealized cross-plot of column height vs. percent oil in the column, by column height, if traps came
to equilibrium with the single-seal strength model (solid curve), or the dual-seal strength model (dashed
curve). With the single-seal strength model, trap frequency should have a peak at ~100% oil by column height,
reflecting a nearly pure oil column as the preferred class 3 column. With the dual-seal strength model, trap fre-
quency should have a peak at 60–80% oil by column height, reflecting that preferred column for class 3 traps.
Both models should have a second peak at ~0% oil, reflecting gas-flushed class 1 traps. Class 2 traps, because
they are transitional and should occur less frequently, should normally occur as a frequency “saddle” between
these peaks in both models. As Figure 6 shows, the single-seal strength model appears to be the more valid.
Compare also with Figure 4.

In summary, seal theory has concentrated on condi- CLOSURE, SPILLPOINT, AND


tions leading to initial capillary breakthrough, but con- FILLED-TO-SPILL
ditions after breakthrough are more important. An
important point has been overlooked: traps that spill Normally, closure and spillpoint are considered to
both fluids (class 1) never test seal strength; traps that influence size and economics of a prospect, but not
leak fluids (class 2 and class 3) immediately go beyond mechanics by which its fill-type is determined. In this
initial breakthrough and equilibrate to a steady-state; paper, it is suggested that closure and spillpoint play a
no trap ever “hangs” at the condition of initial capil- dominant role, in combination with seal strength, in
lary breakthrough. Likewise, if the main thesis of this determining what kind of fluid the trap is capable of
paper is correct, few traps drain catastrophically after holding. This determination pivots on whether a trap is
initial capillary breakthrough. filled, which is critical in determining trap class. Unfor-
tunately, closure and spillpoint may be hard to evaluate
from field and literature data. To quote Zeigler (1992,
WHAT HAPPENS WITH p. 503): “Determining, with any confidence, whether
CONDENSATES? traps were full to their structural spillpoints or were
underfilled was impossible from the oil and gas field
Condensates do fit into this trap classification. In catalogs. Consequently this important aspect cannot be
successively deeper traps, more gas can be dissolved used to help evaluate the potential for leakage through
in oil, until there is no distinction between phases (Fig- the interconnected pore throats of the sealing beds.”
ure 3). This eliminates class 2 traps. There can still be I reject that it is impossible, but admit that it is diffi-
class 1 condensate traps, with excess seal strength that cult. The determination is simpler in structures with
spill condensate updip. There can still be class 3 con- four-way dip closure than in faulted structures, and
densate traps with excess closure that leak condensate simpler in faulted structures than in stratigraphic
rather than spill it. Although a two-phase charge is traps. Different workers use differing analytical and
precluded in the traps in question, condensate spilled reporting procedures. Obvious traps with the largest
or leaked to shallower levels can exsolve back to sepa- columns, “sheepherder anticlines,” were discovered
rate phases. early, with crude techniques and equipment. Data
66 Sales

% Oil in Column fractures and independent of structural closure (Van


den Bark and Thomas, 1981; Watts, 1983). Such traps
0 20 40 60 80 100 appear to be filled beyond their spillpoint, but are
700 31
actually not filled-to-spill.
In faulted traps (Figure 7), the spillpoint may be the
structural spillpoint, a permeable unit juxtaposed
600 33 against the reservoir across the fault, or a point of
smear gauge failure (Skerlec, 1993). If at the structural
spillpoint, the fault must be cross-sealing. If spilling
through a permeable unit juxtaposed across the fault,
500 this point will normally be higher than the structural
Column Height (m)

spillpoint. Such a trap appears not filled to its struc-


tural spillpoint, but actually is filled-to-spill. Likewise,
400 5 a spillpoint at a cryptic point of smear gauge failure
6 can cause a full trap to appear unfilled. It may appear
2 to be filled either deeper or shallower than the highest
24
9 point of juxtaposition of sand on sand on a fault plane
300 27 diagram (Downey, 1984; Allen, 1989). Because the
22,35
oil/water contact in class 3 traps can also be in any
28
25,34
24,12
position above the structural spillpoint, it may be diffi-
cult to judge whether a cryptic spillpoint, a limit on
200 36 11
8 fluids supplied to the trap, or a limit on seal strength is
26,10,7
3,16,21 36
limiting column height. Indications of leakage above
19,37 4 the trap favor the latter.
100 15 Postdrill, the most meaningful clue to trap class
20,17,18,23 30,14,13 may be performance of the trap itself. If the trap has an
32 29,28 oil charge in an area of apparent exposure to gas or
14B with indications of hydrocarbons over it, the oil/water
0 contact is likely to be a pressure-controlled contact in a
Figure 6. Cross-plot of hydrocarbon column heights class 3 trap, not filled-to-spill. If the trap is gas-
vs. percent oil in the column, by column height, for charged, it is most likely class 1 and filled-to-spill,
Norwegian fields. Numbers are keyed to the field although the spillpoint may be cryptic. If gas was
chapter numbers in Spencer et al., 1987. Italicized encountered at higher horizons and it has two-phase
numbers denote traps that appear to be filled-to- fill, it is probably class 2 and filled to a cryptic spill-
spill; upright numbers denote those not appearing point. If no gas was encountered, that same trap is
to be filled-to-spill. The judgment was based (1) on probably class 1, still evolving, but filled to a spill-
the original author’s statement, (2) assessment of the point. If oil-filled, but in a basin of very low matura-
original author’s map, (3) crest minus lowest closing tion, it may be a class 1 trap that gas has not reached.
contour statement in tabulated data, and (4) pressure
data. Note the skewing of the trap population
toward all-gas and all-oil traps, and the strong over-
SEAL STRENGTH
lap of column heights in these two populations. A trap’s seal strength is controlled by the largest
Compare with Figure 5. flaw in the best-sealing membrane nearest the crest of
the structure. That flaw may rarely be sedimentologi-
cally induced, occasionally diagenetically induced,
such as pressure measurements were not taken or and frequently structurally induced. It may predate
reported for these largest and most important traps. hydrocarbon charging and act passively as the largest
Waste strata (transitional strata between reservoir pore that becomes the failure path. Alternatively, it
and seal) won’t produce in a field-life time frame, but may be hydraulically opened by pressures created by
may leak in a geological time frame. Thus, the actual the buoyancy of the hydrocarbon column added to the
effective spillpoint may be some distance up into this reservoir pressure in the water system (Hubbert and
waste strata, and not be exactly at the top of the pro- Willis, 1957; Watts, 1987).
ducing reservoir (Schowalter, 1979). Such a trap may The best and most used test for estimating seal
be filled, while appearing not to be. strength is the mercury injection test (Berg, 1975;
Fractured traps, such as salt-induced chalk anti- Schowalter, 1979). To the extent that seals are not uni-
clines of the North Sea, give special problems. Bending form, the success of this test may depend on compen-
produces fracturing, the essential permeability of the sating sampling errors. On the one hand, the sample(s)
chalk reservoir. Bending, fracturing, and permeability probably seldom include the best and actual sealing
diminish off-structure. Actual closure in the labyrinth membrane in the thick shale unit. The sample actually
of fractures usually exceeds structural closure of the tests somewhat inferior shale that gives a somewhat
fold. The actual spillpoint is hidden in this labyrinth of lower entry pressure than the best membrane. On the
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 67

Figure 7. Options for


spillpoints in a faulted trap.

other hand, the sample almost surely does not test the were randomly matched, half of the traps would have
exact largest pore in the exact critical failure path in a closure that exceeded seal strength. Because the
that best membrane in the crest of the trap. This critical upper limit on closure exceeds the upper limit on seal
pore would also have an entry pressure somewhat less strength, more than half of all closures should exceed
than the average for the best membrane. In sedimento- their seal strength. (3) Many traps appear to leak (Fig-
logically uniform condensed shales that have not been ure 8). (4) Many of these, while not filled, retain a com-
made brittle by uplift or excessively strained by struc- mercial hydrocarbon column. (5) Statistical studies
turing, a high degree of accuracy can be obtained from (Figure 6) suggest that the strength of seals may vary
mercury injection tests, especially if multiple samples significantly from trap to trap, further increasing the
are tested (Sneider and Neasham, 1993). probability of weak seals paired with large closure. In
While it is statistically improbable to sample this fact, larger closures have greater curvature, larger
single actual failure point in a core, it is sampled auto- faults, more secondary brecciation, and more severe
matically in traps in areas of active generation that are tectonic stressing, which aggravates this tendency. (6)
at seal strength and leaking. Thus, the hydrocarbon Pressure studies show many traps are at seal strength;
column in active class 2 and class 3 traps exactly por- pressure at the top of the hydrocarbon column, as indi-
trays the effective seal strength of that trap. If column cated on RFT tests, coincides with pressure at which
heights and fluid densities are known, seal strength the seals fail, as indicated by leak-off tests.
can be calculated precisely on a pressure-depth plot.
However, that seal strength is not the initial capillary
breakthrough pressure in a water-wet pore. Rather, it EXAMPLES OF TRAP CLASSES
is the steady-state seal strength of the hydrocarbon-
wet leak path after initial breakthrough, and after Class 1 Trap—Troll East, Norwegian North Sea
inflow and outflow have equilibrated. Troll (Gray, 1987) has three blocks (Figure 9). Each
more westerly and downdip block has a slightly deeper
WHY CLOSURE SHOULD OFTEN hydrocarbon/water contact and a thicker oil column.
EXCEED SEAL STRENGTH Troll East, with 230 m of gas and 0–4 m of oil, is the
largest gas trap in Norway. The 0–4-m oil column may
A significant number of traps should have closure indicate a borderline class 2 trap. Alternatively, Troll
that exceeds seal strength: (1) Closure of fault blocks, East may be a class 1 trap and simply have that much oil
anticlines, and wedge-outs of inclined strata that form in transit, underspilling the giant trap. If so, it has possi-
the world’s traps can be several times greater than bilities for quantifying present migration along one of
hydrocarbon columns (see Zeigler, 1992, for a compila- the most prolific routes in Norway. An isopach of the
tion of column heights for the western United States; oil leg might show thinning toward the spillpoint. This
see Figure 6 for hydrocarbon column heights in Nor- is because a slight slope should be required, both at the
way). (2) If hydrocarbon columns and closures had the base and top of the oil, to allow the oil to flow across the
same (probably log-normal) height distribution, but broad undersurface of the gas.
68 Sales

Figure 8. Seismic section across Ekofisk Field in the North Sea, showing the strong gas chimney over that oil-
filled trap, which is positioned over a gas kitchen [from Dangerfield et al. (1992, their figure 1)].

The western two blocks of Troll cannot be class 1 By class 1 reasoning, each successively more easterly
traps or they would be flushed with gas; gas generated block should have proportionally more oil, not pro-
downdip in the Troll kitchen almost surely passed portionally more gas.
through them on its way to Troll East. These western
blocks have columns appropriate for higher-class Class 2 Trap—Oseberg, Northern North Sea
traps. Rather than leaking vertically to higher strati-
graphic levels, they probably leak across the fault Oseberg (Figure 10) has a 375m gas and 602m total
plane into adjacent blocks (Figure 9). Thus, although column, filled-to-spill (Nipen, 1987). As shown in
Troll East is likely to be the largest class 1 trap in Nor- Spencer and Larsen (1990), and shown clearly in the
way, the three blocks are one of the strongest indica- distribution of column heights, it is an exceptional trap
tions that simple class 1 type reasoning is inadequate. for the area (Field 33 on Figure 6). Apparent gas
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 69

Figure 9. Diagrammatic
M cross section of fill relation-
SS ships in the three blocks of
Approx. 40 Km
1300 the Troll Field, Norway.
Class 3 Class 2 Each successively updip
and more easterly block
contains proportionally less
oil and more gas. Class 1
Class 1 reasoning suggests more oil
1400
updip; therefore, the
Troll Troll western blocks of Troll
Troll must be class 2 traps. Troll
West West
East East is Norway’s largest
(Oil) (Gas)
class 1 trap. Note extreme
vertical exaggeration.
[Based on data from Gray
1500 (1987).]

0-4 M
12 M
±25 M
1600

blooms, hydrocarbon shows in Cretaceous limestone large bounding faults may divert gas from it. How-
stringers, and areas of poor seismic quality (Nipen, ever, the trap may have exposure to gas, is not filled,
1987) suggest leakage. If Oseberg were class 1 and and shows reservoir pressure at the fracture gradi-
capable of filling with gas, it would require about 25% ent (Karlsson, 1986). Pockmarks and acoustic distur-
more seal strength than the actual column requires, bances over the field seen on sparker profiles
and could support an equivalent oil column of >1000 (Hoveland and Judd, 1988) suggest that the seal is
m. Because Oseberg already has an exceptional seal for leaking. There were gas-related drilling problems,
the area (Figure 6), it seems more likely to be a class 2 and relief wells were drilled to vent gas. Indications
trap at equilibrium than a class 1 trap still filling. of gas occur in the shallower strata (Miles, 1990, his
figure 13). Gullfaks has a small gas cap. Some of the
Class 3 Traps—Gullfaks, Snorre, and Ekofisk, gas encountered above the reservoir may be bio-
North Sea genic. Thus, the main criteria that define a class 3 oil-
Gullfaks, which has very large structural closure, prone trap are satisfied: (1) crestal leakage, (2) not
may be a class 3 trap. Eriksen et al. (1987) feel that filled, (3) a column dominated by oil, (4) indications

Pressure (PSI x 100) Figure 10. Pressure-depth


0 1 2 3 4 5 6 7 8 diagram showing
0 generalized seal-spill
relationships for the
Act

Oseberg Field. [Based


Ma
ual

1
on data from Nipen (1987).]
xim Class

EGOC
Depth (M x 1000)

Gas

um

(Class 3) Wa
if

2
Col

terG
Gas 1

rad
um

i ent
Col

3
n

um
n

4 Gas-Oil Contact Ac
tua
lO
il C
5 Capacity Gas Column ol.
Spill Point
6
70 Sales

of leakage, and (5) pressure tests that indicate that gas’ [my emphasis]. However, the remark that ‘...some
seal strength has been reached. of the conditions of oil trapped in the region defy satis-
Snorre is a large complex fault block that, in the factory explanation,’ indicates that he is not all
present interpretation, is a class 3 trap not filled-to- together satisfied with the explanation. According to
spill. It has no significant gas cap. Work by Leith the principle of differential entrapment, no anomalous
(1993) has established that the Kimmeridge Clay seal, condition need exist.”
rather than being saturated with indigenous Kim- Considering geometry and the statements of the
meridge hydrocarbons, is resaturated with hydrocar- original workers, the case seems strong against differ-
bons from the oil pool below that have been forced ential entrapment and for closure vs. seal strength rela-
upward through it. This resaturation is differential, tionships being the primary principle that causes a gas
with smaller and more mobile molecules penetrating charge in Wertz but an oil charge in Little Lost Soldier.
farthest up through the seal. Under this interpretation, both Wertz and Lost Soldier
Ekofisk (Figure 8) and many related chalk anticlines were exposed to both oil and gas. Wertz, with its lower
have characteristics that are similar. They have oil amplitude and less-stressed seal, filled with gas as a
charges, but are located over a gas kitchen (Figure 11) class 1 trap and spilled all of its oil and much gas
and under gas chimneys. updip into the Mahogany-Ferris trend by differential
entrapment. The greater closure and higher degree of
Little Lost Soldier and Wertz Fields, Wyoming stressing of the seal by both faulting and bending, pos-
sibly in combination with deeper breaching, caused
With pays from the Upper Cretaceous Dakota For-
Little Lost Soldier to perform as a class 3 trap, with clo-
mation down to and including fractured basement, the
sure exceeding seal strength. It vented its gas, as the
Little Lost Soldier (Lost Soldier) anticline (Figure 12) is
original workers suggested, allowing it to retain its oil
considered perhaps the best oil field in the Rocky
instead of spilling oil as Wertz did. Thus, the Little
Mountains for its size (although fields such as
Lost Soldier anticline became perhaps the best oil field
Rangely, Salt Creek, and Elk Basin are larger). Wertz
for its size in the Rockies because its closure exceeded
dome, only 2.5 mi (4 km) away, is filled instead with
its seal strength, not because only oil reached the trap
gas. Gussow (1954) interpreted Wertz to have been
by differential entrapment.
filled from the Washakie Basin to the west and to have
spilled oil and then much gas updip to the east into the
Mahogany-Ferris trend. It then spilled only oil into ALBERTA FOOTHILLS BELT
Lost Soldier dome by “differential entrapment,” Gus-
sow’s term for class 1 mechanics. The Alberta Foothills belt (Fox, 1959) contains many
Several things are suspect in this interpretation: (1) gas fields and one larger two-phase field, Turner Val-
Wertz does not, in Irwin’s map (1929), appear to be ley (Figure 13). These fields should have equal expo-
filled. Taken at face value, Wertz could have spilled sure to gas. Link (1949) suggested that the presence of
neither oil nor gas to the east into the Mahogany-Ferris oil at Turner Valley might be explained by trapping of
trend nor to the south into Little Lost Soldier anticline. oil in a precursor stratigraphic wedge-out that later
(2) If oil did spill from Wertz to fill Lost Soldier, Wertz was remobilized into the Turner Valley structure. A
had to be filled-to-spill with gas, and it is extremely second option is that Turner Valley received insuffi-
fortuitous that significant gas did not follow it: Lost cient gas to spill all of the oil from the trap by differen-
Soldier had no significant gas cap. (3) Lost Soldier tial entrapment, by Gussow’s reasoning. Neither, I
actually has as good or better access to the Washaki think, is correct.
depocenter than Wertz does: if overample gas entered With the on-trend and surrounding fields filled
Wertz, overample gas should have entered Lost Sol- with gas, it seems likely that Turner Valley has been
dier. (4) It is as unlikely for Wertz to have spilled both exposed to overample gas to fill that trap. This gas
updip into the Mahogany-Ferris trend and into Lost should have displaced the oil from both the strati-
Soldier as it is for a lake to have two outlets. Ironically, graphic and later structural trap at Turner Valley sug-
the correct interpretation (in my judgment) was well gested by Link (1949). The presence of one of the
stated in Gussow’s quotes from the original mappers, world’s largest deep gas traps (Masters, 1984) under
here reproduced from Gussow (1954) as reprinted in and in front of these structures further suggests over-
Foster and Beaumont (1988). ample gas availability: the entire core of the foredeep
“Twenty-five years ago Irwin (1929) attributed the is saturated with gas. Further, Turner Valley has a
apparent ‘anomalous’ trap of hydrocarbons to the very large and well-known natural perpetual gas flare
presence or absence of faults and their tightness. Thus over it.
the absence of a gas cap at Little Lost Soldier was All of these factors point to venting of excess gas
attributed to selective leakage of gas [my emphasis], and away from differential entrapment as the cause of
whereas the occurrence of gas to the exclusion of oil at the oil leg at Turner Valley. The most likely alterna-
Wertz was attributed to the absence of faults or to their tive may be that Turner Valley, perhaps because it
tightness if present,” Gussow (1954, p. 131). [And is one of the largest of the structures, is a trap with
again:] “...Dobbin (1947, p. 811) stated that the lack of small enough seal strength relative to closure that
gas in the Little Lost Soldier field ‘suggests strongly it is forced to vent its gas at the well-known perpet-
that there has been vertical migration of oil and leakage of ual flare (Porter, 1992), rather than retaining and
Figure 11. Position of Ekofisk relative to gas generation [from Cayley (1987)]. Even though Cayley came to a
different conclusion, this figure, in combination with the strong gas chimneys over traps like Ekofisk (see
Figure 8), is strong evidence that gas has access to these traps, despite the fact that they are filled with oil.
72 Sales

R90W R89W R88W R87W

2000

WERTZ

0
40 000
00
2

200
T
0

0
26
N LITTLE 200
0
0
LOST MAHONEY 200
0
SOLDIER
-20
00

T
25 4000
N

Figure 12. Map of the Little Lost Soldier (Lost Soldier) and Wertz fields, central Wyoming. [Modified slight-
ly from Gussow (1954) as taken from Irwin (1929) (the less-applicable eastern third of the diagram has been
omitted)].

accumulating the gas to drive out the oil leg (i.e., it is a strength relative to closure. Gas becomes less dense as
class 2 trap). the confining hydrostatic pressure decreases during
uplift. In contrast, oil density remains more constant,
although it does change some as gas is exsolved.
MAHAKAM DELTA, INDONESIA There is a convergence of the hydrostatic and fracture
The Mahakam Delta in Kalimantan (Borneo) in gradients toward the surface (Figure 15), the feature
Indonesia (Figure 14) is a small but prolific delta that that limits column heights in shallow traps and allows
contains both oil and gas fields. In a map view, they are some very deep traps to have great column heights.
interspersed in what at first seems to be a confusingly Except for a few traps in rare underpressured areas,
random pattern. There is indication, however, that the all traps are confined between these two gradients,
oil fields are the more highly faulted, while the gas fields and pressure at the trap crest can never exceed frac-
are the more perfect anticlines, with less-faulted seals, as ture pressure. If this happens and the trap is class 1
indicated by the following quote (Duval et al., 1992, p. and gas-charged, it will probably become class 2 or
70): “This does not answer why Bekapai and Attaka oil class 3 and have a two-phase or an oil charge, but will
fields exist along the same trend as Tunu (gas and con- in any case have column height degraded. Charge
densate field). The explanation is that vertical migration, economics may, however, improve if a significant oil
in spite of being a subsidiary process, is an important leg is added.
factor and occurred late in these faulted structures. To thoroughly cover the possibilities, both the case
Whereas Tunu is not faulted, Bekapai is explained by for constant and the case for lowered seal strength
dismigration through active faults from a previous gas during uplift must be considered. Seal strength may
condensate pool in deeper horizons.” remain constant if the traps, after uplift, do not
Thus, some of the oil fields in the Mahakam Delta impinge on the fracture gradient. They will almost
may have enough structural damage to lower their surely degrade if they do. In the following discussion,
seal strength relative to their closure so they perform it is assumed that the uplift does not involve tilting
as class 3 traps, venting gas so that they can retain oil. and that closure does not change. First, assume a con-
The gas fields may have enough seal strength that they stant seal strength during uplift.
perform as class 1 traps, spilling oil updip. Class 1 traps, with reserve-seal strength for gas, can
be gas-flushed. Class 3 traps, with little or no seal
strength for gas, before or after uplift, shouldn’t
EFFECT OF UPLIFT ON THE change much. Class 2 traps should gain room for oil at
TRAP CLASSES the expense of gas: it takes a shorter column of more-
buoyant gas after uplift to match the buoyancy stress
Each trap class should be affected differently by exerted by the longer column of less-buoyant gas that
uplift because each has differing characteristics of seal was in equilibrium with the seal strength of the trap
Figure 13. Regional map showing the southern part of the Alberta Foothills belt gas fields, western Canada [modified from Fermor and Moffat (1992)].
Blowup of the Turner Valley Field is from Link (1949, his figure 3). According to Link (1949): “The structure relief or effective closure, from the highest
known contour on the gas cap to the water-oil boundary, is slightly more than 5000 feet.” Note that while closure is definitive on the east, west, and
south, splay faults at the north end of the structure probably contain a spillpoint. These factors, in combination with its perpetual gas flare due to crest-
al leakage, suggest that Turner Valley’s prime reason for not being flushed with gas, in a province that is dominated by gas, is that it is a class 2 trap
that vents gas to retain room for oil.
74 Sales

Figure 14. Map of the


Mahakam delta, redrafted
from Duval et al. (1992, their
figure 2). Note the close and,
at first glance, random
association of oil and
gas fields.

before uplift (Figure 16). However, the spillpoint inter- gas caps. The oil leg of such class 2 traps would at first
venes before the oil leg gets high enough to use all of expand because class 2 traps have excess seal strength
the oil’s seal strength. Therefore, class 2 traps spill for oil. If uplift is arrested at this stage, these traps
their oil rather than leak it. Thus, there is both room would have improved. If uplift and lowering of seal
and enough seal strength for the oil column to expand strength continue, however, these class 2 traps at some
to replace the shortened gas column. In an economy- stage should convert to class 3 traps and cease to be
favoring oil, this should improve the economics of filled. They now have lost their ability to support a
these class 2 traps. free gas cap and to spill oil. They should now start to
Seal strength may diminish as, during uplift, the leak oil steady-state. These former class 2 traps would
weight of the water column above the seal is lowered, now start to decline economically, as their oil column
while pressure in the reservoir, being more confined, height and volume start to reduce.
may respond more gradually. This may have differing Note that wholesale gas-flushing and catastrophic
influences on the trap classes. Class 1 traps might lose loss of oil to the surface during uplift require all class 1
enough seal strength to convert to class 2, improving traps and no diminution in sealing strength. It was
economically by gaining an oil leg at the expense of appropriate for Gussow (1954) to suggest this possibil-
shrinking gas caps. Class 3 traps should degrade, as ity, because he only dealt with class 1 traps. Because it
diminished seal strength allows less oil column height. is the most easily visualized, industry has picked up
Class 2 traps might be forced toward, and eventually on this and often automatically assumes that uplift
into, class 3 configurations, shrinking their equilibrium equates with gas-flushing. It should not be considered
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 75

PRESSURE ➟ Figure 15. Pressure-depth


diagram showing conver-
gence of the hydrostatic and
SAME TRAP EXCEEDS FRACTURE fracture gradients toward the
GRADIENT AFTER UPLIFT surface. As shown, this may
AMOUNT OF UPLIFT

cause traps that did not


THEREFORE HYDROCARBON COLUMN intersect the fracture gradi-
HEIGHT WILL BE LOWERED AT LEAST ent before uplift to exceed it
THIS MUCH, AND TRAP WILL BECOME after uplift, forcing them to
CLASS 2 OR 3
DEPTH

have lower column heights.


TRAP TOP IS WELL BELOW
FRACTURE GRADIENT
BEFORE UPLIFT

UPLIFT MAY FORCE


TRAPS TO EXCEED
OIL GRADIENT
FRACTURE GRADIENT, IN TRAP
LOWERING
COLUMN HEIGHTS

axiomatic, however, because all of the other alterna- updip from the higher-class trap in question may
tives are equally possible. Further, the presence of have some secondary gas-flushing.
just one class 2 or class 3 trap near the downstream
end of a long migration chain of class 1 traps can
vent gas from the chain and possibly save all traps PREDICTION
updip for oil. This is not axiomatic, however, Implications for Migration Chains
because additional gas may be coming out of solu-
tion in each trap affected by the uplift. Thus, The systematic way in which the three classes of trap
depending on the volume of this exolved gas, traps distribute oil and gas has implications for migration

PRESSURE ➟
GAS LEAKS
Figure 16. Pressure-depth
diagram showing the effect
G of uplift on a class 2 trap.
SEAL STRENGTH TOP OF TRAP Class 2 traps gain room for
GAS GRADIENT oil as the gas cap is forced to
(POST-UPLIFT) CLASS 2 TRAP shrink, this is because a
more-buoyant shorter gas
DURING UPLIFT column after uplift must
balance the less-buoyant
W trappable gas column that
GAS/OIL AT
ER characterized the trap before
GAS GRADIENT
DEPTH

CONTACT GR
uplift, raising the gas/oil
AD (PRE-UPLIFT)
(POST-UPLIFT) IE
NT contact. Thus, class 2 traps
may be improved by uplift.

GAS/OIL O
IL OIL
CONTACT G
RA SPILLS
(PRE-UPLIFT) DI
EN
T

OIL LEG OIL LEG O


AFTER BEFORE
UPLIFT UPLIFT
SPILL POINT
76 Sales

Figure 17. Three diagram-


matic cross sections of a
CLASS 1 linear migration chain, in
TRAP which it is assumed that
the chain has been exposed
in each case to enough gas
to flush the entire chain.
The lowest trap in the
CLASS 2 upper cross section is class
TRAP 1; in the middle cross
section, class 2; and in the
lower cross section, class 3.
Note that middle and bot-
tom cross sections have an
CLASS 3
TRAP enhanced chance for oil
even when exposed to an
excess of gas.

(Figure 17). The upper cross section of Figure 17 shows Example 1


a migration chain consisting of all class 1 traps A drilling well just encountered a trap, filled with
exposed to an excess of both oil and gas, but with all gas, but not to its apparent spillpoint (Figure 18). Such a
oil flushed to the sea floor by the gas. trap could have been sourced by spillage from a class 1
The middle cross section shows the same chain trap downdip at the same stratigraphic level. Such a
under the same conditions, but with the lowest trap downdip trap, however, should normally have spilled
being class 2 rather than class 1. Because class 2 traps oil ahead of the gas, and space to have preserved some
vent gas and spill only oil updip, the chain above is of that oil remains in the trap that has just been drilled.
saved for oil. Alternatively, the just-drilled trap was filled from a
The bottom cross section shows the same chain subjacent class 2 trap: the ideal class 2 trap leaks only
under the same conditions, but with a class 3 trap as gas to higher stratigraphic levels and spills all of its oil
the lowest in the chain. Because such traps leak both updip in the main carrier bed. Such a subjacent trap
oil and gas and spill neither, the chain updip should be should contain oil as well as gas because, in principle,
water-wet. Alternatively, tributary migration routes class 2 traps leak only gas and have an oil leg.
might bring oil generated at shallower levels into the Under these constraints, the logical next target is to
upper part of the chain. Thus, in the presence of class 2 deepen the well that is drilling to try for that oil leg. If
and class 3 traps, the migration chain updip has an found, it implies that additional oil has spilled updip
increased probability for oil, even though the lower at this deeper level: test for this after the first well is
end of the chain has had access to excessive gas to completed.
flush oil from the entire chain.
There are also implications for traps stratigraphi- Example 2
cally above the one in question. Such a stratigraphi- The well just drilled encountered gas, filled-to-spill
cally higher trap has increased probability of being (Figure 19). This enhances the possibility that the
water-wet in the upper cross section and full of gas trap was filled by gas spilled from a trap downdip,
in the lower two cross sections. There is a difference in although filling from a subjacent class 2 or class 3
the lower two cross sections, however. The shallow trap at a deeper stratigraphic level is not precluded.
trap in the upper of the two should have received only Now, however, oil spilled from a trap downdip
gas, and still-shallower traps updip from it have could have preceded the gas into the just-drilled
enhanced probability for gas. In the lowest cross sec- trap, but has been displaced updip by the gas. This
tion, the shallow trap should have received both oil possibility is enhanced if there is residual oil staining
and gas, and the fact that it now contains only gas sug- within the gas column. This scenario increases the
gests that oil has been spilled updip. Thus, other shal- possibility that the trap updip has received oil,
low traps updip from it have an enhanced probability although it, like the trap just drilled, may also have
for oil. In the uppermost cross section, these shallow been flushed by gas. If so, you might have to “chase”
traps are most likely water-wet. the oil to still higher traps.
Because the alternative, that the just-drilled trap
Implications for Exploration Programs
was filled from a trap downdip, is now logical, it
These concepts also dictate different drilling somewhat decreases the probability that the just-
sequences, depending on the type of trap encountered. drilled trap was charged from a subjacent class 2 or
The following examples assume access to both fluids, class 3 trap. However, the possibility that the subjacent
that oil is the more economic, and that migration routes trap is a class 3 trap, charged with oil, is also enhanced.
connecting the discussed traps are simple and linear. Class 3 traps leak both fluids, but there is no room for
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 77

Figure 18. Exploration


scenario when a trap that
appears to be only partly
CLASS 1 FILLING? filled to its spillpoint with
WATER
gas is discovered.

JUST DISCOVERED
A TRAP PARTLY FILLED
WITH GAS WATER

GAS OIL
NOT FILLED-TO-SPILL

GAS OIL
G
G

GAS
OIL O

GAS MORE LIKELY


2 OPTIONS
OPTION

the oil in the gas-filled upper trap. If this is the correct Example 4
alternative, it also enhances the chance for oil updip The trap just drilled contains dominantly oil, but with
at the level of the first-drilled shallower trap. The oil gas-related problems encountered while drilling, and it
that leaked vertically from the subjacent class 3 trap appears not to be filled-to-spill (Figure 21). It may either
had no trap space in the gas-filled trap above and be a still-evolving class 1 trap or a class 3 trap that has
must have underflowed it and migrated updip. the capability of supporting some free gas cap. A third
Thus, this scenario has two attractive remaining tar- possibility is that subsidence since filling has com-
gets: (1) deepen the present hole while in it and (2) pressed the gas cap and put more of the gas in solution
drill the trap updip for the possibility of shallow oil. in the oil. Such a trap could have been charged from a
Example 3 subjacent class 3 oil-prone trap that has leaked both
fluids into it. Alternatively, it may have been charged
A trap just drilled contains both oil and gas filled-to- from a class 1 trap downdip that spilled oil and then
spill (Figure 20). It may be a class 1 trap still in the gas into it. The trap downdip is most likely filled with
process of having its oil displaced by gas that is still gas, the excess of which spilled into the just-drilled
accumulating. Alternatively, it may be a class 2 trap trap. The trap updip should be water-wet, since the
at equilibrium and leaking gas. Indications of shal- just-drilled trap is not filled. With the trap downdip
low gas while drilling enhance the class 2 trap inter- likely to be gas-filled, and the trap updip likely to be
pretation. In either interpretation, the trap just water-wet, the best procedure would be to deepen the
drilled could have been sourced from oil and gas present hole and test for the deeper class 3 trap.
spilled from a trap downdip or from oil and gas
leaked from a deeper subjacent class 3 trap. In either
case, the trap updip has an enhanced probability for AN INITIAL ATTEMPT AT STATISTICAL
containing oil because gas has not yet been spilled VERIFICATION: NORWAY
updip from the trap just drilled.
Given these constraints, the best chance for find- These concepts, formulated while working in Nor-
ing additional oil is to drill the trap updip. However, way, were motivated by a feeling that there was more
since you are already in the hole, the best sequence is gas available than recognized, and more oil pre-
probably to deepen the present hole first to see if the served than should have been by class 1 reasoning.
just-discovered trap was charged from a subjacent Many of Norway’s traps leak, and failure paths must
class 3 trap at a deeper stratigraphic level that has be hydrocarbon-wet. Ekofisk, for example, is above a
leaked both oil and gas up into the trap that you gas kitchen, below a gas chimney, and filled domi-
have just discovered. After that, one would drill the nantly with oil (Van den Bark and Thomas, 1981).
trap updip to tap the oil that should have spilled Pressure regimes preclude filling of some traps; full,
updip from the trap that was just drilled. Note that they would exceed fracture pressure as indicated by
the drilling sequence is the same as in the case above, leak-off tests. Gullfaks has these characteristics (Erik-
even though the reasoning differs. sen et al., 1987). Caillet (1993), Lieth (1993) and Lieth
78 Sales

Figure 19. Exploration


scenario when a gas trap
appearing to be filled-to-
CLASS 1 TRAP spill is discovered.
OIL

JUST DISCOVERED
A TRAP FILLED-TO-SPILL GAS
WITH GAS
OIL O
OIL
O
GAS
FILLED-TO-SPILL
G
C-1 O
GAS G OIL
G O
O
OIL
G O

OIL 2 OPTIONS

et al. (1993), suggest that Snorre and possibly Statfjord Each except Oseberg is problematic, but all
may have these characteristics. Thus, some traps in Nor- seem to have significant pressure-enhancement
way have “failed” by capillary breakthrough, yet are of the seal.
commercial and, indeed, the best and largest oil fields. 6. Gas traps have less signs of leakage. Several
To assess whether closure vs. seal strength relation- traps, especially the Ekofisk group, located close
ships played a part in distributing oil and gas in Nor- to gas generation, support spectacular gas chim-
way, cross-plots based on data in the Oil and Gas neys, yet are filled dominantly with oil. Signifi-
Fields Volume (Spencer et al., 1987) were made. A sim- cantly, a few of these are gas-charged, but appear
ple technique was used to safeguard objectivity: work- to have no more access to gas. The oil traps are
ing backward through the book, making decisions as gas-saturated, and produce significant gas, but
to whether the trap was filled without knowing which have little free gas cap.
field it was until the decisions were made. The deci-
sion was based on: (1) author’s statement, (2) my inter- If all of these traps were class 1, with fill determined
pretation of author’s map, (3) fluid contact levels vs. by volume of each fluid delivered, and fluids limited,
lowest closing contour figures from tabulated data fill would correlate with distance from kitchens and
(there is one contour interval of looseness in this trap depth: it does not. Likewise, with excess fluids, all
method), and (4) pressure considerations: would a traps would be gas-flushed: they are not. With just the
filled trap exceed the fracture gradient? right amount of oil and gas, the right stage of migra-
Norway, compared with other prolific areas, has tion could fortuitously produce the skewed distribu-
relatively simple plumbing (Spencer and Larsen, tion seen (about half gas traps and half oil traps, with a
1990), with one dominant source bed (Upper Jurassic few two-phase traps). Then, however, gas should be
Kimmeridge equivalent) and one dominant reservoir- concentrated downdip from oil, and it appears not to
carrier bed (Middle Jurassic Brent equivalent). In that be, either when the total trap distribution is consid-
sense it is an ideal place for such a study. Preliminary ered or when only the Middle and Late Jurassic field
analyses suggest: distribution is considered.
Significant scatter and overlap of column heights
1. Little correlation between fill character and dis- between oil- and gas-dominated traps might be
tance from generating kitchens (Figure 22) or explained in two ways. The gas side of the distribu-
depth of the trap (Figure 23). tion, with traps that tend to be filled, might be
2. A subequal number of gas- and oil-dominated explained by distribution of closure heights. Equal
traps, with few traps with subequal fills of both scatter occurs on the oil-dominated side of the distri-
fluids. bution, where it is less likely to be related to closure,
3. Gas traps tend to be filled; oil traps are not. because this set of traps tends not to be filled. Tenta-
4. Significant overlap and scatter in column heights tively, imperfections in seals, in combination with
between oil- and gas-dominated fields (Figure 6). pressure limitations on column height, explain this
5. A cutoff of column heights at ~400 m. Only a few variation in seal strength and resulting column height.
columns (Oseberg, Agat, Smorbukk) are higher. Had the principles worked perfectly and had there
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 79

Figure 20. Exploration


scenario when a trap with a
two-phase fill that appears
GAS PROBLEMS CLASS 2 TRAP OIL to be filled-to-spill is
WHILE DRILLING? discovered.

JUST DISCOVERED
A FULL TRAP WITH
BOTH OIL AND GAS
G OIL

GAS
WATER
OIL O FILLED-TO-SPILL

O
GAS WATER
G G

O
OIL
G
NOT FILLED-TO-SPILL

OIL 2 OPTIONS

NOT FILLED-TO-SPILL

been no variation in seal strength, a simpler, idealized contribute significantly as well. A corollary of the
distribution might be expected (Figure 24). Had that present analysis is that deep class 2 and class 3 traps
been the case, we might have had a high degree of pre- can be expected to leak gas from traps in these as yet
dictability as to fill type prior to drilling; as with most undrilled older strata up into the commercially
simplistic solutions, this seems elusive. This explana- exploited shallower traps.
tion probably does not cover every trap. Class 1 type Thus, a significant proportion of Norway’s traps
reasoning is the probable cause in some cases. Exam- may have been exposed to an excess of both oil and
ples of differentially filled class 1 migration chains, gas. They equilibrated with a fill appropriate to their
such as suggested by Thomas et al. (1985) seem rea- closure vs. seal strength relationships. Traps such as
sonable. Nevertheless, an alternative explanation may Troll East were flushed with gas (class 1), traps such as
be necessary for many of Norway’s traps. Gullfaks and Ekofisk that couldn’t be flushed with gas
As a working hypothesis, many traps in Norway retained oil, while gas continues to leak through these
have been exposed to excess oil and gas. The main traps steady-state (class 3). A few traps, with Oseberg
source is Kimmeridge-equivalent Late Jurassic (hot) possibly being an example, reached seal strength with
shales. Thomas et al. (1985) suggest “vast oversupply” a balanced fill, part oil and part gas. They now spill
in relation to the northern Viking Graben. Bernard and their oil but leak their gas (class 2).
Bastow (1991) suggest that “...the oil productivity of
the Kimmeridge Formation in the Central Graben,
Moray Firth, and southern Viking Graben areas of the DISCUSSION
North Sea is estimated to be >250 billion bbl, of which
25% is accounted for in traps discovered to date.” If the best basins are characterized by overproduc-
These authors (in their figure 4) suggest oil productiv- tion of both oil and gas, and still contain oil, they are
ity in excess of 100,000 bbl/acre in depocenters. likely to do so because they contain traps that leak gas
Kimmeridge shales are also a main source of gas in while retaining oil (class 2 and especially class 3 traps).
gas kitchens along the axis of the Viking Graben Thus, while the volumetrics of source bed richness
(Thomas et al., 1985) and Central Graben (Cayley, and maturation, expulsion efficiency, and details of
1987). Kimmeridge gas potential is augmented, possi- migration may be paramount in basins with margin-
bly even exceeded, by the combination of multiple ally developed hydrocarbon systems, closure vs. seal
gas-prone source rocks at deeper stratigraphic levels strength relationships may dominate in determining
and by biogenic gas. This deeper gas source potential fill character in the best (oversupplied) basins. To the
exists in the Middle Jurassic Brent Formation coals and extent that these rich basins dominate world supply
Lower Jurassic Statfjord Formation coals. Addition- and determine the world mix of oil vs. gas, it is possi-
ally, Triassic and especially Carboniferous units have ble that this relationship dominates in determining the
gas-producing potential in England, Germany, Green- world mix of oil vs. gas.
land, and Svalbard. Although too deep for exploration Unfortunately, the principle, even if one considers it
in the Norwegian offshore, these deeper sources must completely valid, does not lead easily to predictivity.
80 Sales

Figure 21. Exploration


scenario when a trap with
GAS PROBLEMS
CLASS 1 (FILLING) oil that appears not to be
WHILE DRILLING? filled-to-spill is discovered.
OR CLASS 3
WATER
O AT EQUILIBRIUM
JUST DISCOVERED
A TRAP PARTLY FULL
G WATER
OF OIL

OIL
NOT FILLED
WATER

G
GAS
O WATER
OIL O

OIL

GAS 2 OPTIONS

OIL O

% Oil in Column In the few places analyzed, it seems that there is high
0 20 40 60 80 100 variability in seal quality, even in adjacent traps. This
does not seem to be related to depositional factors, but
14
50 to diagenetic and especially tectonic imperfections
25A,29A
implanted later and unique to each trap. The Wood-
35(K) ford Shale, for example, a recognized sourcing and
sealing formation of the U.S. midcontinent, contains
Distance from Gas Gen. (Km)

28A calcite concretions >5 ft (1.5 m) in diameter (Kirkland


34C 26
40 et al., 1992). These concretions could disrupt seal conti-
nuity and provide failure paths. Nevertheless, the
dominant reason is likely to be the variable tectonic
stressing of the individual trap’s seal.
34B After the referenced statistical study, I felt the prin-
30
ciples were sound but that seal strength wasn’t consis-
tent and/or spillpoints were not being treated
consistently. Yet considerable data (Sneider and
Neasham, 1993) suggested great consistency in the
32A
20 21
32B,28 mercury injection test data. The alternative seemed
that this consistent data was not testing actual seal
31
strength; it tested the perfect pores and not the cracks
7 in the seals. Yet Sneider and Neasham’s data (1993)
22 could not work as well as it does if it were not the cru-
10 33
cial data in the fields that they are dealing with.
19 2,3,8,16, 11,10 Uplift may be the key. The fields that Sneider and
36 17,18,20,25 12,9,6,2,4 13 Neasham (1993) are dealing with are mostly in the
23 24
27 Gulf Coast, and the seals have had a history of contin-
0 ual burial. Shales are created by and attuned to over-
burden pressure and remain plastic as long as burial is
Figure 22. Cross-plot of distance (near edge of continuous. Under these conditions, the sedimento-
Kimmeridge gas generation to the near edge of trap) logical pore-throat size, as reflected in the mercury
vs. fill character for the Norwegian North Sea. Actual injection tests, should be the critical parameter. In Nor-
migration distances for Kimmeridge-generated gas way, on the other hand, there has been uplift of indi-
could have been considerably greater, but migration vidual structures by footwall uplift during rifting and
distances for gas generated from deeper horizons could by the later Neogene uplift of the entire area that cur-
have been less. Note lack of relationship between fill rently uplifts Norway. This may have made seals brit-
character and distance from gas generation. tle and fractured. Fractures, rather than the perfect
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 81

trap-configuration domains the sedimentological pore


Oil Fields 0 Gas Fields size is critical and in which the crack-induced pore size
is critical.

CONCLUSIONS
Interplay among closure, seal strength, and buoy-
ancy provides a fundamental control on proportions of
oil and gas in traps, migration chains, basins, and possi-
bly the world. In a trap exposed to excess of both fluids,
closure vs. seal strength relationships are the dominant
control on fill character. Three trap classes are based on
1 this relationship, using the height of the longest possi-
ble gas column and the deeper longest possible (oil-
dominated) column as thresholds. Position of
34 spillpoints in relation to these thresholds determines
trap class. Each class has different fill character and
implications for traps updip and at shallower strati-
Depth (m x 1000)

15 graphic levels. Conversely, shallow traps give clues


16 about traps downdip or at deeper stratigraphic levels.
26,32 Class 1 (gas-prone) traps have excess seal strength,
17 and class 3 (oil-prone) traps have a deficiency of seal
18,20 strength relative to closure. Class 2 (two-phase-prone)
33
2 traps have intermediate seal strength relative to clo-
19,37 sure. Class 1 traps will contain gas; class 2 will contain
14 21,34 both gas and oil; and class 3 will contain dominantly
27,28 oil, even when exposed to excess gas. Class 1 traps
spill gas and oil and leak neither; class 2 spill oil and
11,24,30 leak gas; and class 3 leak oil and gas, and spill neither.
7,29 Salient features are given in Table 1.
6 Fluid contacts can be spill-, charge-, or pressure
31 controlled. Pressure-controlled contacts equilibrate in
3 midtrap, despite movement of oil and gas through the
5,9 trap. With too much closure to allow gas flushing,
3 class 2 and class 3 traps have enhanced probability for
4,12 2 oil, even when exposed to excess gas, as long as they
8 are not so deep that oil is cracked to gas in the trap.
Thus, large closure and/or low seal strength are insur-
10 22,23 ance against gas flushing. Leakage of gas is required to
maintain room for oil in a trap exposed to excess gas.
36 Rather than being negative, signs of gas leakage
25 enhance the probability for oil.
Gussow (1954) considered all traps to be class 1.
While he made a landmark contribution, he dealt with
only one aspect of the total spectrum of trap processes
4 and thus oversimplified. Some of his examples appear
13 valid, some suspect. All are ripe for review in a
broader conceptual framework.
The northern North Sea, with too much exposure to
Figure 23. Graph of trap depth vs. fill character for gas and generally good migration potential, still has oil
the Norwegian North Sea. Note that the shallowest in the majority of traps. There are enough class 2 and
trap (Troll East) is gas and the deepest trap (7/11-5 class 3 traps along the migration pathways to vent suffi-
Discovery) is oil, and that there is no apparent rela- cient gas to increase the proportion of oil relative to gas
tionship between trap depth and fill character. Data in the area’s traps. In such areas, with overabundance of
from Spencer et al. (1987) keyed to field numbers in both fluids, present generation, and long migration
that volume. routes, trap processing may be paramount in control-
ling trap charge. In marginal areas, volumetrics may be
more important. In the North Sea, Troll East is class 1;
sedimentological pore, mayprovide the critical failure Oseberg is class 2; and Gullfaks, Snorre, Ekofisk, and
path in such areas. There is need for studies that both most of the chalk fields are class 3. (Examples in
test this hypothesis and define in which geographic or Wyoming, Alberta, and Indonesia are discussed.)
82 Sales

700 John Snedden, Nick Theis, and Joana Vizgirda com-


Hypothetical mented on early versions of this manuscript. A.G. Dorés
Distribution if editing improved a previously published version (Sales,
All Traps Had 1993). Anonymous reviewers of both versions helped
600
Same Seal Never Filled greatly. I thank them all, but retain responsibility for
errors that remain. I thank Mobil Oil Corporation for
Capacity
permission to publish this paper.
500
REFERENCES CITED
HC Column Height (M)

All
Class 3
Allen, U.S., 1989, Model of hydrocarbon migration and
Oil Traps
400 entrapment within faulted structures: AAPG
Mixed Traps Bulletin, v. 73, p. 803–811.
Class 2
Berg, R.R., 1975, Capillary pressure in stratigraphic
traps: AAPG Bulletin, v. 59, p. 939–956.
300 Bernard, P.C., and M.A. Bastow, 1991, Hydrocarbon
Capacity Gas Column generation, migration, alteration, entrapment and
mixing in the Central and Northern North Sea, in
W.A. England and A.J. Fleet, eds., Petroleum migra-
200 tion: Geological Society of London Special Publica-
Gas Traps
(Class 1)

tion 59, p. 167–190.


Caillet, G., 1993, The caprock of the Snorre Field, Nor-
way: a possible leakage by hydraulic fracturing:
100 Marine and Petroleum Geology, v. 10, p. 42–50.
Cayley, G.T., 1987, Hydrocarbon migration in the cen-
tral North Sea, in J. Brooks and K. Glennie, eds.,
Petroleum geology of northwest Europe: London,
0
Graham & Trotman, p. 549–555.
0 20 40 60 80 100
Dangerfield, J., I. Knight, and H. Farrell, 1992, Character-
% of Oil by Column Height ization of faulting and fracturing in Ekofisk Field from
seismic, log, and core data; in R.M. Larsen, H. Brekke,
Figure 24. Idealized pressure-depth plot of relations B.T. Larsen, and E. Talleraas, eds., Structural and tec-
if all traps had the same seal strength and ample tonic modelling and its applications to petroleum
access to both fluids. geology: Norwegian Petroleum Society Special Publi-
cation 1: Amsterdam, Elsevier, p. 397–407.
Dembicki, H., and M.J. Anderson, 1989, Secondary
Trap classes respond differently to uplift. Assuming migration of oil: experiments supporting efficient
constant seal strength, class 1 may be gas-flushed as movement of separate, buoyant oil phase along lim-
gas expands in the trap and in traps downdip. Class 2 ited conduits: AAPG Bulletin, v. 73, p. 1018–1021.
should improve, gaining a higher oil column at the Dobbin, C.E., 1947, Exceptional oil fields in the Rocky
expense of the gas cap, as leakage of the less-dense gas Mountain region of the United States: AAPG
accelerates. Class 3 might modestly improve if the Bulletin, v. 31, p. 797–823.
only gas was in solution and it is forced out of solution Downey, M.W., 1984, Evaluating seals for hydrocar-
by uplift. A different scenario applies if seal strength bon traps: AAPG Bulletin, v. 68, p. 1752–1763.
decreases during uplift. Durand, B., 1988, Understanding of HC migration in
Prediction on many scales based on these concepts sedimentary basins (present state of knowledge):
seems possible, although working with the data is Organic Chemistry, v. 13, p. 445–459.
inherently difficult. To the extent that oversupply of Duval, B.C., G.C. de Janvry, and B. Loir, 1992, Detailed
both fluids characterizes the world’s best provinces, geoscience reinterpretation of Indonesia’s Mahakam
the provinces that dominate the world’s supply, clo- Delta scores: Oil and Gas Journal, August 10, p. 67–71.
sure vs. seal strength relationships may dominate in England, W.A., A.S. Mackenzie, D.M. Mann, and T.M.
determining the world’s mix of trapped oil and gas. Quigley, 1987, The movement and entrapment of
More practically, a knowledge of the characteristics of hydrocarbon fluids in the subsurface: Journal of the
the trap classes may influence exploration rationale Geological Society of London, v. 144, p. 327–347.
and the order in which targets are drilled. Eriksen, T., M. Helle, J. Henden, and A. Rognebakke,
1987, Gullfaks, in A.M. Spencer, E. Holter, C.J. Camp-
ACKNOWLEDGMENTS bell, S. Hanslien, P.H.H. Nelson, E. Nysæther, and E.
Ormaasen, eds., Geology of the Norwegian oil and
Steve Buck, George Claypool, Jim Dixon, Pete gas fields: London, Graham & Trotman, p. 273–286.
Hansen, Jim Helwig, Doug Kirkland, Gene Leavitt, Pam Fermor, P.R., and I.W. Moffat, 1992, Tectonics and
Luttrell, Randy Martinsen, Tim Matava, George Moekel, structure of the Western Canada foreland basin, in
Seal Strength vs. Trap Closure—A Fundamental Control on the Distribution of Oil and Gas 83

R.W. MacQueen and D.A. Leckie, eds., Foreland Sales, J.K., 1993, Closure vs. sealing strength—a funda-
basins and fold belts: AAPG Memoir 55, p. 81–106. mental control on the distribution of oil and gas:
Foster, N.H., and E.A. Beaumont, compilers, 1988, Basin modeling: advances and application: Norwe-
Traps and seals: AAPG Treatise of Petroleum Geol- gian Petroleum Society Special Publication 3,
ogy, Reprint series 7, 2 vol., 42 papers, 1064 p. p. 399–414.
Fox, F.G., 1959, Structure and trap of hydrocarbons in Schowalter, T.T., 1979, Mechanics of secondary hydro-
southern Foothills, Alberta, Canada: AAPG carbon migration and entrapment: AAPG Bulletin,
Bulletin, v. 43, p. 992–1025. v. 63, p. 723–760.
Gray, I., 1987, Troll, in A.M. Spencer, E. Holter, C.J. Schowalter, T.T., 1991, Secondary hydrocarbon migra-
Campbell, S. Hanslien, P.H.H. Nelson, E. Nysæther, tion in basin modelling: International conference on
and E. Ormaasen, eds., Geology of the Norwegian basin modelling; advances and applications (abs.):
oil and gas fields: London, Graham & Trotman, Norwegian Petroleum Society, p. 23.
p. 389–401. Skerlec, G.M., 1993, Seals: risking fault seal and top
Gussow, W.C., 1954, Differential entrapment of oil and seal: PetroQuest International Inc., Informal Course
gas: a fundamental principle: AAPG Bulletin, v. 38, Notes, Mobil Oil Corp., Dallas, Texas.
p. 816–853. Sneider, R.M., and J. Neasham, 1993, Comparison of
Hoveland, M., and A.G. Judd, 1988, Seabed pock- seal strength determinations: cores vs. cuttings
marks and seepages: London, Graham and Trot- (abs.): AAPG Seals and Traps Conference, Abstracts
man, 293 p. with Program, Addendum, June 21–23, 1993,
Irwin, 1929, quoted in W.C. Gussow, 1954, Differential Crested Butte, Colorado.
entrapment of oil and gas: a fundamental principle: Spencer, A.M., and V.B. Larsen, 1990, Fault traps in the
AAPG Bulletin, v. 38, p. 131. northern North Sea, in R.F.P. Hardman and
Hubbert, M.K., and D.G. Willis, 1957, Mechanics of J. Brooks, eds., Tectonic events responsible for
hydraulic fracturing: Petroleum Transactions of the Britain’s oil and gas resources: U.S. Geological
American Institute of Mining and Metallurgical Society Special Paper, v. 55, p. 251–298.
Engineers, v. 210, p. 153–168. Spencer, A.M., E. Holter, C.J. Campbell, S. Hanslien,
Karlsson, W., 1986, The Snorre, Statfjord and Gullfaks P.H.H. Nelson, E. Nysæther, and E. Ormaasen, eds.,
oil fields and the habitat of hydrocarbons on the 1987, Geology of the Norwegian oil and gas fields:
Tampen Spur, offshore Norway, in A.M. Spencer et London, Graham & Trotman, 493 p.
al., eds., Habitat of hydrocarbons on the Norwegian Thomas, B.M., P. Moller-Pedersen, M.F. Whitaker, and
Continental Shelf: Norwegian Petroleum Society, N.D. Shaw, 1985, Organic facies and hydrocarbon dis-
London, Graham & Trotman, p. 181–197. tributions in the Norwegian North Sea, in B.M.
Kirkland, D.W., R.E. Denison, D.M. Summers, and J.R. Thomas et al., eds., Petroleum geochemistry in explo-
Gormly, 1992, Geology and organic geochemistry of ration of the Norwegian Shelf: London, Graham &
the Woodford Shale in the Criner Hills and western Trotman, p. 3–26.
Arbuckle Mountains, Oklahoma: Oklahoma Geo- Tissot, B., 1987, Migration of hydrocarbons in sedimentary
logical Survey Circular 93, p. 38–69. basins: a geological, geochemical and historical per-
Leith, T.L., 1993, Caprock leakage on the Tampen spective, in B. Doligez, ed., Migration of hydrocarbons
Spur, Norwegian North Sea: IKU Poster at AAPG in sedimentary basins: Paris, Editions Technip, p. 1–20.
Traps and Seals Conference, June 21–23, 1993, Ungerer, P., J. Burrus, B. Doligez, P.Y. Chenet, and
Crested Butte, Colorado. F. Bessis, 1990, Basin evaluation by integrated two-
Leith, T.L., I. Kaarstad, J. Connan, J. Pierron, and G. dimensional modeling of heat transfer, fluid flow,
Caillet, 1993, Recognition of caprock leakage in the hydrocarbon generation, and migration: AAPG
Snorre Field, Norwegian North Sea: Marine and Bulletin, v. 74, p. 309–335.
Petroleum Geology, v. 10, p. 29–41. Van den Bark, E., and O.D. Thomas, 1981, Ekofisk: first
Link, T.A., 1949, Interpretations of Foothills structures, of the giant oil fields in western Europe: AAPG
Alberta, Canada: AAPG Bulletin, v. 33, p. 1475–1501. Bulletin, v. 65, p. 2341–2363.
Masters, J.A., ed., 1984, Elmworth—case study of a Watts, N.L., 1983, Microfractures in chalks of
deep basin gas field: AAPG Memoir 38, 316 p. Albuskjell Field, Norwegian Sector, North Sea: pos-
Miles, J.A., 1990, Secondary migration routes in Brent sible origin and distribution: AAPG Bulletin, v. 67,
sandstones of the Viking Graben and East Shetland p. 201–234.
Basin: AAPG Bulletin, v. 74, p. 1718–1735. Watts, N.L., 1987, Theoretical aspects of cap-rock and
Nipen, O. 1987, Oseberg, in A.M. Spencer, E. Holter, fault seals for single- and two-phase columns:
C.J. Campbell, S. Hanslien, P.H.H. Nelson, E. Marine and Petroleum Geology, v. 4, p. 274–307.
Nysæther, and E. Ormaasen, eds., Geology of the Xiaowen, F., L. Chatzis, and A.L. Dullien, 1992, An
Norwegian oil and gas fields: London, Graham & experimental study of oil migration: AAPG
Trotman, p. 379-388. Bulletin, v. 76, p. 638–650.
Porter, J.W., 1992, Early surface and subsurface inves- Ziegler, D.L., 1992, Hydrocarbon columns, buoyancy
tigations of the Western Canada Sedimentary Basin, pressures and seal efficiency: comparisons of oil
in R.W. MacQueen and D.A. Leckie, eds., Foreland and gas traps in California and the Rocky Mountain
basins and fold belts: AAPG Memoir 55, p. 125–158. area: AAPG Bulletin, v. 76, p. 501–508.
Kaldi, J.G., and C.D. Atkinson, 1997, Evaluating seal potential:
example from the Talang Akar Formation, offshore north-
west Java, Indonesia, in R.C. Surdam, ed., Seals, traps, and
the petroleum system: AAPG Memoir 67, p. 85–101.

Chapter 6

Evaluating Seal Potential: Example from the


Talang Akar Formation, Offshore Northwest
Java, Indonesia
J.G. Kaldi
ARCO Indonesia
Jakarta, Indonesia

C.D. Atkinson
ARCO British Ltd.
Guildford, United Kingdom


ABSTRACT
The seal potential of various lithologies in the Upper Oligocene Talang Akar
Formation (TAF) is evaluated in the BZZ area of offshore northwest Java. Seal
potential comprises (1) seal capacity (the calculated amount of hydrocarbon
column height a lithology can support); (2) seal geometry (the structural posi-
tion, thickness, and areal extent of the lithology); and (3) seal integrity (rock
mechanical properties such as ductility, compressibility, and propensity for
fracturing). Seal capacity is determined by mercury injection capillary pressure
(MICP) analyses. Seal geometry is derived by integrating seismic data, core,
detailed well correlations, regional sedimentological/stratigraphic relation-
ships, and comparisons to known depositional analogs. Seal integrity is evalu-
ated qualitatively by core examination, borehole imaging, and petrographic
studies. These three variables were integrated and the totals were “ranked.” In
the BZZ area, deltaic distributary channel sandstones and delta-front/mouth
bar heterolithic sandstones comprise the main reservoirs. Possible seals include
prodelta, delta-front, and delta-plain shales; channel abandonment silts; and
transgressive shelf carbonates in both the upper and lower TAF. Seal potential
is best in the delta-front shales, which have high seal capacity and are thick, lat-
erally continuous, and very ductile. Seal potential is moderate in the thicker
(upper TAF) transgressive carbonates. These rocks have high seal capacity and
excellent lateral continuity, but are brittle and, hence, prone to fracturing. Delta-
plain shales and prodelta shales are poor seals due to their limited seal capacity
(delta-plain) or because they are too thin (prodelta shales). Channel abandon-
ment siltstones have even poorer seal potential because of small lateral extent
and limited seal capacity. The least favorable seal potential occurs within the
thin (lower TAF) carbonates. These rocks are relatively thin, as well as being
prone to fractures.

85
86 Kaldi and Atkinson

INTRODUCTION seal thickness and seal effectiveness. Another impor-


tant factor in seal potential is the structural position of
Recent drilling successes in the Talang Akar Forma- the sealing lithology. All other variables being equal, a
tion (TAF) in the offshore northwest Java (ONWJ) lithology on the crest of a tightly folded structure has a
region have proved the formation to be an important higher probability of having developed tensile frac-
and viable exploration target. To date, most Talang tures, leading to hydrocarbon escape, than the same
Akar prospects have been drilled on the basis of struc- lithology on a broad, open structure.
tural closure, often as extensions to wells drilled for The seal potential of a lithology is a combination of
stratigraphically higher objectives. Few stratigraphic the calculated amount of hydrocarbon column height
plays have been drilled, despite the fact that the depo- supported (seal capacity); the structural position,
sitional origin of the formation, particularly that of the thickness, and areal extent of the lithology (seal geom-
main reservoir sandstones (i.e., deltaic channel bod- etry); and rock mechanical properties such as ductility
ies), favors stratigraphic trapping. Maxus’s (1988) dis- (seal integrity). Seal capacity is measured in the labo-
covery of the large Widuri field in the nearby Asri ratory by using mercury injection capillary pressure
basin confirms this belief (Young et al., 1991). This (MICP) analysis of representative rock samples. Seal
field, which contains in-place reserves >500 MBO, geometry can best be determined by integrating seis-
comprises a succession of stacked, interconnected mic data, borehole information, regional sedimento-
channel sandstone reservoirs trapped by a combina- logical/stratigraphic relationships, detailed well
tion of gentle, three-way fault dip closure and updip correlations, and comparisons to known depositional
sandbody pinch-out (Young et al., 1991). analogs. Seal integrity can generally be evaluated from
Normal faulting has resulted in a variety of litholo- rock mechanics studies and/or petrographic analysis.
gies positioned updip against the reservoir intervals:
which of the many potential sealing lithologies, juxta- Study Objectives
posed against which reservoir units, provide the most
favorable reservoir-seal couplets? In addition, there is The overall goal of this study is to evaluate the
good stratigraphic prospectivity within the TAF potential variability in seal quality within the TAF in
because the reservoir sandstones are interbedded with the BZZ field and its applicability to understanding
good source rocks and a variety of potential sealing reservoir/seal relationships in the TAF throughout the
lithologies. This untested stratigraphic potential, in ONWJ area. This goal was subdivided into the follow-
combination with the excellent reservoir quality of the ing specific objectives:
sandstones, makes the Talang Akar one of the few
remaining plays in the highly explored ONWJ basin • to subdivide potential seal lithologies within the
that may contain sizable recoverable reserves. In order TAF, according to depositional origin;
to optimize drilling success, a comprehensive knowl- • to combine the sealing capacity results with thick-
edge of the geological characteristics of the formation ness and estimated areal extents of each lithology
is a necessity at this stage in exploration. to determine overall seal potential;
In this study, an assessment is made of the sealing • to develop criteria for ranking the seal potential of
potential of the various lithologies known to exist various lithologies in the TAF; and
within the formation. A similar study (Noble et al., • to provide recommendations concerning seal qual-
1991) uses geochemical methods for testing whether ity and distribution to guide future exploration of
seal capacity has been exceeded. TAF reservoirs throughout the ONWJ area.

Seals and Seal Potential


Database and Methodology
Any lithology can form a seal in the subsurface,
provided that the minimum displacement pressure of The database for the study comprises >20% explo-
the potential seal rock is greater than the established ration and development wells from the BS-BTS-BZZ
buoyancy pressure of the hydrocarbons within the area of the central Ardjuna basin, ONWJ (Figure 1;
accumulation (Schowalter, 1979; Sneider, 1987; Vavra Table 1). In this area, detailed core studies of >76 m
et al., 1992). The capacity of a seal to hold back hydro- (250 ft) of conventional core and comprehensive
carbons is thus controlled by the size of the largest interwell correlations have been carried out, primar-
interconnected continuous pore throats and the rela- ily in the BZZ field (Suria, 1991). Since this area is
tive densities of the hydrocarbon (oil or gas) and for- one of the few to date in which TAF reservoirs are
mation water. For a seal to be truly effective, however, under development, it represents the only area in the
it also needs to be relatively thick, laterally continuous, ONWJ region where enough data are present to be
relatively homogeneous, and fairly ductile (Downey, used as the basis for a study of this type. In both the
1984). Murris (1980) documents the interdependence BTS and BZZ fields, the main hydrocarbon accumu-
of these factors in his discussion of the Upper Jurassic lations occur within the uppermost part of the
Hith anhydrite and Albian Nahr Umr shale seals of the deltaic TAF member. While most available conven-
Middle East. In their appraisal of more than 160 tional core has been obtained from this stratigraphic
prospects, Sluijk and Nederlof (1984) and Sluijk and level, well-logs and detailed correlation sections
Parker (1986) find an empirical relationship between from the entire succession were used.
Evaluating Seal Potential: Example from the Talang Akar Formation 87

Figure 1. Location of BS, BTS, and BZZ fields within central Ardjuna region of ARII’s offshore northwest Java
(ONWJ) contract area.

The methodology of the study first involved using cycle). Routine MICP analyses involve a time-deter-
sedimentological criteria to determine the depositional mined incremental increase of injection pressure.
environment for all potential seal and reservoir However, the extremely low porosity and perme-
lithologies observed within the conventional cores. ability of most of these samples (Table 1) caused
Comparisons to wells from other areas suggest that injection rates to be very slow. Therefore, injection
the interpreted environments are typical for the TAF pressures were held at each pressure point and
throughout the ONWJ region. Thus, any conclusions increased only after equilibration had been reached
developed on the basis of the BS-BTS-BZZ data should at that point. This method is recommended by Snei-
be applicable on a regional scale. der (1987).
From each determined lithology or “lithofacies,” a
series of representative 1-in. (2.5-cm) diameter plug
samples were cut and used for plug porosity/perme- STRATIGRAPHY AND DEPOSITIONAL
ability and MICP analysis. Helium porosity and per- SETTING OF THE TALANG AKAR
meability to nitrogen gas were measured on each FORMATION
sample at room temperature and under a confining
pressure of 500 psi. The measured permeability of The Talang Akar Formation (TAF), along with the
some samples has been affected by natural fractures overlying Batu Raja carbonates, comprises the lower
and/or partings shown on Table 1. Mercury injection portion of the Cibulakan Formation in the ONWJ
capillary pressure (MICP) testing was performed by region (Figure 2). In most previous studies, the Talang
RUI Analytical by using the Micromeritics AutoPore Akar has been assigned formation status and has been
9201 mercury injection machine and specialized soft- subdivided into a series of informal members based
ware developed by RUI Analytical. All samples loosely upon depositional environment. These are,
underwent mercury injection to 30,000 psi (drainage from youngest to oldest: marine Talang Akar, deltaic
88 Kaldi and Atkinson

Table 1. Sample Database.

Depth Porosity Permeability Grain


Sample Well (ft) Facies* (%) (md) Density
1 BZ-2 7264.9 CAS 4.70 2.24** 2.73
2 BZ-2 7270.3 CAS 3.25 0.25 2.67
3 BZ-2 7262.3 CAS 4.75 3.26 3.01
4 BZ-2 7256.7 DCS 23.21 381.00 2.53
5 BZ-2 6983.6 SC 0.17 <0.001 2.83
6 BZ-2 6980.6 PDS 9.87 0.003 3.14
8 BZ-2 6970.8 DFS 5.67 1.05** 2.72
9 BZ-2 8058.3 DFS 1.29 0.003 2.73
10 BZZA-7 7898.0 DPS 2.10 0.008 2.67
11 BZZA-7 7897.0 DPS 1.05 0.54** 2.56
13 BZZA-7 7892.2 PDS 1.40 44.60** 2.79
14 BZ-2 6968.6 DMBS 9.77 2.00 2.71
17 KLN-1 8031.0 SC 0.66 <0.001 2.87

* CAS = Channel abandonment siltstones; DCS = Distributary channel sandstone (reservoir rock); SC = Shelf carbonate; PDS = Prodelta shale;
DFS = Delta-front shale; DPS = Delta-plain shale; and DMBS = Distributary mouth bar sandstone (reservoir rock).
** Fracture present in sample.

Talang Akar, and the so-called Talang Grits. The lim- coals upward through the succession generally marks
ited amount of available biostratigraphic data pro- the transition from deltaic to marine Talang Akar.
hibits a more refined and accurate subdivision of the Shales and siltstones comprise the dominant lithologies
formation at this time. However, the data do confirm of the succession and have been interpreted to repre-
that the sediments accumulated from at least the late sent a variety of subenvironments. Channel abandon-
Oligocene until the early Miocene. ment siltstones typically occur at the top of most
The general vertical sequence through the forma- distributary channel fills. If regression occurred, these
tion is transgressive. Largely nonmarine deposits at siltstones commonly grade upward into paleosol-
the base grade upward into transitional deltaic sedi- modified delta-plain shales. In contrast, where aban-
ments, which in turn give way to increasingly more donment was associated with transgression, delta-
marine/shelfal deposits at the top (Figure 3). Final pas- front, drifted coals, thin prodelta shales, and limestones
sage into the overlying Batu Raja carbonates appears may have developed above the siltstones.
to reflect the local cessation in deltaic supply and, con-
sequently, is interpreted as a diachronous contact Reservoir Lithofacies
throughout the region (Figure 3, lower left). Most Two main reservoir rock units occur within the
hydrocarbon accumulations to date have been discov- Talang Akar succession: distributary channel and
ered in the marine and deltaic Talang Akar. Conse- delta-front distributary mouth/tidal bar sandstones.
quently, all the available conventional cores upon
which the study is based have been obtained from Distributary Channel Sandstones
these two members. These sandstone bodies range from 10 to 27 m
Sedimentological studies, in combination with inter- (30–90 ft) in thickness and are common throughout the
pretations based upon wireline log shape, reveal that deltaic Talang Akar succession. In core (Figure 5B),
the marine and deltaic Talang Akar successions com- they possess either fining-upward or uniform, vertical
prise an interbedded sequence of sandstones, shales, grainsize profiles which, on wireline logs, are reflected
siltstones, coals, and thin limestones (Figure 4). Adopt- by either bell-shaped or blocky log-curves. Both single
ing the present-day Mahakam Delta of eastern Kali- and multistory bodies have been recognized, the latter
mantan as a depositional analog (Allen et al., 1979; of which tend to have thicknesses >12 m (35 ft). The
Allen, 1987), the sandstones are interpreted as the sandstones are fine to coarse grained, and in places
deposits of both deltaic distributary channels and may be locally conglomeratic. Thin shale beds are
coarsening-upward, delta-front/tidal bars. The lime- common within the bodies, especially toward the tops
stones are of shallow marine origin and, when present of the fills (Figure 4). The dominant sedimentary struc-
in the deltaic member, signify the temporary abandon- ture is small- to medium-scale trough cross-stratifica-
ment of the delta during periods of either reduced or tion, the foresets of which (in some of the more tidally
no clastic sediment supply. At least two coal types are influenced channels) are often highlighted by thin clay
recognized: “in-situ” coals associated with an underly- and organic-rich drapes.
ing sear earth, or paleosol (fossil soil), and “drifted” Based upon detailed correlations in the BZZ field
coals resulting from the accumulation of large amounts and comparison to the present-day Mahakam Delta,
of plant debris in a waterlain environment. The loss of the channel sandstones appear to comprise elongate,
Evaluating Seal Potential: Example from the Talang Akar Formation 89

Figure 2. Generalized
stratigraphy of offshore
northwest Java, with
eustatic sea level variation
over the Late Paleogene and
Neogene (modified from
Haq et al., 1987). Note
overall eustatic sea level
rise during upper Talang
Akar–lower Batu Raja depo-
sition (30–25 Ma), which
results in a generally trans-
gressive vertical
succession through the
sequence.

elliptical-shaped bodies ranging from 0.5 to 1.5 km interpret in the subsurface. They comprise 1–5-m- (3–15-
(1500 ft–1 mi) in width and ≥1 km (≥1.7 mi) long. The ft) thick coarsening- and thickening-upward packages of
sandstones represent former side, median, and lateral very fine to fine grained sandstones, siltstones, and
bar deposits within the distributaries and, as such, shales (Figure 5C). On wireline logs, they are character-
often exhibit complex lateral geometries. This is espe- ized by funnel-shaped curves. Internally, the sandstones
cially the case in those bodies characterized by pro- commonly show intense bioturbation and exhibit vari-
nounced multistory stacking, where several channel able degrees of calcite cementation. In many cases, the
units can be juxtaposed, producing a thicker and more uppermost part of the body is overprinted by later soil
areally extensive sandstone accumulation. Sample 4 formation (rootlet traces, etc.) and is capped by a thin
typifies the distributary channel sandstones of the coal (Figure 4). Because of the disruption caused by the
BZZ area. burrows and rootlets, physically produced sedimentary
structures are rare. However, where they are present,
Delta-Front Distributary Mouth/Tidal Bar small-scale ripple lamination is dominant, with subordi-
Sandstones nate amounts of plane-bed, horizontal lamination.
In comparison to the distributary channel sandstones, In the present-day Mahakam Delta, such bodies
these bodies are less numerous and more difficult to possess at least two types of geometry, depending
90 Kaldi and Atkinson

Figure 3. Schematic representation of Talang Akar sequence development. Each


sequence represents a pulse of deltaic progradation into an essentially shallow
marine transgression. Following delta lobe abandonment, relative sea level rise
initiates transgression. Inset (lower left) demonstrates diachronous nature of
Talang Akar and Batu Raja contact in the region. Larger part of the figure is
detail of inset. Note overall upward decrease in delta-plain and delta-front
muds and silts (stippled pattern) and increase in marine shales (unornamented
area) and carbonates (brick pattern). Coarse dots are distributary channel sand-
stones; dark lines are in-situ coals.

upon their depositional origin. In the lower, tidally is inversely proportional to the thickness of the under-
influenced parts of some distributaries, tidal bars lying channel sandstone, thin (<3 m; 10 ft) where the
develop in median locations within the flared channel sandstone is thick, and thick (≤16 m; ≤50 ft) where the
mouth. They are laterally restricted, because of the sandstone is thin. Since it represents the abandonment
channelized nature of the distributary mouth, and deposits of the former channel course, this lithofacies
tend to range from 0.5 to 1.5 km (1500 ft–1 mi) in width is expected to possess a relatively narrow, elongate,
and ≤3 km (≤2 mi) long. In contrast, where the distrib- sinuous geometry. Based on the dimensions of the
utaries open onto the delta front, flow deceleration associated channel sandstone, such geometries will be
promotes the accumulation of tidal/mouth-bar sand- laterally restricted, ranging from hundreds to thou-
stones with far greater lateral extents. In some places sands of square meters wide.
on the Mahakam Delta, such bodies can have areal
dimensions in excess of several square kilometers. Lat- Delta-Plain Shales
eral juxtapositioning of both types of bar sandstone is This seal lithofacies is common within the deltaic
expected, given the shifting nature of these deposi- member of the TAF, but is rare, if not absent, in the
tional environments. Thus, in the subsurface, the upper marine member. It is made up of mottled, gray-
resulting sandstone bodies may be even more exten- green-purple–colored shales and minor siltstones with
sive than those observed in the present-day delta. obvious carbonized rootlet traces (Figure 5D). Inter-
nally, it often displays distinctive slicken-sided frac-
Potential Sealing Lithofacies tures, which are interpreted to represent former soil
Five potential sealing lithofacies have been recog- “ped” textures (“ped” is a soil-science term that
nized and are displayed in Figure 5: channel abandon- describes the tendency of a soil to develop an internal
ment siltstones, delta-plain shales, delta-front shales, blocky fabric that is produced through alternate
prodelta shales, and shelfal carbonates. swelling and contracting due to repeated wetting and
drying). The delta-plain shales tend to develop above
Channel Abandonment Siltstones the channel abandonment siltstones, and are often
As the name implies, this lithofacies occurs exclu- associated with in-situ coal development. Where
sively in association with distributary channel com- observed in core, the shales tend to range from 3 to 15
plexes. It consists of siltstones and shales with varying m (10–50 ft) in thickness. Based upon correlations in
proportions of interbedded, thin, very fine grained the BZZ field, lateral extents are expected to be in the
sandstones (Figure 5A). The thickness of this lithofacies range of 1 to 10 km2 (<10 mi2).
Evaluating Seal Potential: Example from the Talang Akar Formation 91

Figure 4. Generalized
interpretation of Talang
Akar depositional environ-
ments from BZZ cores. Note
isolated reservoir sand bod-
ies encased in shales of vari-
ous origins.

Delta-Front Shales silt-sand content of the shales. In many cases, the sandier
Delta-front shales occur in both the deltaic and intervals are characterized by extensive calcite and/or
marine members of the TAF, but are especially common dolomite cementation. The shales often tend to grade
and more numerous in the latter. The shales are dark vertically upward into either delta-front/tidal-bar sand-
gray to black and often contain admixtures of organic stones or, in some cases, into thin, in-situ, or drifted coals.
matter (Figure 5E). Siltstones and thin interbeds of very In core, the shales range from 3 to 15 m (10–50 ft) in thick-
fine grained sandstone occur locally. Bioturbation is ness and, based upon the modern Mahakam Delta ana-
ubiquitous and becomes more apparent the higher the log, are interpreted to possess lateral extents of 1–10 km2.
92 Kaldi and Atkinson

Figure 5. Examples of various reservoir and seal lithofacies in core from the BZZ area.

Prodelta Shales occur within an argillaceous matrix. The abundant


Shales of this type are relatively rare in the TAF clay content of these carbonates may provide a degree
sequences. They are intergradational with the associ- of ductility to this lithofacies not commonly associated
ated delta-front shales, but are distinguished by their with carbonates. In the lower deltaic member, this litho-
darker color, lower silt content, absence of sandstone, facies is quite thin (<3 m; 10 ft), whereas in the upper
and presence of siderite nodules or bands (Figure 5F). marine interval, such carbonates can range up to 15 m
Where developed, they are frequently quite thin (<3 (50 ft) in thickness. From correlations in the BS-BTS-
m; 10 ft) and sometimes contain concentrations of BZZ region, the carbonates tend to be locally extensive
drifted coal. By analogy to the modern-day Mahakam and to cover areas in excess of several 10 × km2.
Delta, such shales are expected to be quite laterally
extensive and to cover areas of 1 to 10 km2. SEAL GEOMETRY RESULTS
Shelfal Carbonates In terms of geometry, the shelfal carbonates and
Shelfal carbonate lithologies develop preferentially delta-front shales of the marine and upper deltaic TAF
in the upper part of the deltaic TAF member, but reach and the delta-plain shales of the lower deltaic TAF are
their optimum development in the overlying marine the thickest (3–15 m; 10–50 ft) and most extensive (sev-
member. The carbonates vary from very fine grained, eral km2) of the various seal lithofacies. The shelfal car-
argillaceous, dolomite mudstones (Figure 5G) to skele- bonates associated with the lower deltaic TAF and the
tal lime wackestones. In the more lime-rich wacke- prodelta shales are both quite extensive (several km2),
stones, allochems comprising fragments of coral, but probably have reduced seal effectiveness due to
echinoderms, bivalves, and larger foram tests also their relatively thin nature (<3 m; 10 ft). In contrast, the
Evaluating Seal Potential: Example from the Talang Akar Formation 93

Figure 5. Continued.

channel abandonment siltstones, although relatively contact angles of the brine-hydrocarbon-solid and air-
thick (1.5–15 m; 5–50 ft), are laterally very restricted mercury-solid systems, respectively (Vavra et al., 1992).
(100s m–1 km; 100s ft–0.6 mi) and, consequently, their Next, the relationship between the brine/hydrocar-
seal effectiveness is likely to be limited. These results bon capillary pressure and the height above the free
are summarized in Figure 6 and Table 2. water level (FWL) was calculated using the equation:

Seal Capacity Pc b/hc = h( ρ w – ρ hc ) × 0.433 (2)


Evaluation of seal capacity was done by performing
MICP analysis on each sample (both reservoirs and Seal capacity (the maximum hydrocarbon column
seals) up to an injection pressure of 30,000 psia mer- the seal can hold before it begins to leak) was calcu-
cury (Figure 7). Since the capillary pressure data were lated by:
measured in the air/mercury system, it was necessary
to convert to the brine/hydrocarbon system of the Pds – Pdr
reservoir using the equation: H max = (3)
0.433( ρ w – ρ hc )
σ b / hc Cosθ b / hc
Pc b/hc = Pc a/m × For equations 2 and 3, h = height above FWL; Hmax
σ a / m Cosθ a / m (1)
= maximum height of the hydrocarbon column held
(in feet); Pds and Pdr = the brine/hydrocarbon displace-
where P cb/hc = the capillary pressure in the brine- ment pressures of the seal and reservoir rock (in psi);
hydrocarbon system of the reservoir; Pca/m = the capil- ρw and ρhc = the densities of the reservoir fluids (water
lary pressure in the air-mercury system; σb/hc and σa/m = and hydrocarbon) in grams per cubic centimeter. PVT
the interfacial tensions of the brine-hydrocarbon and analyses indicate ρw and ρhc to be 1.01 and 0.85 (34.6°
air-mercury systems, respectively; θb/hc and θa/m = the API), respectively.
94 Kaldi and Atkinson

Figure 6. Geometries of
potential seal facies in the
BZZ area. Data from BZZ
area core, Mahakam Delta
analog, and literature.

Values for interfacial tension (s) and contact angle mination of actual subsurface rock and fluid proper-
(q) in the air-mercury system are 480 dynes/cm and ties. To account for some of these variables, sensitivi-
140° API, respectively (Core Laboratories, 1982; Snei- ties were tested for different brine/hydrocarbon
der, 1987; Vavra et al., 1992). For the Talang Akar fluid contact angles, interfacial tensions, and displacement
system, laboratory-determined values were obtained pressures. Thus, seal capacity was evaluated by using
from R.M. Sneider (personal communication, 1990). θb/hc 0°–45° API and σb/hc from 20 to 35 dynes/cm, with
The results for the reservoir brine/hydrocarbon sys- weighting toward the laboratory-determined most
tem are: contact angle (θb/hc) = 10° API and interfacial likely values (MLV).
tension (σb/hc) = 30 dynes/cm. Even with the best labo- The critical parameter in the calculations is the
ratory data, however, calculations of Hmax are, at best, determination of displacement pressures. In the reser-
close approximations. Accuracy is limited by uncer- voir, this is the buoyancy pressure required to over-
tainty in sample representativeness and in the deter- come capillary pressure, and thus allow hydrocarbon

Table 2. Seal Facies Geometries.

Facies Thickness (m)* Lateral Extent**


Channel
Abandonment 2–16 *100’s m2–1 km2
Siltstones
Delta-Plain 3–15 1–10 km2
Shale
Delta-Front 3–15 1–10’s km2
Shale
Prodelta Shale <3 1–10’s km2
Shelfal Carbonates 3–15 (Upper TAF) 1–10’s km2
<3 (Lower TAF)

*Data from BZZ area.


**Data from modern-day Mahakam Delta and literature analogs.
Evaluating Seal Potential: Example from the Talang Akar Formation 95

Figure 7. Mercury
injection capillary
pressure curves plot-
ted by facies. Note
reservoir facies in
green.

(the nonwetting phase) to displace water (the wetting 1 psi in the air/mercury system equals 1.15 ft above
phase) from the largest pore throats. In the laboratory, FWL. The displacement pressure of the reservoir rocks
displacement pressure is the pressure at which mer- (Pdr ) was similar for all samples, varying between 6.0
cury (the nonwetting phase) first enters the pore sys- and 6.5 psi air/mercury. The results of MLV seal
tem of the rock. In most reservoir rocks, this is easily capacity measurements for each sample and their cal-
determined as the pressure at which the wetting phase culated range of sensitivities are provided in Figure 9.
saturation first deviates from 100%. However, closure In addition, results of all MLVs plotted by lithofacies
problems are inherent to low-porosity, fracture-prone are shown in Figure 10.
samples typical of seal lithologies. Closure (also
known as conformance) is an artifact of the MICP Seal Capacity Results
experiment wherein mercury tends to fill irregularities
(nicks or fractures) in the sample at low injection pres- In terms of seal capacity (as determined by MICP
sures. Since porosities in most seal rocks are com- measurements), the highest displacement pressures
monly <2%, very small fractures or nicks can (Pd ) are associated with shelf carbonates (Pd 2300–9000
constitute significant proportions of the measured psi mercury). This corresponds to a calculated seal
pore volume. However, these irregularities constitute capacity (Hmax) of 2600–10,000 ft. The next highest Pd
only a small fraction of the sample’s bulk volume, and occurs in the delta-front shales (1000–1400 psi) and
the volume of mercury entering them is not represen- corresponds to Hmax of 1150–1600 psi); this calculates
to H max of 310–2000 ft. The lowest Pd are associated
tative of matrix pore-throat size distribution. Simi-
with delta-plain shales (80–90 psi) and channel aban-
larly, the point at which mercury first enters the
donment siltstones (90–300 psi), which equate to Hmax
irregularities does not represent the actual displace-
of 90–100 ft and 100–350 ft, respectively. These results
ment (entry) pressure of the matrix (true Pds).
are summarized in Table 3.
In this study, potential variations in closure were
accounted for by testing the sensitivities of selected
different displacement pressures. These ranged from DISCUSSION
0.5 to 5% nonwetting-phase (mercury) saturation (Fig-
ure 8). Another method to account for closure is to Combining all the variables affecting seal potential
examine the shape of the logarithmically plotted MICP is difficult even when all of these variables are known.
curve and pick the injection pressure of the first major When many of the intrinsic rock and fluid properties
inflection (FMI) of the curve. If the FMI falls some- are unknown, the results are that much less reliable.
where between the 0.5%–5% saturation values, there is To account for some of these potential inaccuracies, a
greater confidence that pressure represents actual rock range of values, rather than exact values, is presented
matrix Pds , and that point is considered the MLV dis- for the seal potential in the Talang Akar. The sensitivi-
placement pressure. ties discussed account for the possible spectrum of
Allowing for the wide range of variables detailed variation in measured rock and fluid properties in the
above, the hydrocarbon column height (H max) that analysis of seal capacity. Proper sampling and sample
each potential seal lithology could hold was calculated handling are critical issues: do the samples analyzed
on a sample-by-sample basis. Based on equations 1–3, represent the full extent of potential reservoir and
96 Kaldi and Atkinson

Figure 8. Effect of varying


closure (conformance) on
the selection of displace-
ment (energy) pressure.
Note that displacement
pressure varies from <1 psi
at the point of first diver-
gence from 0% nonwetting-
phase saturation (NWPS), to
1.5 psi at 0.5% NWPS, to 6
psi at 5% NWPS. The rec-
ommended point for selec-
tion of displacement
pressure for most samples is
usually between 0.5% to 5%
NWPS or at the first major
inflection (FMI), which in
this example occurs at 4 psi.

potential seal lithologies? What are the effects of the or estimated values of seal capacity, seal geometry, and
removal of overburden stress and of drying out on the seal integrity of each lithology on a scale based on field-
samples? How consistent are hydrocarbon and brine and formation-specific constraints. For instance, the seal
densities throughout the region? capacity was evaluated relative to the actual structure
Within the limits of these caveats, it was possible to in the field (400 ft in the BZZ area). Thus, the rocks hav-
apply a seal potential (SP) ranking of the seal potential ing seal capacity ≥400 ft can hold a column of hydrocar-
of the various fine-grained lithologies in the Talang bon equal to or greater than the structural spillpoint;
Akar of the BZZ area. This semiquantitative ranking these lithologies have an SP for capacity of 1.0. Rocks
system is derived as a product of the individual actual with capacities <400 ft have SPs proportional to their

Figure 9. Range of sensitivities of maximum hydrocarbon column heights (Hmax) supported


by each sampled potential sealing lithology. The key to samples is in Table 1.
Evaluating Seal Potential: Example from the Talang Akar Formation 97

Figure 10. Range of maxi-


mum hydrocarbon column
heights (Hmax) held by each
potential sealing facies
sampled in this study.

measured value (e.g., in this area, rocks able to hold a The SP for area of the potential seal was determined rel-
column of 200 ft would have an SP of 0.50). ative to the area needed to entirely drape closure over
For seal geometry, the depositional characteristics of the specific BZZ structure.
the rocks in question, and the general structural frame- For seal integrity, rock mechanical properties were
work in which they occur, can be fairly accurately qualitatively determined based on petrographic criteria
deduced from core studies, well-logs, and regional and the presence or absence of visible fractures in core
seismic data. The lateral extents and thicknesses of and plugs. It is well understood that the presence of
potential seal lithologies were interpreted from the fractures within any lithology diminishes its seal poten-
examination of a relatively limited number of cores and tial, and that certain rock types are more susceptible to
logs and from comparisons to modern analogs. The fracturing than others (Landes, 1959). In general, salt,
maximum ranking for thickness (SP 1.0) was a 50-ft cut- sulfates, and organic shales have low fracture potential
off. This figure is based on the work of Sluijk and Ned- or heal fractures rapidly under geological conditions. In
erlof (1984) and Sluijk and Parker (1986), who used an contrast, shaly siltstones, tight limestones, and fine-
empirical approach to equate increasing thickness with grained lithic sandstones have a propensity to fracture
decreasing likelihood of through-going fractures. Their at geological strain rates, and fine-grained lithologies
conclusions were that thicknesses >50 ft were generally such as tight dolostones or well-cemented quartz sand-
required for effective seals. They equated this to the stones, due to their very brittle nature, are rarely found
general limits of seismic resolution: fault offsets >50 ft unfractured (Stearns and Friedman, 1972). In this
can be distinguished seismically (and thus used to iden- respect, the less-brittle, organic-rich, silt-free shaly
tify breached seals), while smaller offsets are not visible. lithologies of the TAF, such as the delta-front, prodelta,

Table 3. Seal Capacity Results from MICP.

Displacement Seal Capacity


Facies Pressure (Pd in psi) (Hmax in ft)
Delta-Plain Shale 80–90 90–100
Channel Abandonment Siltstones 90–300 100–350
Prodelta Shale 1000–1400 1150–1600
Delta-Front Shale 1000–1400 1150–1600
Shelf Carbonates
Upper TAF 2300–9000 2600–10,000
Lower TAF 2300–9000 2600–1000
98 Kaldi and Atkinson

Figure 11. Individual component and total seal potential (SP) of all Talang Akar lithofacies. Seal potential is
boldface number at base of each histogram, and is derived as a product of individual components.

and delta-plain shales, have SPs for seal integrity ~1.0. This seal lithofacies is relatively thick, laterally exten-
In contrast, more brittle lithologies, including the shelfal sive, and exhibits measured seal capacities >400 ft
carbonates and channel abandonment silts, have their required to fill the structure to spillpoint. In addition,
ranking for integrity downgraded due to the potential the shales are relatively homogeneous (silt-free) in
for fracturing. The summary of the seal potential rank- thin section and ductile (probably due to their high
ing is displayed in Figure 11. organic content) in core.
On the basis of the criteria presented above and in Although of similar lateral extent and having even
Figure 11, the best sealing potentials within the TAF higher seal capacity than the delta-front shales, the
exist within the more marine, transgressive units that upper shelfal carbonates have a lower sealing poten-
occur preferentially in the upper TAF. This concept is tial (SP .50). This is due to their poor seal integrity
supported by the fact that most hydrocarbon discover- (high potential to fracture). Several long (>10 cm),
ies to date occur in the uppermost TAF. The rocks with open (aperture of >1 cm) fractures were observed in
best seal potential are the delta-front shales (SP .95). at least one shelfal carbonate unit in the BTS-4 well
Evaluating Seal Potential: Example from the Talang Akar Formation 99

Figure 12. Open fractures


result in poor seal potential
of a transgressive shelfal
carbonate unit in the lower
TAF. Circumferential bore-
hole induction log image of
BTS-4 well clearly shows
the SE-dipping open frac-
tures at depths X962 and
X975 ft. This zone tested
1100 BOPD. Note that
zones of higher porosity
(X965–X971) are not frac-
tured.

(Figure 12), indicating that these fractures are open in channel abandonment siltstones (SP .15) also have
the subsurface. low seal potential. In the case of the delta-plain
Prodelta shales have even lower seal potential (SP .18). shales, this is a result of their relatively low mea-
These sediments are of similar lateral extent as the sured seal capacities. In contrast, channel abandon-
delta-front shales and have good seal integrity (low ment siltstones exhibit higher measured seal
fracture potential), but have relatively lower measured capacities, but form poor seals because of their lim-
seal capacities and are thin. This thin geometry increases ited lateral extent. The poorest seal potential in the
the risk of faults breaching the unit, thus juxtaposing TAF occurs within the lower shelfal carbonates (SP .10).
nonsealing lithologies. Delta-plain shales (SP .18) and Like the upper shelf carbonates, these rocks have the
100 Kaldi and Atkinson

risk of fractures, but also have additional risks due to This thin geometry increases the risk of faults breach-
their relatively thin nature. Prodelta shales are of ing the unit, thus juxtaposing nonsealing lithologies.
similar lateral extent as the shelf carbonates and Delta-plain shales (SP. 18) and channel abandonment
delta-front shales, but have relatively lower mea- siltstones (SP .15) also have low seal potential. In the
sured seal capacities and are thin. This thin geome- case of the delta-plain shales, this is a result of their
try increases the risk of subseismic (<50 ft) faults or relatively low measured seal capacities. In contrast,
fractures breaching the unit, thus lowering the seal channel abandonment siltstones exhibit higher mea-
potential. sured seal capacities, but form poor seals because of
Some of the regional exploration implications of their limited lateral extent. The poorest seal potential
these results are that the more marine influenced in the TAF occurs within the lower shelfal carbonates
delta-front shales in the TAF comprise the rocks with (SP .10). Like the upper shelf carbonates, these rocks
the most favorable seal potential. The transgressive have the risk of fractures, but also have additional risk
nature of the TAF succession results in such seals due to their relatively thin nature.
being better developed toward the top of the forma- Because the transgressive nature of the TAF succes-
tion. Unless additional seal lithologies (with high seal sion results in seals being better developed toward the
potential) are encountered in the lower TAF, seals in top of the formation, the rocks with the best seal
this less marine influenced interval are likely to have potential in the TAF are the more marine influenced,
only poor-to-moderate seal potential, therefore limit- upper shelfal carbonates and shales. This concept is
ing the chances of major hydrocarbon discoveries. supported by the fact that most hydrocarbon discover-
Another aspect of the evaluation of seals is the ies to date occur in the uppermost TAF. Unless addi-
potential for reservoirs updip of rocks with poor seal tional seal lithologies (with high seal potential) are
capacity to be charged. In other words, if the hydrocar- encountered in the lower TAF, seals in this less marine
bon column thickness exceeds the seal capacity of a influenced interval are likely to have only poor-to-
lithology, that lithology will be able to transmit hydro- moderate seal potential, therefore limiting the chances
carbons to overlying rocks. If these overlying rocks are of major hydrocarbon discoveries.
reservoir quality and are effectively sealed, accumula-
tions can be expected. A case in point is the delta-plain
shale lithofacies. This rock type has a seal capacity of
ACKNOWLEDGMENTS
<100 ft (Figure 9). Since known hydrocarbon columns We would like to thank a number of individuals for
in the TAF are typically >100 ft, the shale should be oil- their contributions to this study. Bob Sneider (R.M.
saturated (“waste zone” of Downey, 1984) and, of Sneider Exploration Inc.) provided critical fluids data
more significance to explorationists, reservoirs updip and introduced us to the basic premise of seal studies.
of delta-plain shales should be hydrocarbon bearing. Tom Shackleton (ARCO International Oil & Gas Co.)
A method to test for the hydrocarbon saturation of the provided the introduction to the geology of the area.
seal as a test for column exceeding seal capacity is dis- Eric Frodesen, Alan Hart, and Firman Yaman (ARCO
cussed by Noble et al. (1993). Reservoir-quality rocks Indonesia) supported the conduction of the study and
that occur updip from good seals (delta-front shales or facilitated its go-ahead; Chandra Suria (ARCO Indone-
the thicker shelf carbonates) are not as likely to have sia) provided helpful discussions on correlations in the
economic hydrocarbon accumulations. BZZ field. Bambang Wiradi and Simon-Robertson’s
Indonesia helped sample the cores. Discussions with
CONCLUSIONS Chuck Vavra, Mark Scheihing, Phil Lowry, and Jim
Lorsong (ARCO Exploration & Production Technol-
Seal potential in the Talang Akar Formation (TAF) ogy) helped crystallize some of the concepts on seal
can be evaluated in terms of seal capacity capacity and depositional settings. Vavra and Schei-
(MICP–derived membrane seal properties), seal geom- hing’s review of earlier versions of the manuscript
etry (lateral and vertical continuity), and seal integrity improved it considerably.
(rock mechanical properties). The various seals were
ranked using a semiquantitative seal potential (SP) REFERENCES CITED
model. The best sealing potential within the TAF exists
within delta-front shales (SP. 95). This seal lithofacies Allen, G.P., 1987, Deltaic sediments in the modern and
is thick and laterally extensive, and exhibits high mea- ancient Miocene Mahakam Delta: Jakarta, India,
sured seal capacities. In addition, delta-front shales Indonesian Petroleum Association/Total Exploration
appear relatively homogeneous (silt-free) and ductile. Laboratory Field Trip Guide, February 1987, 59 p.
Although of similar lateral extent and relatively high Allen, G.P., D. Laurier, and J. Thouvenin, 1979, Étude
seal capacity, the upper shelfal carbonates have a sedimentologique du delta de la Mahakam: Notes
lower sealing potential (SP .50) because of their et Memoires 15, Paris, Compagnie Française des
propensity to be fractured. Prodelta shales have even Petroles, 156 p.
lower seal potential (SP .18). These sediments are of Core Laboratories, 1982, A course in special core
similar lateral extent as the delta-front shales and have analysis: CoreLab Report, Dallas, Texas, 210 p.
good seal integrity (low fracture potential), but have Downey, M.W., 1984, Evaluating seals for hydrocarbon
relatively lower measured seal capacities and are thin. accumulations: AAPG Bulletin, v. 68, p. 1742–1763.
Evaluating Seal Potential: Example from the Talang Akar Formation 101

Haq, B.U., J. Hardenbol, and P.R. Vail, 1987, Chronol- methods and applications: AAPG Studies in Geol-
ogy of fluctuating sea levels since the Triassic: ogy, v. 21, p. 55–58.
Science, v. 235, p. 1156–1167. Sneider, R.M., 1987, Practical petrophysics for explo-
Landes, K.K., 1959, Petroleum Geology, 2nd ed: New ration and development; AAPG Education Depart-
York, John Wiley, 357 p. ment Short Course Notes.
Murris, R.J., 1980, Middle East: stratigraphic evolution Stearns, D.W., and M. Friedman, 1972, Reservoirs in
and oil habitat: AAPG Bulletin, v. 64, p. 597–618. fractured rocks, in R.E. King, ed., Stratigraphic oil
Noble, R.A., J.G. Kaldi, and C.D. Atkinson, 1997, Oil and gas fields—classification, exploration methods
saturation in shales: applications in seal evaluation, and case histories: AAPG Memoir 10, p. 82–106.
in R.C. Surdam, ed., Seals, traps, and the petroleum Suria, C., 1991, Development strategy in the BZZ
system: AAPG Memoir 67, p. 13–29. field and the importance of detailed depositional
Schowalter, T.T., 1979, Mechanics of secondary hydro- model studies in the reservoir characterization of
carbon migration and entrapment: AAPG Bulletin, Talang Akar channel sandstones: 20th Annual
v. 63, p. 723–776. Indonesian Petroleum Association Convention,
Sluijk, D., and M.H. Nederlof, 1984, Worldwide geo- Jakarta, October 1991, v. 1, p. 419–452.
logical experience as a systematic basis for prospect Vavra, C.L., J.G. Kaldi, and R.M. Sneider, 1992, Geo-
appraisal, in G. Demaison and R.J. Murris, eds., logical applications of capillary pressure: a review:
Petroleum geochemistry and basin evaluation: AAPG Bulletin, v. 76, p. 840–850.
AAPG Memoir 35, p. 15–26. Young, R., G. Juniarto, and B. Thomas, 1991, Widuri
Sluijk, D., and J.R. Parker, 1986, Comparison of Field, offshore southeast Sumatra: sandbody geome-
predrilling predictions with postdrilling outcomes, tries and the reservoir model: 20th Annual Indone-
using Shell’s Prospect Appraisal System, in D.D. sian Petroleum Associaton Convention, Jakarta,
Rice and W.R. James, eds., Oil and gas assessment— October 1991, v. 1, p. 385–418.
Hippler, S.J., 1997, Microstructures and diagenesis in North
Sea fault zones: implications for fault-seal potential and
fault-migration rates, in R.C. Surdam, ed., Seals, traps,
and the petroleum system: AAPG Memoir 67, p. 85–101.

Chapter 7

Microstructures and Diagenesis in North Sea


Fault Zones: Implications for Fault-Seal
Potential and Fault-Migration Rates
S.J. Hippler1
Exxon Production Research Company
Houston, Texas, U.S.A.

ABSTRACT
The deformation mechanisms and diagenetic events that determine the
fluid-flow properties of North Sea faults have been established by characteri-
zation of faults in core and thin section. The results of this study have impli-
cations for predicting fault-seal potential and fault-migration rates.
In the northern North Sea, faulting of weakly lithified sand and shale
units was accomplished via independent particulate flow, in which grains
are disaggregated and reoriented without grain fracture. In the southern
North Sea, faulting of lithified sandstone units was dominated by grain frac-
ture and grain size reduction. Capillary pressure measurements of the
deformed samples indicate that both independent particulate flow and cata-
clastic deformation mechanisms result in relatively low permeability fault
zones in comparison to the surrounding reservoir rocks. Cements are com-
monly observed within the North Sea fault zones and also contribute to the
permeability reduction. Simple calculations using the fluid-flow measure-
ments suggest, however, that faults with good sealing potential can allow
hydrocarbon migration over geologic time scales.

INTRODUCTION relatively small throw faults in sandstone (Pittman,


1981; Antonellini and Aydin, 1994), or large throw
Much progress has been made in identifying the thrust faults in crystalline rocks (Forster et al., 1993). In
deformation mechanisms and diagenetic events that this study, fluid-flow property measurements were
cause faults to seal (Knipe, 1992; Hippler, 1993; obtained from cored faults with a range of throws (mil-
Antonellini and Aydin, 1994). However, while a fault’s limeters to ~1 km) from a well-explored hydrocarbon
fluid-flow properties determine its sealing or migration province. The fluid-flow property measurements are
potential, such data are generally limited to permeabil- used to estimate fault-seal potential and calculate fault-
ity measurements from outcrop exposures of either migration rates. A simple model developed from the

1
Now with Exxon Exploration Company, Houston, Texas

103
104 Hippler

JURASSIC CRETACEOUS P EO O MIO


0
Cormorant Field Brent
Rifting Event
Field Norway
2,000

Burial Depth (Feet)


Northern North Sea
4,000

6,000

8,000

200 180 160 140 120 100 80 60 40 20 0


Ma

Figure 2. Schematic burial-history plot of the Brent


Southern North Sea Group reservoir in the North Sea. The extensional
faults in the Brent and Cormorant fields were active
North
Valiant during a Late Jurassic rifting event, and are cross-cut
Field at the crest of the structures by the Base Cretaceous
erosional event. The burial history indicates that the
Great Brent Group was buried to depths of ≤2000 ft (≤500 m)
Britain at the time of faulting.
Amsterdam
London
100 miles
pressure tests (Vavra et al., 1992), from which fault-
seal potential and fault-migration rates were calcu-
Figure 1. Location map of the North Sea hydrocar- lated.
bon fields that contain the faulted cores discussed in
this paper. Northern North Sea Examples
Cores from the Brent and Cormorant fields contain
faults formed during a Jurassic rifting event. Large
results of the calculations explains one mechanism by normal faults of the same age set up the tilted normal
which faults may allow hydrocarbon migration over fault blocks that are characteristic of the hydrocarbon
geologic time scales while also sealing significant traps in the northern North Sea. The cored faults inter-
hydrocarbon columns. sect the sand-dominated deltaic sequence of the Brent
Group, the major reservoir in this area, which was
deposited during the rifting event. The crests of the
FAULT ZONE CHARACTERIZATION tilted fault blocks were eroded by the Base Cretaceous
unconformity, which marks the end of the major rift-
Samples and Analyses
ing event. Burial-history analysis and seismic data
The fault-zone cores studied are from hydrocarbon indicate that the faults were active when the reservoir
fields in the northern North Sea and the southern was buried to depths ≤2000 ft (≤500 m) (Figure 2).
North Sea, and contain numerous minor faults (with A deviated well at the crest of the Brent Field inter-
throws <10 m) and several major faults (with throws sects listric faults that are imaged on seismic data
≥10 m) that affect clastic facies sandstones and shales. (Figure 3) and other minor fault zones that are below
Three examples that are representative of the range of seismic resolution. Small throw (≤1 m) shale and sand
faults observed in the suite of cores are discussed in fault zones are common throughout the cored interval
detail in this paper (Figure 1). where thin shale beds are present in the Brent Group
The fault zones were characterized by micro- (Figure 4). The fault zones are characterized by the
structural analysis from optical transmission elec- intercalation of sand and shale clasts that are smeared
tron (Barber, 1985), scanning electron (Lloyd, 1985), into parallelism with the fault planes. Optical
and cathodoluminescence (Marshall, 1988) micro- microscopy indicates that the sand clasts are
scopy. The microstructural observations were used deformed via independent particulate flow, because
to determine the deformation mechanisms and dia- fractured grains and grain size reduction are not
genetic processes that contribute to each fault’s observed (Figure 5).
fluid-flow properties. The fluid-flow properties Fault-zone samples from the Cormorant Field are
were measured with mercury intrusion capillary from a deviated well that intersects the crest of Block
Microstructures and Diagenesis in North Sea Fault Zones 105

Brent Field Figure 3. Schematic cross


section based on an east-
west seismic line through
the Brent Field, North Sea,
W E
showing location of the
WELL WITH CORED FAULTS well with the cored faults.
y
nformit
eous Unco
Base Cretac
roup
rent G
Top B
rou p
un lin G
Top D

1 Km

IV (Figure 6), one of the field’s four discrete oil accu-


mulations in the Brent Group. Syndepositional faults
within Block IV have further segmented the reser-
voir and have created at least four compartments
with different original oil-water contacts. A west-
dipping antithetic fault with ~60 m of throw at the
core location juxtaposes Dunlin Group shale in the
footwall with the Brent Group reservoir in the hang-
ing wall. The cored fault separates two subblocks
with different original oil-water contacts.
Excellent core recovery indicates that fault-zone
deformation is localized to ~2 m on either side of the
fault plane. The Dunlin Group contains a penetrative
scaly fabric, a common deformation feature in fine-
grained sediments (Moore et al., 1986). Deformation
in the Brent Group is characterized by lozenges of
sediment bounded by pyrite-filled shear planes ori-
ented subparallel to the fault plane (Figure 7). A thin
lozenge of the coarse-grained Broom Formation of
the Brent Group is preserved directly adjacent to the
fault plane (Figure 7). Directly above, a shear zone
(several grain diameters in width) juxtaposes the
finer grained Rannoch Formation with the Broom
Formation.
Optical microscopy of the coarse-grained Broom
Formation directly adjacent to the fault zone reveals a
limited number of angular quartz grains, suggesting
minor grain breakage and grain size reduction (Fig-
ure 8). However, the extent of the cataclastic defor-
mation is limited. For example, at the Broom/
Rannoch shear-zone boundary, there are several
large quartz grains from the Broom Formation inter-
mixed with the finer grained Rannoch Formation. Figure 4. Core sample of a minor fault in the Brent
The dominance of unfractured grains in the fault Field (viewed approximately perpendicular to the
zone suggests that slip has taken place via grain dis- fault plane). Undeformed Brent Group sandstone is
aggregation and independent particulate flow. The on the left, bounded by a narrow, shale-rich fault
particulate flow has resulted in the repacking of zone on the right. Note the elongated lenses of sand
grains and, in combination with pyrite cement, has intermixed with the shale.
significantly reduced the pore space and size in the
reservoir sediments in the fault zone. which the Rotliegende Group, the primary reservoir,
was laid unconformably on eroded Carboniferous sedi-
Southern North Sea Example
ments. The Rotliegende Group is a primarily aeolian
Fields in the southern North Sea are characterized sandstone deposited in a continental desert environ-
by a northwest-southeast structural trend that was ini- ment. During early Cretaceous and mid-Tertiary inver-
tially set up by a Permian-Triassic rifting event, during sion events, many of the Permian-Triassic extensional
106 Hippler

Figure 5. Optical photomontage through the fault zone shown in Figure 4. Undeformed Brent sandstone is
on the left, the sand/shale fault zone is on the right. The sand-rich portion of the fault zone is one of the
lenses of sandstone seen in Figure 4. Note that the shale fabric has been smeared into parallelism with the
fault plane. Also note the loss of porosity from left to right, and the absence of grain size reduction in the
sandstone.

faults were reactivated in a reverse or strike-slip sense reverse and normal offsets ≥10 cm. The fault zones are
(Farmer and Hillier, 1992). Reservoir quality is gener- altered to pale green or white and thus stand out in com-
ally poor, due to the compaction and diagenesis that parison to the red-stained reservoir rock (Figure 10).
accompanied deep burial of the sediments prior to the Optical microscopy indicates that the fault zones
early Cretaceous inversion event (Figure 9). are characterized by grain size reduction and an
The North Valiant Field contains multiple faults that increase in matrix material in comparison to the unde-
separate the field into two compartments with differing formed rock (Figure 11). High-resolution scanning
gas-water contacts. Cores from the reservoir rock inter- electron microscopy indicates that the matrix material
sect numerous minor fault zones, indicating both consists of illite and fine-grained quartz (Figure 12a).

Figure 6. Schematic cross


Cormorant Block IV
section based on an east-
west seismic line through
Cormorant Block IV, North
W E Sea, showing the location of
WELL WITH CORED FAULT
the well with the cored
fault.
it y
form
c on
Base Cretace ous Un
up
T o p B r en t G r o Bren
t

Top Dunlin Gro up

1 Km
Microstructures and Diagenesis in North Sea Fault Zones 107

hydrocarbon column heights that can be held by capil-


lary top seals (Vavra et al., 1992). This work has shown
that capillary pressure measurements on a few repre-
sentative samples may give a good approximation of
seal capacity in stratigraphic units with great lateral
continuity and vertical homogeneity.
In contrast, field observations indicate that fault
zones are often composed of a complex and heteroge-
neous distribution of deformed rocks, the origins of
which are often difficult to determine. This suggests
not only that fluid-flow properties of fault zones are
potentially extremely variable, but that to begin pre-
dicting this variability, the fault-zone properties must
first be placed in the context of the fluid-flow proper-
ties of the rocks from which they formed. Thus, the
approach taken in this study was to define the contrast
in fluid-flow properties between the deformed sam-
ples in the fault zone and their undeformed counter-
parts outside the fault zone with capillary pressure
testing on a range of faulted lithologies.
Compilation of the capillary pressure test data from
all of the North Sea cores (Figure 13) indicates that the
capillary entry pressures of deformed samples from
the fault zone are generally higher than the capillary
entry pressures of undeformed counterparts outside
the fault zone. This is consistent with fault-zone char-
acterization observations that indicate that the pore
size and space in North Sea fault zones are reduced
through disaggregation and repacking of grains, grain
size reduction, and cementation. This is particularly
noticeable in the reservoir-quality rocks, as their
mercury-air capillary entry pressures are increased by
more than an order of magnitude, to the 200–300 psi
range (Figure 13). A capillary pressure seal of this
quality can support a 30–55-m oil column or a 15–30-m
Figure 7. Core sample of a major fault from
gas column, depending on hydrocarbon density
Cormorant Block IV, North Sea (viewed approxi-
(Schowalter, 1979).
mately perpendicular to the fault plane). The fault
plane is indicated; the footwall shales lie below
and out of picture. In the hanging wall, a sliver of ESTIMATING PERMEABILITY FROM
coarse-grained Broom Formation is smeared direct-
ly adjacent to the fault plane and is juxtaposed CAPILLARY PRESSURE
against the fine-grained Rannoch Formation MEASUREMENTS
(Figure 8). The Rannoch Formation contains
numerous black, anastomosing shear planes orient- Thompson et al. (1987) discuss in detail how the
ed subparallel to the fault plane. The black shear dynamic transport properties of porous rocks, such as
planes are filled with pyrite. absolute permeability, can be approximated from the
results of static mercury injection capillary pressure
measurements by applying percolation theory. For
example, by using the following relationship, we can
Cathodoluminescence, in combination with the scan- calculate the absolute permeability (K) reduction that
ning electron microscopy (SEM), indicates that even the corresponds to the capillary pressure increase in the
large quartz grains have been fractured and cemented fault zones (Thompson et al., 1987):
by quartz (Figure 12b). The grain fracture and grain size
reduction indicate that cataclasis was the dominant
deformation mechanism during faulting. ( )
K = ( c) lc 2 (σ / σ o )
(1)

CAPILLARY PRESSURE where K = permeability (darcies); c = a constant that


MEASUREMENTS connects the conductivity of the capillary structure of
equivalent size lc to the permeability of the same path
Mercury injection capillary pressure measure- [the work of Thompson et al. (1987) indicates that c ≈
ments are now used almost routinely to quantify the 1/226]; l c = the capillary diameter (in micrometers)
108 Hippler

Figure 8. Optical photo-


montage through the fault
zone shown in Figure 7. The
coarse-grained Broom
Formation lies directly
adjacent to the fault plane
(bottom) and is smeared
against the fine-grained
Rannoch Formation (above).
The black cement is pyrite.
Note only minor grain
fracturing and lack of
grain size reduction in the
fault zone.

associated with the first connected path in a mercury MIGRATION-RATE CALCULATION


intrusion test (= the pore throat diameter, d, deter-
mined by the Washburn equation); σ = electrical con- The fluid-flow property measurements can pro-
ductivity of the rock saturated in brine solution; and vide a general understanding of fault-migration
σo = brine conductivity [(σ/σo) can be approximated rates if the fault-zone properties are assumed to be
by φ2, where φ is porosity]. homogeneous and isotropic; the following equa-
Figure 14 shows that while the calculated perme- tion is used to estimate the average velocity (V) of
abilities of undeformed silts and silty shales were an oil layer behaving as a thin film in the fault
decreased only slightly as a result of the faulting, the zone:
permeabilities of some undeformed reservoir-quality
rocks were reduced by three orders of magnitude, in
some cases to the 10–4–10–5 d range. V = K / µ × g × ∆ρ × sin (Θ ) (2)
Microstructures and Diagenesis in North Sea Fault Zones 109

CAR PERM TRIAS JURASSIC CRET P EO O MIO


0

2,000
Burial Depth (Feet)

4,000 Inversion Events


Rifting Event

6,000

8,000

10,000

300 250 200 150 100 50 0


Ma

Figure 9. Schematic burial-history plot of the


Rotliegende Group, southern North Sea. The exten-
sional faults observed in the seismic data from the Figure 10. Core sample of a minor fault in the
southern North Sea were active in an extensional Rotliegende Group, North Valiant Field, southern
sense during the Permian rifting event, and reacti- North Sea (viewed approximately perpendicular to
vated in a reverse or strike-slip sense during inver- the fault plane). The undeformed reservoir is stained
sion events in the Late Cretaceous and mid-Tertiary. red due to iron-oxide grain coatings. Bedding is
The burial history indicates that the Rotliegende subhorizontal, and the fault zone (pale green/white)
Group was buried to depths of ~6000–8000 ft (~2–3 trends from the top left to the bottom right.
km) at the time of the fault reactivation.

where V = average oil velocity (m/sec); K = k/φ, the (in degrees); and ∆ρ = density contrast of water and oil
ratio of effective permeability to porosity; K = perme- (in grams per cubic centimeter).
ability (in square meters); φ = porosity; µ = viscosity (in This calculation considers only the movement of
kilograms per meter per second); g = gravitational a single drop of oil along the fault pathway, and
acceleration (9.8 m/sec2); Θ = average dip of the fault therefore the calculated rates are independent of

Figure 11. Optical micrograph through


the fault zone shown in Figure 10.
The fault zone (trending from top
left to bottom right) is dominated by
fine-grained matrix material (brown),
while the undeformed reservoir rock
(top right) has visible porosity and
limited grain fracture.
110 Hippler

Figure 12. (a) Back-


scattered, scanning
electron micrograph
(SEM) of the fault zone
shown in Figure 11.
Quartz grains (gray)
appear relatively unde-
formed. (b) Combined
back-scattered, scan-
ning electron and
cathodoluminescence
micrograph of same
area of fault zone
shown in (a).
Cathodoluminescence
reveals that the quartz
grains are extensively
fractured and contain
quartz cement (slightly
darker gray material
between lighter gray
quartz clasts).

fault thickness (width) and strike length. The calcu- caused by the density contrast between the oil and
lation assumes: (1) that oil travels as a separate the water; and (3) Darcy’s Law applies.
phase in a thin layer within the fault zone; (2) the Figure 15 shows the rate of hydrocarbon migration as
resistance to flow is equal to the buoyancy force a function of fault-zone permeability for oil and gas
Microstructures and Diagenesis in North Sea Fault Zones 111

Figure 13. Comparison of


capillary entry pressures of
undeformed and deformed
samples. Each data point
represents the results of the
capillary pressure measure-
ments on a set of paired
samples, one undeformed
and one deformed. The
heavy line on the graph
represents a one-to-one
relationship of the capillary
entry pressure between
undeformed and deformed
samples; the thin line
indicates one order of
magnitude increase in
capillary entry pressure
from undeformed to
deformed samples. For
consistency, entry pressure
was picked at 5% mercury
saturation in each sample.

migrating through a fault dipping 70° API. For a fault- DISCUSSION


zone permeability in the 10–4 to 10–5 d range, similar to
that observed in this study, oil migrates at a rate of 1 Although limited, the fluid-flow property measure-
km/100,000–1 m.y., while gas migrates at a rate of ments suggest that the deformation mechanisms that
1 km/1000–10,000 yr. In other words, the calculation sug- operate in a sand-shale sequence can result in fault-zone
gests that fault-migration rates are geologically instanta- material with an intermediate mercury-air capillary
neous for oil when fault-zone permeability is >10–5 d, entry pressure (200–300 psi), and a correspondingly low
and for gas when fault zone permeability is >10–7 d. permeability (10–4 to 10–5 d) relative to an undeformed

Figure 14. Comparison


of calculated absolute
permeabilities between
undeformed and deformed
samples. The permeabilities
were calculated from the
capillary entry pressure data
shown in Figure 13. The
heavy line on the graph
represents a one-to-one
relationship between the
undeformed and deformed
samples; the thin lines
represent one, two, and
three orders of magnitude
decrease in permeability
from undeformed to
deformed samples.
112 Hippler

Figure 15. Graph showing


the rate of hydrocarbon
migration (years per
kilometer) as a function of
fault-zone permeability if φ
= 0.10; µ = 0.003 kg/m/sec
for oil and 0.001 kg/m/sec
for gas; g = 9.8 m/sec2; Θ =
70° API; and ∆ρ = 100 kg/m3
for 26° API oil and 1000
kg/m3 for 45° API gas.

Figure 16. A simple conceptual model of fault migration in an interbedded sand-shale sequence. Three gently
dipping beds are reservoir-quality rocks connected via a fault zone (indicated). The reservoirs and fault zone
are surrounded by shale with high seal capacity. The capillary entry pressure of the fault zone is intermediate
between the reservoir and the shale. The gray scale is a schematic representation of the location of the hydro-
carbons in the hypothetical system. A = Hydrocarbons enter the fault and reservoir system via the lower reser-
voirs and the hydrocarbons flow up toward the fault due to buoyancy. B = Hydrocarbon column accumulates
at the reservoir–fault interface. C = Hydrocarbons enter the fault zone when the buoyant force of the growing
hydrocarbon column exceeds the capillary entry pressure of the fault zone, and migrate upward.
Microstructures and Diagenesis in North Sea Fault Zones 113

reservoir rock. The simple calculations suggest that REFERENCES CITED


faults with such fluid-flow properties can allow
hydrocarbon migration over geologic time scales, Antonellini, M., and A. Aydin, 1994, Effect of faulting
while simultaneously sealing significant hydrocarbon on fluid flow in porous sandstones: petrophysical
columns. properties: AAPG Bulletin, v. 78, p. 355–377.
These results can be used to develop a simple Barber, D.J., 1985, Dislocations and microstructures, in
model that explains one mechanism by which hydro- H.-R. Wenk, ed., Preferred orientation in deformed
carbon migration can take place in faulted, interbed- metals and rocks: an introduction to modern tex-
ded sand-shale sequence (Figure 16). If hydrocarbons ture analysis: London, Academic Press, p. 149–182.
enter a hypothetical fault and reservoir system via Farmer, R.T., and A.P. Hillier, 1992, The Barque Field
the lower reservoirs, the hydrocarbons flow up Blocks 48/13a, 48/14, UK North Sea, in I.L. Abbotts,
toward the fault due to buoyancy. When hydrocar- ed., United Kingdom oil and gas fields, 25 years
bon columns accumulate at the reservoir–fault inter- commemorative volume: U.S. Geological Society
face, the buoyant force of the growing hydrocarbon Memoir 14, p. 395–400.
column may ultimately exceed the capillary entry Forster, C.B., J.V. Goddard, and J.P. Evans, 1993, Per-
pressure of the fault zone, and hydrocarbons begin meability structure of a thrust fault, in S. Hickman,
to enter the fault and migrate upward at the rate R. Sibson, and R. Bruhn, eds., U.S. Geological Soci-
corresponding to the permeability of the fault-zone ety Proceedings of Workshop LXIII, The mechanical
material. In this manner, the fault acts as both a trap involvement of fluids in faulting: U.S. Geological
seal and a migration pathway to shallower reser- Society Open-File Report 94–228, p. 216–223.
voirs. This model assumes that there is immediate Hippler, S.J., 1993, Deformation microstructures and
imbibition of the wetting phase (water) back into the diagenesis in sandstone adjacent to an extensional
fault zone once the buoyancy of the hydrocarbons fault: implications for the flow and entrapment of
drops below the initial capillary entry pressure of hydrocarbons: AAPG Bulletin, v. 77, p. 625–637.
the rocks. Knipe, R.J., 1992, Faulting processes and fault seal, in
R.M. Larsen, H. Brekke, B.T. Larsen, and E. Taller-
aas, eds., Structural and tectonic modelling and its
CONCLUSIONS application to petroleum geology: Norwegian
Petroleum Society Special Publication 1, p. 325–342.
The fault-migration calculations presented in this Lloyd, G.E., 1985, Review of instrumentation, tech-
paper assume a simple fault-permeability structure in niques, and applications of SEM in mineralogy, in
which fluids flow through homogeneous, isotropic, J.C. White, ed., Applications of electron microscopy
porous media. Characterization of fault zones in core in the earth sciences: Mineralogical Society of
and outcrop indicates that fault-zone permeability Canada Short Course, v. 11, p. 151–188.
structures are commonly more complex and are likely Marshall, D.J., 1988, Cathodoluminescence of geologic
to be influenced by the presence of a fault-related frac- materials: London, Unwin Hyman, 146 p.
ture network. This study represents only a first step Moore, J.C., S. Roeske, N. Lundberg, J. Schoonmaker,
toward understanding fluid flow in more complicated D.S. Cowan, E. Gonzales, and S.E. Lucas, 1986,
fault-permeability structures. Scaly fabrics from Deep Sea Drilling Project cores
In addition, the calculations assume that fault- from forearcs, in J.C. Moore, ed., Structural fabrics
zone permeability is static. Current research is in Deep Sea Drilling Project cores from forearcs:
focused on determining the relative timing of fault- Geological Society of America Memoir 166,
ing and cementation to better understand whether p. 55–73.
the cements precipitate geologically instantaneously Pittman, E.D., 1981, Effect of fault-related granulation
as a result of focused fluid flow during fault activity, on porosity and permeability of quartz sandstones,
or over longer geologic time scales as a result of Simpson Group (Ordovician), Oklahoma: AAPG
posttectonic fluid flow. Bulletin, v. 65, p. 2381–2387.
Schowalter, T.T., 1979, Mechanics of secondary hydro-
carbon migration and entrapment: AAPG Bulletin,
ACKNOWLEDGMENTS v. 63, p. 723–760.
Thompson, A.H., A.J. Katz, and C.E. Krohn, 1987,
I am indebted to several collegues at Exxon Produc- The microgeometry and transport properties of
tion Research Company: A.E. Bence and K.E. Green sedimentary rock: Advances in Physics, v. 36,
for providing the migration-rate calculation, and G.G. p. 625–694.
Gray for a thorough review. O. Poix is also thanked for Vavra, C.L., J.G. Kaldi, and R.M. Sneider, 1992, Geo-
a constructive review that much improved the clarity logical applications of capillary pressure: a review:
of this paper. AAPG Bulletin, v. 76, p. 840–850.
Leith, T.L., and A.E. Fallick, 1997, Organic geochemistry
of cap-rock hydrocarbons, Snorre Field, Norwegian
North Sea, in R.C. Surdam, ed., Seals, traps, and the
petroleum system: AAPG Memoir 67, p. 115–134.

Chapter 8

Organic Geochemistry of Cap-Rock


Hydrocarbons, Snorre Field, Norwegian
North Sea
T.L. Leith
IKU Petroleum Research
Trondheim, Norway

A.E. Fallick
Scottish Universities Research and Reactor Centre
East Kilbride, Glasgow, Scotland

ABSTRACT
This study investigates the organic geochemistry of the cap-rock succes-
sion overlying the Triassic–early Jurassic sandstone reservoir of the Snorre
Field in the Norwegian sector of the North Sea. Understanding the relation-
ships between hydrocarbon-bearing reservoirs and their cap rocks is of fun-
damental importance to prospect generation and reservoir studies.
The results confirm that vertical leakage of oil from the reservoir of the
Snorre Field into overlying claystones and marls of the Late Cretaceous
Shetland Group has occurred. The relatively high concentrations of residual
oil present in the cap rock are compositionally related to the reservoired
Snorre oil, but are enriched in asphaltenes and polar compounds. The con-
centration of the residual oil gradually decreases toward the top of the cap-
rock unit. Molecular biomarker data allowed monitoring of mixing between
leaked residual oil and traces of immature indigenous bitumen in the gener-
ally organic-lean cap rock.
Residual oil in the cap rocks overlying the Snorre Field must have been
emplaced directly from the reservoir by bulk-flow processes. Although the
occurrence of a fracture zone in the cap rock may lead to locally high resid-
ual oil concentrations, there is no evidence that major fracture systems are
responsible for emplacing the oil found in the cap rock. The sealing capacity
of the cap rock is therefore suggested as being related to a combination of
lithological variation in the cap rock, microfracturing, and hydraulic equilib-
rium with the reservoir. The occurrence of calcareous cements with a partial-
ly organic carbon isotope signature suggests that the transit of oil through a
cap rock succession may enhance the sealing capacity of that cap rock under
certain circumstances.

115
116 Leith and Fallick

(a) Figure 1. (a) Summary of


published data on leakage
No Report in reservoir cap rocks in
46%
Yes
some Norwegian oil fields.
37% (b) Summary of published
data on reservoir cap-rock
leakage on the Norwegian
Shelf organized by play
type. [All data primarily
summarized from
references in Spencer et al.
(1987).]

Possible
17%

(b)

100%

80%

60% No report
Possible
Yes
40%

20%

0%
K-T Chalk Paleogene L. Cret. U. Jur. L.-M. Jur.
Sst. Sst. Sst. Sst.
Play

INTRODUCTION accumulations known to date, and are usually consid-


ered a prerequisite in evaluating a prospect. However,
Evidence for the leakage of hydrocarbons is a com- in spite of their importance to reservoir systems, cap
mon feature of many oil fields in the Norwegian sector rocks are usually only rather cursorily evaluated dur-
of the North Sea (Figure 1). This evidence may take the ing an exploration or appraisal program and are rarely
form of seismic “bright spots” where gas has accumu- cored. In this paper, “cap rock” will be applied as
lated in a porous horizon, such as Albuskjell (D’Heur, referring to the physical position of a rock succession
1987), or a seismic “gas chimney” where the succession in relation to a reservoir horizon. Consequently, “seal”
above the reservoir contains gas, such as Ekofisk (Pekot will be used to refer to the ability of a horizon to
and Gersib, 1987); Valhall (Leonard and Munns, 1987); impede the passage of hydrocarbons. Therefore, a
and Hild (Rønning et al., 1987). In addition, oil shows fine-grained horizon overlying a reservoir can be a cap
have been reported from strata of Cretaceous and Ter- rock without being a seal to that reservoir.
tiary age overlying the Valhall (Leonard and Munns, The capacity of a cap rock to seal a hydrocarbon
1987), Gullfaks (Irwin, 1989), and Oseberg fields (Dahl accumulation is determined by the capillary pressure
and Speers, 1985, 1986; Nipen, 1987). More recently, in the cap rock (Berg, 1975; Schowalter, 1979; Downey,
Leith et al. (1993) described significant concentrations of 1984; Davis, 1987). This capillary pressure is deter-
“wet” cuttings gas and oil shows from Cretaceous and mined by the radius of the largest interconnected pore
Tertiary strata overlying the Snorre Field (Figure 2). throats in the cap rock, the interfacial tension between
A “cap rock” or “seal” may be defined as a “layered the trapped/migrating oil/gas and pore water, and the
geologic unit capable of impeding hydrocarbon move- wettability of the rock in question. The role of wetta-
ment” (Downey, 1984). Such geologic units are integral bility in hydrocarbon reservoir systems is poorly
parts of the vast majority of commercial hydrocarbon understood in relation to the interactions between
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 117

Figure 2. Map showing


location of Snorre Field and
wells included in this study.

A 2
34/4
4 9
Snorre
1

BERGEN A1

NORWAY

STAVANGER
Vigdis

Inner Snorre Fault


ABERDEEN

Tordis

34/7

hydrocarbons and rock. The presence of significant dichloromethane containing 7% (v/v) methanol as a sol-
amounts of oil-like hydrocarbons in cap-rock succes- vent for ~18 h. After extraction, the solvent was evapo-
sions above economically important oil fields suggests rated by using a Büchi Rotavapor. The asphaltene
that the relationship between a reservoir and its seal fraction of the extractable organic matter (EOM) was
may be more dynamic than previously thought precipitated out using excess n-pentane. The remaining
(Roberts, 1980; Welte et al., 1984). maltenes were separated into saturated, aromatic, and
nonhydrocarbon fractions using a standard medium
ANALYTICAL PROCEDURES pressure liquid chromatography (MPLC) system (after
Radke et al., 1980), with n-hexane as an eluant.
The cuttings samples were washed and reverse- The saturated hydrocarbon fraction was further ana-
picked for contaminants. The pulverized samples were lyzed by using a Hewlett-Packard 5730A gas chromato-
extracted using Soxhlet apparatus with boiling graph with a 15 m DB-1 capillary column. Hydrogen

Table 1. Stable Isotope Data from Carbonate Cements at the Base and Top of the
Shetland Group: Well 34/7-1.*

Depth δ13C δ18O δ18O


(m) (‰ PDB) (‰ PDB) (‰ SMOW)
1840 –4.96 –5.46 25.29
–4.95 –5.50 25.24
2381 –3.54 –5.40 25.34
–3.82 –5.80 24.93

* Including duplicate analyses.


118 Leith and Fallick

(a) Figure 3. (a) Lithostratigraphy of the


Snorre Field area (Hollander, 1987).

Formation
(b) Schematic geologic cross section

System

Group
through the Snorre Field showing
the relative positions and structural
context of the study wells.
0 Legend

Claystone

Shale
Nordland

Sandstone

Marl

Volcanics
Tertiary

Utsira
1000
Hordaland
Depth (m)

W E
Jorsal-
fare
Rogaland

Balder
Lista /
Sele
Kyrre
Cretaceous

2000
Shetland

Trygg-
vason

Blodøks
Cromer Knoll
Dunlin
Jurassic
Svarte

Statfjord
Triassic

Lunde
Lunde

3000

was used as a carrier gas, with a flow rate of ~1.5 mL/min. Hewlett-Packard 5890 Series II gas chromatograph
Injections were performed in splitless mode. A temper- was fitted with a fused-silica DB-5 capillary column
ature program of 80°–280°C at 4°C/min. was used. with helium as a carrier gas. Injections were carried
Gas chromatography–mass spectrometry (GC-MS) out in splitless mode. The temperature program used
was performed on a VG TRIO 1 GC-MS system. A went from 120° to 280°C at a rate of 4°C/min. The
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 119

(b)
34/4-2 34/7-4 34/7-9 34/7-1

Nordland Group

Inner Snorre Formation


Hordaland Group

Rogaland Group
Jorsalfare Formation
?
Shetland Group ?
Kyrre Formation
? ?
?
on
Statfjord Formati Hegre Group Approximate
OWC

Figure 3. Continued.

saturated hydrocarbon fraction was analyzed in multi- Upper Jurassic rocks over the Snorre structure (Fig-
ple ion mode (MID) with a scan cycle time ~2 s. The ure 3). The present geometry of the area is related to
mass spectrometer operated at 70 eV electron energy Cretaceous differential subsidence but does not cor-
with an ion source temperature of 200°C. respond to pre-Cretaceous relief (Karlsson, 1986).
Total organic carbon (TOC) determinations were Karlsson (1986) also suggests that faults developed
carried out by using a Leco CS 244 carbon/sulfur ana- during the Middle Jurassic–Early Cretaceous tec-
lyzer. Rock-Eval pyrolysis measurements were per- tonic phase may have had surface expression, at
formed on whole-rock samples by using a Rock-Eval II least during Early Cretaceous times. Caillet (1993)
instrument at standard conditions. showed that a number of faults penetrate Creta-
Samples for stable isotopic analysis were plasma- ceous/Tertiary strata, particularly around well
ashed to remove organic contaminants (Hogg et al., 34/7-1.
1993). Carbon dioxide was released by reaction with The Triassic–early Jurassic reservoir rocks of the
phosphoric acid for 3 hr at 25°C (McCrea, 1950; Wachter Snorre Field (Figure 3) are usually overlain by a con-
and Hayes, 1985) and the carbon and oxygen isotopic densed succession that rarely exceeds 20 m in thick-
compositions determined on a VG SIRA 10 mass spec- ness, representing the Lower Cretaceous Cromer
trometer calibrated against international reference mate- Knoll Group. In the Snorre area, this unit is organic
rials [i.e., precision is ± 0.1‰ (1σ) or better and NBS19 lean (TOC ~0.3%) and grades upward from a calcite-
gives δ13C = 1.92‰ and δ18O = –2.19‰, both relative to cemented sandstone or siltstone lithology at the base
PDB]. A carbon dioxide–carbonate oxygen isotope frac- to a marl at the top of the unit.
tionation factor of 1.01025 at 25°C was assumed (Rosen- The rocks of the Cromer Knoll Group are uncon-
baum and Sheppard, 1986). The data are given in Table formably overlain by the predominantly argillaceous
1, where the δ18O values are listed for both the PDB and succession of the Shetland Group, which varies in
SMOW scales. thickness from 400 to 700 m over the Snorre Field. Of
the five formations recognized by Hole et al. (1989)
and Nybakken and Bäckstrøm (1989), only two, the
GEOLOGY OF THE CAP-ROCK Kyrre and Jorsalfare, are presently recognized over the
SUCCESSION Snorre Field (Figure 3). In addition, the upper half of
the Kyrre Formation appears to be represented by only
Erosion associated with the Cimmerian tectonic the upper half of the formation over the Snorre Field
phase has led to the complete removal of Middle and (34/7-1 and 34/7-4).
120 Leith and Fallick

(a) (b)
22 1000
Type I
20
Hordaland (34/7-4)
18
800

Rock-Eval Hydrogen Index (mg/gTOC)


Hordaland (34/7-1)
16
Rogaland (34/7-4)
Number of Samples

14 Rogaland (34/7-1) Type II


600
12 Jorsalfare (34/7-4)

Jorsalfare (34/7-1)
10
Kyrre (34/7-4) 400
8
Kyrre (34/7-1)
6

4 200 Type III

0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 400 420 440 460 480 500
Total Organic Carbon (%) Rock-Eval Tmax (°C)

Figure 4. Summary of organic richness, maturity, and kerogen type from rocks of the Shetland Group over the
Snorre Field.

The claystones of the Kyrre Formation are typically the field and the postfilling equilibration processes that
more organic rich (0.7–1.0%) than those of the more affect the oil (England et al., 1987; Karlsen and Larter,
calcareous Jorsalfare Formation (0.5–0.7%). In addi- 1989; Horstad et al., 1990; Larter et al., 1990). The reser-
tion, the rocks of the Shetland Group show a slight voir horizons in the Snorre Field are lithologically hetero-
westward enrichment in organic carbon content geneous and typically consist of oil-bearing sandstone
(0.5–0.8%). Rock-Eval Tmax data suggest that thermally horizons, which are typically only a few meters thick.
immature, type III/IV kerogen is typical of all the These horizons are separated by mudstones, siltstones,
rocks analyzed from the Shetland Group (Figure 4). and fine-grained, well-cemented sandstones.
The rocks of the Shetland Group are further overlain The lithological heterogeneity of the Snorre reservoir
by ~1800 m of largely argillaceous strata representing is reflected in the composition of the extractable organic
the Rogaland (Lista/Sele and Balder formations) and matter (EOM) isolated from the rocks. Regardless of
Hordaland groups. Total organic carbon contents in the reservoir age, the highest EOM concentrations occur in
Hordaland Group show an upward increase from 0.3% coarse-grained, poorly cemented sandstones. This
to 0.5% to 1.0% to 1.7% in 34/7-1 and 1.0% to 2.0% in EOM, from wells 34/7-1 and 34/7-4 (Figures 5, 6), is
34/7-4. Thin limestone or dolomite beds may be recog- usually rich in hydrocarbons (>80%) with a high satu-
nized at the base of the Lista/Sele Formation, particu- rate/aromatic ratio (2–3) and a low asphaltene content
larly in well 34/7-1. In well 34/7-4, the lower 60 m of the (<5%), similar to the drill-stem test (DST) oils. Gas chro-
Hordaland Group is represented by interbedded clay- matography data from these extracts are characterized
stones and fine to medium-grained sandstones (Figure by C15+ n-alkane compositions, which are almost identi-
3). The Hordaland Group is overlain by mid-Miocene to cal to those of the DST oil; that is, a smooth, unimodal
Recent claystones and sand of the Nordland Group. distribution with a steady decrease in peak height from
a maximum around C15 or C16 (Figure 7).
In contrast, finer grained reservoir lithologies tend to
DISCUSSION OF RESULTS contain much lower concentrations of more asphaltene-
Composition of Reservoir Hydrocarbons and rich (i.e., ~20%) EOM. In general, the composition of the
Relation to the OWC hydrocarbon fraction of the EOM from these lithologies
is similar to that observed in the more porous reservoir
Although the filling history of the Snorre Field is lithologies (e.g., generally similar saturate/aromatic
not the main object of this study, some control on the ratios). Leythaeuser and Rückheim (1989) reported that
reservoir hydrocarbons is required if the significance the n-alkane composition of finer-grained arenaceous
of the oil shows in the overlying mud rocks is to be lithologies is often biased toward the lower-molecular-
properly investigated. It has been demonstrated else- weight compounds than are the n-alkanes in the coarse-
where that the molecular composition of reservoir oil grained lithologies. A similar bias was observed in
in a field may vary according to the filling history of some of the samples from the study wells, and was
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 121

Figure 5. Summary diagram showing variation in extractable organic matter concentration and composition,
and organic carbon content in well 34/7-1. (Depth axis is not to scale.)

most pronounced in a siltstone from 34/7-4 (e.g., below the interval of highest oil saturation. Taken with
2538.50 m) and well-cemented, fine-grained sandstones the lack of variation in other maturation parameters
in 34/7-1 (e.g., 2580.45 m) and 34/4-1 (e.g., 2560.95 m). (e.g., %C29 20S steranes and %C32 22S hopanes), these
This bias was least well developed in the EOM from the variations appear to reflect the increasing alteration of
two lowermost samples in well 34/7-4, both of which the oil rather than variations in maturation. Care is
show evidence of degradation attributable to water- therefore needed in the interpretation of biomarker
washing (Figure 7). data from reservoir intervals. The pattern is not as well
This trend in n-alkane composition does not, how- displayed in well 34/7-1 due to the occurrence of a sig-
ever, hold for mudstones within the reservoir. Within nificant concentration (3500 ppm) of undegraded, oil-
the mudstones, the n-alkane composition varies like EOM in sandstone ~5 m below the oil-water
widely from a low-molecular-weight n-alkane bias contact (OWC).
(Figure 7) to a high-molecular-weight n-alkane bias. It As mentioned previously, significant concentra-
is therefore obvious that the distribution of oil-derived tions of apparently undegraded EOM are present
EOM within the heterogeneous reservoir rocks of the below the OWC in wells 34/7-1 and 34/4-1. The n-
Snorre Field is more complex than described by alkane composition of this material is similar to that of
Leythaeuser and Rückheim (1989). the DST oils, while the biomarker composition of the
The C 27, C 28, C 29 ααα-sterane composition of the sample from 34/7-1 (2590.55 m) is identical to that of
extracts shows no significant variation within the the DST oils and of EOM from the rich, coarse-grained
reservoir samples, suggesting that the EOM shares a lithologies above the OWC. These data may suggest
common source similar to what produced the DST oils that either the OWC has moved upward or that the
(Figure 8). In well 34/7-4, the highest oil saturation OWC has been wrongly calculated in these two wells.
occurs above 2560 m, below which there is a tendency
for the Tm/Ts ratio and the relative abundance of tri- Composition of Hydrocarbons in the
cyclic diterpanes to increase with depth (Figure 9). Cretaceous–Tertiary Succession
These contradictory indications of thermal maturity The presence of significant (780–1850 ppm) EOM
coincide with a decrease in EOM concentration and an concentrations in claystones of the Shetland Group
increase in the degradation of the EOM with depth was demonstrated in the “dry” well 34/7-2 to the
122 Leith and Fallick

Percent Composition EOM Fraction (ppm) TOC (%)


0% 50 % 100 % 0 4000 8000 12000 16000 0.0 0.5 1.0 1.5 2.0
1270.00
Hordaland
1570.00
Gp.
1720.00
1830.00 Rogaland Gp.
1865.00
1922.00
1982.00
2090.00
2180.00
Shetland
2252.00
Gp.
Depth (m)

2324.00
2396.00
2432.00
2486.00
2533.05
Dunlin Gp.
2538.50
2545.00 Oil Column
Statfjord Fm.

2548.50
2555.05
2562.00
2573.00
2621.00
2689.00
Lunde Fm.

ASPH (%) Polar (%) HC (%) ASPH (ppm) EOM (ppm)

Figure 6. Summary diagram showing variation in extractable organic matter concentration and composition,
and organic carbon content in well 34/7-4. (Depth axis is not to scale.)

south of the main Snorre Field (Leith et al., 1993). Vari- the relatively mature sterane signature of the EOM.
ations in EOM content were accompanied by changes The maturities implied by these data [i.e., similar to
in the n-alkane composition of the EOM. This changed that of the DST oils (Figure 9)] are clearly inconsistent
from a bimodal pattern in the upper half of the Dunlin with the maturity of the kerogen.
Group to a more oil-like unimodal pattern at the base With the exception of the uppermost and lower-
of the Shetland Group (Figure 10). The observed most samples from the Shetland Group in 34/7-1 and
decrease in EOM concentration in the Shetland Group 34/7-4, the EOM concentration shows relatively little
was accompanied by an increase in the prominence of variation (Figure 5). The extractability of the calcare-
the low-molecular-weight n-alkanes on the saturated ous rocks at the base of the Shetland Group is similar
hydrocarbon gas chromatograms. to that found in marls of the Cromer Knoll Group (823
Equally surprising after analysis of samples from ppm). The marked decrease in EOM concentration
the Cretaceous–Tertiary succession in wells 34/7-1 between the two uppermost samples from the Shet-
and 34/7-4 was the similarity in the pattern of EOM land Group, (that is, from ~300° ppm at 1920–1930 m
concentration in these two wells and in 34/7-2. In all to ~500–800 ppm at 1840–1860 m) in both study wells
three cases, EOM concentrations ≤2000–3000 ppm coincides with a change to the more calcareous lithol-
occur in claystones of the Shetland Group. The clay- ogy of the upper Jorsalfare Formation. Extractable
stones of the Shetland Group are too immature and organic matter (EOM) concentrations in the overlying
contain too little generation potential to account for Tertiary Rogaland and Hordaland groups show a
such high concentrations of EOM (Figure 4). Given the more gradual decrease from the levels observed at the
lack of any other obvious source horizons, the most top of the Shetland Group.
obvious origin for these hydrocarbons is within the The carbonate content of the claystones in the
reservoir interval. Shetland Group appears to play a significant role in
Similar isoprenoid (Figure 11) and C 27 , C 28 , C 29 determining EOM content. This is not unexpected,
ααα-sterane (Figure 8) compositions from the Shet- since calcareous claystones tend to be less porous than
land Group in all three wells suggest a common source noncalcareous claystones. This is supported by the
for the EOM. The similarity between the EOM and the occurrence of another EOM-poor interval at 2292–2201
Snorre DST oils suggests that the source for the EOM m in 34/7-1, which coincides with a more calcareous
in the Shetland Group and the DST oils was the same. interval evident on the resistivity and sonic logs (Fig-
Additional evidence for a nonindigenous source for ure 12). It might be suggested, based on the above
the EOM in the Cretaceous–Tertiary succession lies in observations, that the 50–80- m- thick calcareous
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 123

35 Figure 7. N-alkane profiles


from reservoir intervals in
Reservoir (Triassic Lunde Fm.) wells 34/7-1 and 34/7-4.
30
Well 34/7-1 Profiles from typical Snorre
percent composition

25
2397.40 DST oils are shown for
2575.65 comparison.
20 2580.45
2585.55
15 2590.55
2595.05
10 2612.95

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34
n-alkane carbon number

25

Reservoir (Lower Jurassic Statfjord Fm.)


20 Well 34/7-4
percent composition

2538.5
15 2548.5
2555.05
2562
10 2621
2689

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34

n-alkane carbon number

10

Typical DST oils


8 (Lunde & Statfjord Reservoirs)
Percent composition

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34

n-alkane carbon number


124 Leith and Fallick

Figure 8. C27, C28, C29 ααα-


sterane composition of
extractable organic matter
compared to DST oils from
the Snorre Field.

interval at the top of the Jorsalfare Formation plays an to remain relatively constant throughout the Cretaceous–
important role in determining the EOM content of the Tertiary succession. Asphaltene-rich EOM might be
overlying Tertiary succession. expected to block up pore throats, thereby restricting
Carbon isotope data from carbonate cements at the porosity and permeability in otherwise undercom-
top and base of the Shetland Group show clear indica- pacted mudrocks (Dahl and Speers, 1986; Larter et al.,
tions of a significant contribution from organic carbon 1990). A decrease in porosity due to the filling of pore
(Table 1). A marine (bi)carbonate source would have space by asphaltene-rich EOM would not necessarily be
δ 13C ~0‰, whereas organic-derived carbon would evident from well-log data, as is evidenced by the prob-
likely have δ13C between –15 and –25‰. The carbon- lems encountered in successfully predicting tar-mat
ate cements from the top of the Shetland Group appear occurrence (Dahl and Speers, 1986).
to be slightly more depleted in 13C than do those at the The similarity in the composition of the EOM in the
base of the unit, whereas the oxygen isotope composi- Cretaceous–Tertiary succession of wells 34/7-1, 34/7-4,
tions are broadly similar. The significance of this is and 34/7-2 is not reflected in the much more variable n-
currently unclear, although a more detailed carbonate alkane and biomarker compositions of the EOM.
cement isotopic study could elucidate the role of In well 34/7-4, a marked change in n-alkane dis-
hydrocarbon leakage. tribution toward a low-molecular-weight–biased
The indication of an organic carbon contribution to distribution occurs from the Cromer Knoll Group
carbonate cements at what may have been the sedi- and upward into the Shetland Group (Figure 13).
ment–water interface when filling of the Snorre Field This trend in the 34/7-4 succession is accompanied
started (Leith et al., 1993) may suggest that the leakage by the appearance of ββ-hopane and hopene peaks
of hydrocarbons through a poorly compacted argilla- in samples from the Shetland groups (Figure 14). The
ceous cap rock could have contributed to the sealing proportion of these compounds steadily increases
capacity of the cap-rock succession (i.e., the develop- upward through the Cretaceous–Tertiary succession
ment of a largely impermeable layer at the paleo- and reaches a maximum in samples from the Roga-
sediment/water interface). Such a situation would be land and Hordaland groups. Saturated-hydrocarbon
consistent with the lack of evidence for an input of gas chromatograms from Tertiary samples are typi-
mature oil to the Tertiary succession, which overlies cally distinguished from samples in the Shetland
the Shetland Group. Group by a nearly bimodal n-alkane distribution and a
One obvious feature of the EOM in the Creta- marked odd-carbon number preference (Figure 13).
ceous/Tertiary rocks that overlie the two study wells There is a tendency for the relative abundance of
(Figure 5) is the increased asphaltene content the ββ-hopane and hopene compounds to increase
(>30%–50%) relative to that in the reservoir EOM with distance from the top of the reservoir interval and
(<5%). This high relative asphaltene content appears with the organic richness of the enclosing rocks. The
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 125

(a) Figure 9. Maturation-


7 0.0 0 dependent biomarker ratios
illustrating similarity in
maturity of extractable
6 0.0 0
% C 3 1 + exte nd e d h op a n es

organic matter from wells


34/7-1 and 34/7-4, and DST
5 0.0 0 oils from the Snorre Field.
(See Table 2 for definition
of the biomarker ratios.)
4 0.0 0

3 0.0 0

Tertiary Mudstones
2 0.0 0
Cretaceous Mudstones

Reservoir Rocks
1 0.0 0
DST Oils

0.00

20 .0 0 40 .0 0 60 .0 0 80 .0 0
% C 2 9 2 0 S ste ra n es
(b)
1 .00
Tertiary M u dstones
R e l ab u n re arran g e d ste ran e s

C re taceous M udsto nes


0 .80
R es ervoir R ocks

D ST O ils

0 .60

0 .40

0 .20

0 .00

0 10 20 30 40 50 60 70 80 90
% C 2 9 ββ ste ra ne s

extreme immaturity associated with these com- The hypothesis that the EOM in the Cretaceous–
pounds (Simoneit, 1986) makes it unlikely that their Tertiary succession represents a mixture of leaked oil
relative contribution could be enhanced by migra- and indigenous EOM is supported by the generally
tion. Most experimental (Zhao-an and Philp, 1987; mature sterane composition of the EOM (Figure 10).
Bonilla and Engel, 1988) and natural (Seifert and The relatively minor variation in sterane composition
Moldowan, 1978; Seifert et al., 1979) observations through the Cretaceous–Tertiary succession suggests
indicate preferential enrichment of more thermally that the input of immature contaminant is biased
stable biomarkers with migration. It therefore seems toward the hopanes. This is emphasized by the rapid
likely that the presence of these compounds repre- increase in Tm/Ts ratio up the succession (Figures 15,
sents a contribution from traces of indigenous imma- 16), and by the increasing divergence of the C 31
ture EOM in the enclosing rocks. 22S:22R ratio and the C 32 22S:22R ratio. The latter
126 Leith and Fallick

Table 2. Biomarker Ratios Used in This Study.

Biomarker Ratio Ratio Definition


%C31 extended hopanes 100 × average of C31αβS
(C31αβS + C31αβR)
…C34αβS
(C34αβS + C34αβR) in m/z 191

%C29 20S steranes 100 × C29ααS


(C29ααS + C29ααR) in m/z 217

Relative abundance of rearranged steranes 100 × C27δβS


(C27δβS + C27ααR) in m/z 217

%C29 ββ steranes 100×2 × (C29ββR + C29ββS)


[C29ααS + C29ααR + 2(C29ββR + C29ββS)] in m/z 217

Tricyclic Terpanes C24 tricyclic/hopane in m/z 191

Tm/Ts C27Tm/C27Ts in m/z 191

Bisnorhopane/hopane C28 28,30-bisnorhopane/C30 hopane in m/z 191

C31 22S Ratio 100 × average of C31αβS


(C31αβS + C31αβR) in m/z 191

C32 22S Ratio 100 × average of C32αβS


(C32αβS + C31αβR) in m/z 191

observation reflects input of a compound that coelutes The biomarker composition of the EOM from the
with the C31 22R peak, possibly gammacerane. Shetland Group in well 34/7-1 shows more similarity
The n-alkane composition of samples from the with that observed in well 34/7-4. As in 34/7-4, peaks
Cretaceous–Tertiary succession in well 34/7-1 shows a representing ββ-hopanes and hopene compounds
more complex pattern of variation than that observed begin to appear in the mass fragmentograms of the
in 34/7-4. This might be expected given that the evi- lowermost Shetland Group. These compounds show a
dence from cuttings gas and oil shows suggests that tendency to increase in relative abundance toward the
more leakage occurred from the 34/7-1 reservoir inter- top of the Shetland Group and into the Tertiary succes-
val (Leith et al., 1993). In well 34/7-1, the change to a sion, although the content of these compounds is not as
consistently low-molecular-weight n-alkane distribu- pronounced in the 34/7-1 well. This may reflect the
tion only occurs across the boundary between the generally lower organic richness of the Cretaceous/Ter-
Shetland and Rogaland groups. Although samples tiary succession in 34/7-1. The gradual increase in the
from 2030–2039 m and 2156–2165 m in the Shetland relative abundance of ββ-hopanes and hopenes is
Group also show similar n-alkane distributions, the punctuated in 34/7-1 by intervals with a more conven-
reason remains unclear. tional hopane/moretane–dominated terpanoid com-
Samples from the Shetland Group in well 34/7-1 position (e.g., 2066–2075 m, 2120–2129 m, and
are typically characterized by a bimodal or high- 1830–1840 m).
molecular-weight–biased n-alkane distribution similar The more mature biomarker distributions found in
to that observed in the Dunlin Group in well 34/7-2 the Shetland Group coincide with either an organic-
(Leith et al., 1993). No systematic change in n-alkane lean enclosing rock (1830–1840 m) or an interval con-
composition can be recognized in the samples from taining intervals of anomalously high interval transit
the Shetland Group. Nevertheless, there is a tendency time (2066–2075 m and 2120–2029 m). Since there is
for the higher-molecular-weight compounds to no obvious evidence available to this study of prob-
become more prominent toward the top of the Shet- lems encountered during logging of the lower two
land Group (Figure 13). intervals, the increased interval transit time may be
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 127

35 Figure 10. Variation in


n-alkane profile through
Tertiary Mudstones the Cretaceous–Tertiary
30
Well 34/7-1 succession in well 34/7-1.
percent composition

25
1440
20 1600
1660
15 1760
1820
10

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34
n-alkane carbon number

18

16 Cretaceous Mudstones (Jorsalfare Fm.)


Well 34/7-1
14
percent composition

12 1840

10 1931
1967
8 1985
2039
6

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34

n-alkane carbon number

12

Cretaceous Mudstones (Kyrre Fm.)


10 Well 34/7-1 2075
percent composition

2093
8 2129
2165

6 2201
2237
2273
4
2309
2345
2

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34

n-alkane carbon number


128 Leith and Fallick

Figure 11. Isoprenoid


composition of extractable
organic matter and DST
oils from the Snorre Field.

cautiously interpreted as indicating increased poros- and the wells are relatively remote from any major
ity (Figure 12). In the absence of any indications for fracture systems. If fracturing were responsible for the
abrupt changes in lithology, it can be tentatively sug- distribution of EOM in the Shetland Group, then dif-
gested that these intervals represent fracturing in the ferent EOM concentration patterns would be expected
drilled section. This is not unreasonable, since 34/7-1 in the three wells. It seems more likely that the pres-
is located close to the heavily faulted area associated ence of fracturing in the Cretaceous–Tertiary succes-
with the Snorre Escarpment/Inner Snorre Fault. As sion merely modifies the composition of the EOM by
was noted in well 34/7-4, the intervals showing a sig- creating intervals of variable porosity.
nificant content of ββ-hopanes and hopenes coincide The evidence discussed in this section suggests
with anomalously high Tm/Ts ratios and divergence that leakage of hydrocarbons from the reservoir into
of the 22S:22R ratios for the C31 and C32 αβ-hopanes the overlying succession has occurred. Biomarker
(Figures 15, 16). data indicate that contamination of the leaked
The content of tricyclic diterpanes is relatively con- hydrocarbons by trace amounts of immature indige-
stant throughout the Shetland Group in 34/7-1 (Figure nous EOM has also occurred. The degree of contami-
10), only increasing in the uppermost Tertiary sample. nation increases away from the top of the reservoir
In contrast, in 34/7-4, the relative abundance of tri- interval, except where intervals of increased porosity
cyclic diterpanes increases across the boundary or reduced organic matter content allow accumula-
between the Dunlin and Cromer Knoll groups, and tion of uncontaminated leaked oil. Consistent varia-
decreases only gradually up the Shetland Group to tions in the concentration and composition of EOM
reach a minimum in the Tertiary samples. In these two in the suprareservoir succession of a number of wells
wells, a high relative abundance of tricyclic diterpanes suggest that it is difficult to propose fracturing as the
coincides with the change to an n-alkane distribution main leakage mechanism. Rather, movement and sub-
dominated by low-molecular-weight compounds. The sequent fractionation of oil through an undercom-
occurrence of this change at a lower interval in 34/7-4 pacted succession seems to be more consistent with
would be consistent with the less extensive leakage the analytical data obtained during this study.
suggested by cuttings gas data. Relationship Between EOM Occurrence and
Given the sample spacing used in this study, it is dif- Composition and Variation in Cuttings Gas- and
ficult to reach any definite conclusions regarding the Oil-Show Records
role of fracturing in the distribution of EOM derived
from the reservoir oil. However, the similarity of the The distribution of EOM suggests a division of the
trends in EOM concentration and composition in the Shetland Group into an approximately 400-m- thick
Cretaceous–Tertiary succession of the study wells interval characterized by relatively high EOM concen-
tends to argue against an important role for fracturing trations (2000–4000 ppm) and bounded by more calcare-
in these wells. The succession penetrated by wells ous intervals at top and bottom. EOM concentrations in
34/7-2 and 34/7-4 shows no evidence of fracturing, these calcareous intervals are ~1 order of magnitude
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 129

Figure 12. Selected


log curves through
the Cretaceous–
Tertiary succession
in well 34/7-1.
The solid lines are,
from left to right:
gamma, sonic (dT),
and density
(RHOB). The bro-
ken lines are, from
left to right: caliper,
resistivity (Rild),
and calculated shale
porosity (after
Magara, 1978).

lower than is observed in the intervening less calcareous be correlated with any variations in either EOM concen-
interval. This contrasts with the absence of visible oil- tration or composition.
show records above ~2070–2090 m across the Snorre The generally decreasing concentration of C2–C 4
Field (Table 3). This absence of visible oil shows cannot and C5+ hydrocarbons in the cuttings gas above the

30 Figure 13. Variation in


n-alkane profile through
Tertiary/Cretaceous Mudstones: the Cretaceous–Tertiary
25
Well 34/7-4 succession in well 34/7-4.
percent composition

1270
20 1830
1865
15 1982
2180
2396
10
2522

0
nC15
nC16
nC17
nC18
nC19
nC20
nC21
nC22
nC23
nC24
nC25
nC26
nC27
nC28
nC29
nC30
nC31
nC32
nC33
nC34

n-alkane carbon number


130 Leith and Fallick

Composition Composition
0% 50 % 100 % 0% 20 % 40 % 60 % 80 % 100 %
1600.00 1270.00
Hordaland Hordaland
Gp. Gp.
1820.00
Rogaland 1830.00 Rogaland
Gp. Gp.
1840.00

Jorsalfare
1865.00

Fm.
1931.00

Jorsalfare Fm.

Shetland Gp.
2180.00
1967.00

Kyrre Fm.
2039.00

Shetland Gp.
2396.00

2075.00
2522.00
Depth (m)

Depth (m)

Dunlin Gp.
2129.00
Kyrre Fm.

2533.05
2201.00

Oil Column
2345.00 2548.50

2397.40

Statfjord Fm.
D 2555.05
Oil Column

2580.45
Lunde Fm.

2562.00
2585.55
E 34/7-4
34/7-1 2621.00
2590.55

Hegre Gp.
2595.05 2689.00

Tricyclic ββ-hopanes

αβ-hopanes Hopenes

Moretanes
Figure 14. Terpanoid composition summary diagram for wells 34/7-1 and 34/7-4. Compound group percentages
are based on peaks identified on m/z 191 mass fragmentograms. (Depth axis is not to scale.)

reservoir interval describes a curve similar to what pattern has occurred, probably reflecting the pressure
might be expected from diffusion as described by system associated with the underlying oil column.
Leythaeuser et al. (1982), Krooss and Leythaeuser The distribution of gaseous and extractable hydro-
(1988), and Krooss et al. (1988). This gradual decrease carbons in the Cretaceous/Tertiary strata of the study
in the C2+ cuttings gas components is punctuated in wells seems to reflect two different processes. England
wells 34/7-1 and 34/7-4 by rapid changes over short and Mackenzie (1988) have illustrated the different
intervals at different levels in the Shetland Group. The time spans required for changes in different oil fractions
occurrence of these events can be related to the height to occur. The apparent relationship between cuttings gas
of the reservoir oil column (Leith et al., 1993), suggest- and the oil column may suggest that the cuttings gas dis-
ing that modification of an otherwise normal diffusion tribution more directly reflects the situation in the oil
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 131

Ratio Value Ratio Value


0.0 0.5 1.0 1.5 2.0 0 20 40 60 80 100

1600.00

1820.00

1840.00

1931.00

1967.00

2039.00

2075.00
Depth (m)

2129.00

2201.00

2345.00

2397.40

2580.45

2585.55

2590.55

2595.05

Tricyclic terpanes Tm/Ts C31 22S Ratio C32 22S Ratio


Bisnorhopane/hopane abun C27 diasteranes %C29 20S steranes %C29 ββ steranes

Figure 15. Summary diagram showing vertical variation in selected biomarker ratios calculated from m/z 191
and m/z 217 mass fragmentograms for well 34/7-1. The position of the reservoir oil column is indicated by the
shaded box. (Depth axis is not to scale. See Table 2 for definition of the biomarker ratios.)

column. In contrast, the generally even distribution of strong control on the distribution of EOM. Isotopic evi-
EOM through undercompacted strata of relatively dence suggests that migrating hydrocarbons have prob-
constant porosity appears to reflect a more gradual ably contributed to the development of at least those
process analogous to the filling of a reservoir horizon, carbonate cements at the top of the Shetland Group.
albeit of very poor quality. This opens the possibility that vertical cap-rock leakage
may, in fact, enhance the sealing capacity of a structure
CONCLUSIONS under certain conditions. The timing and distribution of
these calcareous cements in the Shetland Group and
The results obtained during this study indicate that their significance to the sealing capacity of the Creta-
the relationships among a reservoir horizon, the reser- ceous succession have important implications to the
voir oil, and the overlying sealing cap-rock horizon development of more subtle prospects.
may be complex. In the Snorre Field, this complexity is Differences in the variation of gaseous hydrocar-
evident in the relatively even distribution of signifi- bons and liquid hydrocarbons in the Cretaceous–
cant concentrations of oil-derived EOM throughout Tertiary strata overlying the Snorre Field suggest that
undercompacted mud rocks of Late Cretaceous age. two processes may be controlling these fractions. Dif-
The absence of the lower three formations of the fusion appears to be important in controlling the vari-
Shetland Group may have influenced the poor sealing ation of the gaseous hydrocarbons, although the
capacity of the Shetland Group over the Snorre Field, concentration of C2+ hydrocarbons may be modified
especially since the lowest unit, the Svarte Formation, is by other factors, either lithological or pressure-related.
typically composed of marls and interbedded lime- In contrast, the EOM concentration appears to be con-
stones (Nybakken and Bäckstrøm, 1989). The calcareous trolled by the carbonate content of the Cretaceous
content of the Shetland Group appears to represent a mudstones, sufficient time having passed to permit
132 Leith and Fallick

Ratio Value Ratio Value


0.0 0.5 1.0 1.5 2.0 2.5 3.0 0 20 40 60 80 100

1270.00

1830.00

1865.00

2180.00

2396.00
Depth (m)

2522.00

2533.05

2548.50

2555.05

2562.00

2621.00

2689.00

Tricyclic terpanes Tm/Ts C31 22S Ratio C32 22S Ratio


Bisnorhopane/hopane abun C27 diasteranes %C29 20S steranes %C29 ββ steranes

Figure 16. Summary diagram showing vertical variation in selected biomarker ratios calculated from m/z 191
and m/z 217 mass fragmentograms for well 34/7-4. The position of the reservoir oil column is indicated by the
shaded box. (Depth axis is not to scale. See Table 2 for definition of the biomarker ratios.)

some degree of homogenization to occur. The leak- If the EOM in the Shetland Group is considered an
age of hydrocarbons to the Shetland Group may extension of the reservoir oil column, then the accompa-
therefore involve a rapid diffusive process affecting nying attenuation of the oil column may permit the
the gaseous hydrocarbons, and a slower mass trans- uppermost calcareous interval of the Shetland Group to
port process affecting the EOM, possibly related to act as a seal. This is supported by the high proportion of
microfracturing. indigenous input to the EOM in the Tertiary strata.

Table 3. Summary of Cuttings Gas and Oil Shows in the Study Wells.*

NOCS** NOCS NOCS


34/7-1 34/7-4 34/4-1
Gas Wetness >50% 2066 2396 ?
Gas Wetness >5% 1790 1937 ?
C5+ = 0 1430 2135 ?
C5+ > 5000 µl/kg 2066 2995 ?
C5+ >10000 µl/kg 2084 2995 ?
Uppermost oil show 2093 2078 2050

*Depths given in meters.


**NOCS = Norwegian Continental Shelf.
Organic Geochemistry of Cap-Rock Hydrocarbons, Snorre Field, Norwegian North Sea 133

ACKNOWLEDGMENTS Hollander, N.B., 1987, Snorre, in A.M. Spencer,


E. Holter, C.J. Campbell, S. Hanslien, P.H.H.
Elf Aquitaine Norge A/S is gratefully acknowl- Nelson, E. Nysæther, and E. Ormaasen, eds., Geol-
edged for funding most of the work reported in this ogy of the Norwegian oil and gas fields: London,
chapter. The license committee for the Snorre Field is Graham & Trotman, p. 307–318.
acknowledged for permission to publish these results. Horstad, I., S.R. Larter, H. Dypvik, P. Aagard, A.M.
The Scottish Universities Research and Reactor Centre Bjornvik, P.E. Johansen, and S. Eriksen, 1990,
is supported by Natural Environmental Research Degradation and maturity controls on oil field
Council (N.E.R.C.) and the Consortium of Scottish petroleum column heterogeneity in the Gullfaks
Universities. I would also like to thank D.A. Leith, Field, Norwegian North Sea: Organic Geochem-
D.K. Baskin, and an anonymous reviewer for their istry, v. 16, p. 497–510.
assistance in improving this manuscript. Irwin, H., 1989, Hydrocarbon leakage, biodegradation
and the occurrence of shallow gas and carbonate
REFERENCES CITED cement (abs.): Norwegian Petroleum Society (NPF),
Shallow Gas and Leaky Reservoirs Conference,
Berg, R.R., 1975, Capillary pressures in stratigraphic Stavanger, p. 21.
traps: AAPG Bulletin, v. 59, p. 939–956. Karlsen, D.A., and S.R. Larter, 1989, A rapid correla-
Bonilla, J.V., and M.H. Engel, 1988, Chemical alter- tion method for petroleum mapping within indi-
ation of crude oils during simulated migration vidual petroleum reservoirs: applications to
through quartz and clay minerals:. Organic Geo- petroleum reservoir description, in J.D. Collinson,
chemistry, v. 13, p. 503–512. ed., Correlation in petroleum exploration: London,
Caillet, G., 1993, The cap-rock of the Snorre Field (Nor- Graham & Trotman, p. 77–85.
way): possible leakage by hydraulic fracturing: Karlsson, W., 1986, The Snorre, Statfjord and Gullfaks
Marine and Petroleum Geology, v. 10, p. 42–50. oilfields and the habitat of hydrocarbons on the
Dahl, B., and G.C. Speers, 1985, Organic geochemistry Tampen Spur, offshore Norway, in A.M. Spencer, E.
of the Oseberg Field (I), in B.M. Thomas, A.G. Doré, Holter, C.J. Campbell, S. Hanslien, P.H.H. Nelson,
S.S. Eggen, P.C. Home, and R.M. Larsen, eds., E. Nysæther, and E. Ormaasen, eds., Habitat of
Petroleum geochemistry in exploration of the Nor- hydrocarbons on the Norwegian Continental Shelf:
wegian Shelf, Norwegian Petroleum Society: Lon- London, Graham & Trotman, p. 181–197.
don, Graham & Trotman, p. 185–195. Krooss, B.M., and D. Leythaeuser, 1988, Experimental
Dahl, B., and G.C. Speers, 1986, Geochemical charac- measurements of the diffusion parameters of
terization of a tar-mat in the Oseberg Field, Norwe- hydrocarbons in water-saturated sedimentary
gian Sector, North Sea: Organic Geochemistry, rocks—II: Results and geochemical significance:
v. 10, p. 547–558. Organic Geochemistry, v. 12, p. 91–108.
Davis, R.W., 1987, Analysis of hydrodynamic factors Krooss, B.M., D. Leythaeuser, and R.G. Schaeffer,
in petroleum migration and entrapment: AAPG 1988, Light hydrocarbon diffusion in a cap-rock:
Bulletin, v. 71, p. 643–649. Chemical Geology, v. 71, p. 65–76.
D’Heur, M., 1987, Albuskjell, in A.M. Spencer, E. Holter, Larter, S.R., K.O. Bjørlykke, D.A. Karlsen, T. Nedkvitne,
C.J. Campbell, S. Hanslien, P.H.H. Nelson, E. T. Eglinton, P.E. Johansen, D. Leythaeuser, P.C.
Nysæther, and E. Ormaasen, eds., Geology of the Mason, A.W. Mitchell, and G.A. Newcombe, 1990,
Norwegian oil and gas fields, Norwegian Petroleum Determination of petroleum accumulation histories:
Society: London, Graham & Trotman, p. 51–62. examples from the Ula Field, Central Graben, Nor-
Downey, M.W., 1984, Evaluating seals for hydrocarbon wegian North Sea, in A.T. Buller, E. Berg, O. Hjelme-
accumulations: AAPG Bulletin, v. 68, p. 1752–1763. land, J. Kleppe, O. Torsæter, and J.O. Aasen, eds.,
England, W.A., and A.S. Mackenzie, 1989, Some North Sea oil and gas reservoirs—II: London,
aspects of the organic geochemistry of petroleum Graham & Trotman, p. 319–330.
fluids in the subsurface: Geologische Rundschau, Leith, T.L., I. Kaarstad, J. Connan, J. Pierron, and G.
v. 78, p. 291–303. Caillet, 1993, Recognition of cap-rock leakage in the
England, W.A., D.M. Mann, and T.M. Quigley, 1987, Snorre Field, Norwegian North Sea: Marine and
The movement and entrapment of petroleum fluids Petroleum Geology, v. 10, p. 29–41.
in the subsurface: Journal of the Geological Society Leonard, R.C., and J.W. Munns, 1987, Valhall, in A.M.
of London, v. 144, p. 327–347. Spencer, E. Holter, C.J. Campbell, S. Hanslien,
Hogg, A.J.C., M.J. Pearson, and A.E. Fallick, 1993, Pre- P.H.H. Nelson, E. Nysæther, and E. Ormaasen, eds.,
treatment of Fithian illite for oxygen isotope analy- Geology of the Norwegian oil and gas fields, Nor-
sis: Clay Minerals, v. 28, p. 149–152. wegian Petroleum Society: London, Graham &
Hole, P.A., E. Holter, K.S. Isaksen, K.S. Lervik, T. Monsen, Trotman, p. 153–164.
S. Nybakken, J.E. Tellefsen, K. Tonstad, and B.T.G. Leythaeuser, D., and J. Rückheim, 1989, Heterogeneity
Wandas, 1989, Revised Cretaceous lithostratigraphy of oil composition within a reservoir as a reflection
of the Norwegian North Sea, in D. Isaksen and K. of accumulation history: Geochimica et Cos-
Tonstad, eds., A revised Cretaceous and Tertiary mochimica Acta, v. 53, p. 2119–2123.
lithostratigraphic nomenclature for the Norwegian Leythaeuser, D., R.G. Schaeffer, and A. Yukler, 1982, Role
North Sea: Oljedirektoratet Bulletin, no. 5, p. 16–35. of diffusion in primary migration of hydrocarbons:
134 Leith and Fallick

AAPG Bulletin, v. 66, p. 408–429. P. Songstad, 1987, Hild, in A.M. Spencer, E. Holter,
Magara, K., 1978, Compaction and fluid migration: C.J. Campbell, S. Hanslien, P.H.H. Nelson, E.
Practical petroleum geology: Amsterdam, Elsevier, Nysæther, and E. Ormaasen, eds., Geology of the
319 p. Norwegian oil and gas fields, Norwegian Petroleum
McCrea, J.M., 1950, On the isotopic geochemistry of Society: London, Graham & Trotman, p. 287–294.
carbonates and a paleotemperature scale: Journal of Schowalter, T.T., 1979, Mechanics of secondary hydro-
Chemistry and Physics, v. 18, p. 849–857. carbon migration and entrapment: AAPG Bulletin,
Nipen, O., 1987, Oseberg, in A.M. Spencer, E. Holter, v. 63, p. 723–760.
C.J. Campbell, S. Hanslien, P.H.H. Nelson, Seifert, W.K., and J.M. Moldowan, 1978, Applications
E. Nysæther, and E. Ormaasen, eds., Geology of of steranes, terpanes and mono-aromatics to the
the Norwegian oil and gas fields, Norwegian maturation, migration and source of crude oils:
Petroleum Society: London, Graham & Trotman, Geochimica et Cosmochimica Acta, v. 42, p. 77–95.
p. 379–388. Seifert, W.K., J.M. Moldowan, and R.W. Jones, 1979,
Nybakken, S., and S.A. Bäckstrøm, 1989, Shetland Application of biological marker chemistry to
Group: stratigraphic subdivision and regional corre- petroleum exploration: Proceedings of the 10th
lation in the Norwegian North Sea, in J.D. Collinson, World Petroleum Congress, Paper SP8, p. 425–440.
ed., Correlation in petroleum exploration: London, Simoneit, B.R.T., 1986, Cyclic terpenoids of the
Graham & Trotman, p. 253–269. geosphere, in R.B. Johns, ed., Biological markers in
Pekot, L.J., and G.A. Gersib, 1987, Ekofisk, in A.M. the sedimentary record: Amsterdam, Elsevier,
Spencer, E. Holter, C.J. Campbell, S. Hanslien, p. 43–100.
P.H.H. Nelson, E. Nysæther, and E. Ormaasen, eds., Spencer, A.M., E. Holter, C.J. Campbell, S. Hanslien,
Geology of the Norwegian oil and gas fields, Nor- P.H.H. Nelson, E. Nysæther, and E. Ormaasen,
wegian Petroleum Society: London, Graham & 1987, Geology of the Norwegian oil and gas fields,
Trotman, p. 73–88. Norwegian Petroleum Society: London, Graham &
Radke, M., H. Willsch, and D.H. Welte, 1980, Prepara- Trotman, 493 p.
tive hydrocarbon group type determination by Wachter, E.A., and J.M. Hayes, 1985, Exchange of oxy-
automated medium pressure liquid chromatogra- gen isotopes in carbon dioxide phosphoric acid sys-
phy: Analytical Chemistry, v. 52, p. 406–411. tems: Chemical Geology (Isotope Geosciences
Roberts, W.H., 1980, Design and function of oil and Section), v. 52, p. 365–374.
gas traps, in W.H. Roberts and R.J. Cordell, eds., Welte, D.H., W. Stoessinger, R.G. Schaeffer, and M.
Problems of petroleum migration: AAPG Studies in Radke, 1984, Gas generation and migration in the
Geology 10, p. 217–240. Deep Basin of western Canada, in J.A. Masters, ed.,
Rosenbaum, J., and S.M.F. Sheppard, 1986, An isotopic Elmworth: case study of a deep gas basin: AAPG
study of siderites, dolomites, and ankerites at high Memoir 38, p. 35–47.
temperatures: Geochimica et Cosmochimica Acta, Zhao-an, F., and R.P. Philp, 1987, Laboratory biomarker
v. 50, p. 1147–1150. fractionations and implications for migration studies:
Rønning, K., C.D. Johnston, S.E. Johnstad, and Organic Geochemistry, v. 11, p. 169–176.
Allard, D.M., 1997, Fault leak controlled trap fill: rift basin
examples, in R.C. Surdam, ed., Seals, traps, and the
petroleum system: AAPG Memoir 67, p. 135–142.

Chapter 9

Fault Leak Controlled Trap Fill:


Rift Basin Examples
D.M. Allard
Exxon Ventures (CIS) Inc.
Houston, Texas, U.S.A.

ABSTRACT
Interpretation of fault leak and seal enhances the understanding of hydro-
carbon reservoir systems. Useful applications exist for both field develop-
ment and seal risk evaluation for undrilled exploration prospects.
Basic fault seal data (e.g., logs, lithology from wells, stratigraphic models
of lithology variation away from well control, interpretations of bed juxtapo-
sitions across the fault plane, and basic well test results) are needed for the
interpreter to resolve the puzzle of trap fill limits and possible different
reservoir systems within the same faulted field. Interpretation of both pro-
ductive and tested wet structures allows for the development of conclusions
that may be applicable to different proven or potential traps across the basin.
A refined understanding of trap fill limits and hydrocarbon distribution
within a field may be gained by interpretation and calibration of fault gouge
as a seal on a basinwide basis. Evaluation of fault gouge as a seal with the
displaced section analysis methodology is labor intensive and may not
always be an effective exploration tool, considering the data control needed
to develop reliable results.
In exploration, the quality of fault seal interpretation for an undrilled
structure is dependent on map quality, well control, and the ability to predict
stratigraphic facies changes away from wells. The minimum calibration
requirement for exploration applications of a fault seal interpretation of an
undrilled structure is to have completed interpretations of bed juxtapositions
across the fault plane(s) for both productive and nonproductive structures
elsewhere in the basin.

INTRODUCTION and top seal information. Fault gouge or displaced sec-


tion analysis (DSA) (Bouvier et al., 1989) may be
Seal interpretation of a particular feature is an inte- applied to refine the interpretation of trap fill levels, if
grated assessment of trap fill levels (hydrocarbon- a basin-specific calibration has been completed. An
water contacts) based on fault plane profiles (Smith, understanding of geologic influence on top seal qual-
1980; Allen, 1989) and well data, including test results ity may be developed from mercury injection-capillary

135
136 Allard

with FPPs, and dip seal/dipleak, which is interpreted


with DSA.
An overview of the elements of a seal study method
is shown in Figure 2 in order to explain various
aspects that may be studied to better understand trap
seal. The “refine analogs” step (Figure 2) refers to
understanding what seal factors are attributable to
either fields or dry holes, given the resolution of a par-
ticular data set. This results in a seal evaluation cali-
bration to a specific basin or area.

DISCUSSION OF EXAMPLES
The examples used here are from rift basin fields cut
Figure 1. A generalized fault seal model given to by normal faults within an overall extensional struc-
explain fault seal terms. tural style. The implication for seal interpretation is
that the structural configuration is not reactivated from
a previous one, and the normal fault throws observed
pressure analysis of seal rock samples (Pittman, 1992; today are the maximum throws attained. Structural
Vavra et al., 1992). In typical exploration applications, complexity may be a more important issue in basins
data control limits fault seal interpretation to fault with multiple episodes of structuring, especially when
plane profile (FPP) analysis of bed juxtapositions and the different structural episodes overlap with the
does not support application of DSA or capillary pres- migration of hydrocarbons (Schowalter, 1979).
sure concepts. Figure 1 is a general model for graphic Clastic sand reservoirs and lacustrine shale seals
definition of cross seal/crossleak, which is interpreted make up the main lithologies of the example fields

Figure 2. A seal analysis


method focused on
improving prediction
of seal integrity in an
exploration project.
Fault Leak Controlled Trap Fill: Rift Basin Examples 137

Figure 3. Field X structure map. The faulted anticline has a structural spillpoint in the southwest, and
the only well drilled to date is located in the northeast fault block.
138 Allard

Figure 4. SW-NE schematic structural cross section depicting the fieldwide reservoir systems as interpreted
from fault seal analysis.

discussed below. Reservoir deposition occurred prior C1, C2, and C3 zones are interpreted as separate reser-
to the extensional structuring that created the present- voir systems in the east fault block. The fault leak-
day trap configuration. Well-log correlations and point interpretation, depicted in Figure 4, indicates the
depositional models suggest that reservoir sand conti- C2 and C3 in the east fault block are in pressure com-
nuity is good, and sands consistently occur in specific munication over the entire field area. If an appraisal
stratigraphic intervals. well is drilled in the central fault block, it will be
Figure 3 shows Field X in map view. Only the high- important to acquire good pressure information for
est fault block was drilled at the time of the fault seal each of the three C zones to evaluate the current inter-
study. Leak points were interpreted for both the east pretation.
fault (bounding the west side of the drilled fault block) There is no Field X repeat formation test (RFT)
and for the west fault of this faulted anticline. Oil/water pressure survey or pressure-volume-temperature
contacts for the entire structure were interpreted under (PVT) data to aid in the interpretation of this field’s
the assumption that the trap fill is not migration-charge oil/water contacts. Good seismic coverage, good
limited for the drilled and undrilled blocks. Key aspects seismic resolution, and seismic well ties at Field X
to the Field X fault seal interpretation are the cross-leak allow direct comparison of mapped structure depths
points depicted in Figure 4 and the structural spillpoint (in FPP) with well test depths. The fault’s throws are
(–5540; Figure 3) for the top of the C1 sandstone. These mapped consistently with the seismic interpretation
ultimately control the majority of the hydrocarbon vol- and, in addition, the fault throw vs. fault length
ume trapped in the C reservoirs. (Walsh and Watterson, 1988) and the lateral varia-
The well drilled in the east fault block of Field X tion of mapped fault throw are reasonable for this
tested oil in each of the C1, C2, and C3 sand zones. The rift basin structural setting. The interpretation of
Fault Leak Controlled Trap Fill: Rift Basin Examples 139

Figure 5. The Field Z struc-


ture map depicts a large
faulted anticline. The map
surface is the flooding sur-
face (FS) identified at 5844
ft on the log in Figure 6.

hydrocarbon distribution is presented as a simple explanation may involve complicated hydrocarbon


undersaturated oil/water system. migration and structural timing analysis that are
A second example, Field Z, has well control, selec- beyond the scope of this paper.
tive production tests, and pressure data, which allow a Hydrocarbon volumes in the M2 and M3 reservoirs
more refined application of fault seal techniques from are dependent in part on the fault seal concepts as
that used for Field X. In Field Z, the oil column heights depicted in Figure 8, which requires cross seal and dip
trapped in specific reservoirs are apparently con- seal to honor the structure map and well data. Fault
trolled both by leak points of highside-to-lowside jux- gouge dip seal is interpreted to cause the M2 oil/water
taposed sand beds and by fault gouge lithology. contact to occur deeper on the southwest side of the
A Field Z structure map (Figure 5) shows the clo- fault than on the northeast side of the fault, based on
sure to be a faulted anticline, based on mapping a sur- DSA results.
face located immediately above the M1 pay zone. The An important conclusion from study of Field Z, and
field stratigraphy has been simplified for use in the many other drilled structures in this basin, is that
FPP interpretation, as shown on the log in Figure 6. sand-on-sand juxtapositions will leak across faults. The
Reservoirs M2 and M3 have been simplified as shown observation of fault gouge seal control on trapped
on the left edge of Figure 6, to show gross sand-prone hydrocarbon volumes was rare, possibly due to the
zones to help simulate the stratigraphic observation sparse well control in the exploration study area. Assess-
that sand content may vary within the sand-prone ment of fault-dependent exploration prospect trap fill in
zone over the distance or area covered by each fault. this basin may use this sand juxtaposition concept.
Figure 7 is an example of an interpreted FPP from
Field Z, fault 3. Similar FPP interpretation details are
completed for each fault in the field to support the CONCLUSIONS
overall trap fill conclusions. Inspection of Figure 7
shows the field’s M1 oil/water contact occurs deeper An improved understanding of basin-specific seal
than the M1 cross-leak point. The interpretation for the integrity is obtainable from application of the fault seal
M1 reservoir accumulation is that the trapped hydro- evaluation techniques on drilled structures as
carbon volume is controlled not by the fault leak described above. The process of seal interpretation of
points in the productive area, but by other leak points productive structures may result in an improved
elsewhere on the structural trend. An alternative understanding of field hydrocarbon distribution. With
140 Allard

Figure 6. Simplified Field


Z well log. Oil-water
contacts have been
confirmed with
production tests. The
sand zones are topped
by flooding surfaces (FS),
and sequence boundaries
(SB) are interpreted to
occur at the base of some
sand beds. This figure
is shown to demonstrate
a simplication of the
stratigraphy for use in
the fault plane profile
(FPP). The yellow bars on
the left edge of the figure
depict the sand intervals
as applied in the FPP in
Figure 7, which are, in
concept, sand-prone
stratigraphic intervals.
Fault Leak Controlled Trap Fill: Rift Basin Examples 141

Figure 7. Example of a major interval fault plane profile that represents the stratigraphic juxtapositions at
Fault 3 in Field Z.

Figure 8. Field Z schematic


cross section display of
reservoir-specific hydro-
carbon distribution as
interpreted at specific
faults in Field Z, based
on fault seal concepts
and well test results.

seal analysis and calibration complete, the basin explo- the subject matter and/or suggestions for presentation
ration strategy, including a fault-dependent trap seri- of figures and text: R.G. Bellis, C.B. Boutte, C.D.
atim, may be improved. Brotherhood, F.W. Haueter, G.P. Nakayama, J.R.
Schwalbach, W.T. Shea, P.O. Yilmaz, and K. Zaud-
ACKNOWLEDGMENTS erer. In addition to discussions of the subject matter
and suggestions for presentation, the following indi-
I would like to acknowledge the following co- viduals have published Exxon proprietary research
workers, whose contributions include discussions of documents that were significant in the impact to the
142 Allard

final results published in this paper: L.H. Fairchild, Pittman, E.D., 1992, Relationship of porosity and perme-
D.W. Phelps, and A.C. Tuminas. ability to various parameters derived from mercury
Thanks to Exxon Ventures (CIS) Inc. and Exxon injection-capillary pressure curves for sandstone:
Exploration Company, for whom I worked during the AAPG Bulletin, v. 76, p. 191–198.
creation of this document, for support to allow devel- Schowalter, T.T., 1979, Mechanics of secondary hydro-
opment of the concepts and publication of the results. carbon migration and entrapment: AAPG Bulletin,
v. 63, p. 723–760.
REFERENCES CITED Smith, D.A., 1980, Sealing and nonsealing faults in
Louisiana Gulf Coast Salt Basin: AAPG Bulletin,
Allen, U.S., 1989, Model for hydrocarbon migration v. 64, p. 145–172.
and entrapment within faulted structures: AAPG Vavra, C.A., J.G. Kaldi, and R.M. Sneider, 1992, Geo-
Bulletin, v. 73, p. 803–811. logical applications of capillary pressure: a review:
Bouvier, J.D., C.H. Kaars-Sijpesteijn, D.F. Kluesner, AAPG Bulletin, v. 76, p. 840–850.
C.C. Onyejekwe, and R.C. Van Der Pal, 1989, Three- Walsh, J.J., and J. Watterson, 1988, Analysis of the
dimensional seismic interpretation and fault sealing relationship between displacement and dimensions
investigations, Nun River Field, Nigeria: AAPG of faults: Journal of Structural Geology, v. 9,
Bulletin, v. 73, p. 1397–1414. p. 1039–1046.
Boult, P.J., P.N. Theologou, and J. Foden, 1997, Capillary seals
within the Eromanga Basin, Australia: implications for
exploration and production, in R.C. Surdam, ed., Seals, traps,
and the petroleum system: AAPG Memoir 67, p. 143–167.

Chapter 10

Capillary Seals Within the Eromanga Basin,


Australia: Implications for Exploration
and Production
P.J. Boult
P.N. Theologou
Gartrell School of Mining Metallurgy and Applied Geology, University of South Australia
Pooraka, South Australia, Australia

J. Foden
Department of Geology and Geophysics, Adelaide University
Adelaide, South Australia, Australia

ABSTRACT
Studies of producing oil pools of the predominantly fluvial to shallow-
lacustrine Eromanga Basin in central Australia indicate that capillary seals
have a significant control on hydrocarbon occurrence and production. Oil
pools sealed by the Callovian Birkhead Formation at the Bodalla South,
Jackson, and Gidgealpa fields, and by the Berriasian Murta Formation at the
Murteree Ridge fields, are filled to capillary seal capacity. Resultant upward
leakage has taken place, filling oil pools higher up.
The major producing reservoir of the Eromanga Basin is the braided flu-
vial Hutton Sandstone. The transition from this to the Birkhead Formation
seal is marked by two features: (1) a change from braided fluvial to mean-
dering fluvial and (2) a change from craton-derived (CD) sediment to
volcanic-arc-derived (VAD) sediment. The CD and VAD sediments have
Sm/Nd model ages of ~1500 and 800 m.y., respectively, suggesting a major
provenance change. In the most prospective part of the basin, this prove-
nance change and the resultant switch from quartz to lithic arenites control
the location of the effective intra-Birkhead Formation seal. These lithic aren-
ites are the same grain size as the underlying reservoir-quality quartz
arenites; however, their pore-throat sizes have been reduced by diagenesis
and the precipitation of pore-filling authigenic clays.
In the Murteree Ridge area, a lacustrine, upper shoreface, intra-Murta
Formation reservoir is the primary producer. The reservoir that occurs beneath
the Murta Formation is the McKinlay Member. Between the McKinlay Member
reservoir and the intra-Murta Formation, reservoir vertical stacking of thin,
fine-grained, shallow-lacustrine, quartz arenites occurs. This has produced tor-
tuous vertical flow paths, or conduits, which have lower displacement pres-
sures and higher permeabilities than the intervening siltstone layers. These
143
144 Boult et al.

paths have allowed upward migration of oil, by capillary leakage, beyond the
McKinlay Member into the intra-Murta Formation reservoir over geologi-
cal time. These paths have recently allowed the leakage of brine through
this interpool seal during production, providing pressure maintenance to
the otherwise confined higher oil pool from the very strong water drive
of the lower pool.

INTRODUCTION diagenetically altered lithic arenites. The repeated


influx of these sediments is predominantly con-
Fluvial/lacustrine sequences are known to be trolled by the uplift of a sediment source area
extremely heterogeneous and difficult depositional along a volcanic arc to the east, and associated tec-
environments to find and produce hydrocarbons tonic warping within the basin. This volcanic
from. Three-dimensional seismic has recently enabled activity has been documented (Veevers, 1984;
some advances to be made in delineating sand body Watts, 1987; Wiltshire, 1989; Hawlader, 1990; He
geometry, and repeat formation testers have enabled and Conaghan, 1994; Whitford, 1994) to have
an assessment of connectivity of reservoirs within repeatedly provided the source material for vol-
such sequences. However, the sealing/barrier-to-flow canic-arc-derived (VAD) lithic arenites that occur
mechanisms that occur between and within reservoirs across the foreland Surat Basin and eastern part of
remain largely an underresearched area, requiring a the epicratonic Eromanga Basin. This paper
multidisciplinary understanding of geology, engineer- extends the area of deposition of these VAD lithic
ing, the physics of fluids, and seismic interpretation. In arenites further west across the Eromanga Basin
this paper we emphasize the multidisciplinary than previously documented.
approach as a method for understanding oil entrap- •Where VAD lithic arenites have sufficient thick-
ment and solving production optimization problems. ness, diagenesis creates the effective seal in the
The increased knowledge gained about migration examples discussed. The diagenesis of VAD sedi-
pathways through complicated stacked, sand-rich flu- ments involves the blocking of pore throats by
vial/lacustrine sequences should lead to better discov- authigenic clays. This process creates the
ery rates within such sequences. increased capillary pressures, which are gener-
Hydrocarbon migration pathways are predomi- ated within this rock type, that are responsible for
nantly vertical within fluvial/lacustrine sequences the trapping of hydrocarbons beneath them.
due to the lack of lateral continuity of individual litho- • Where shallow-lacustrine facies are well devel-
facies. This may be due to either local or small-scale oped, fine quartz arenites are the effective seal.
unconformities, dewatering structures, or faulting.
The key to understanding the migration of hydrocar- The eastern part of the Jurassic/Cretaceous
bons within fluvial/lacustrine sequences is a detailed Eromanga Basin, together with the underlying Per-
study of facies geometry and the capillary pressures mian/Triassic Cooper Basin, comprises the largest
generated by key lithofacies involved. Also, water and most prolific onshore hydrocarbon province in
drives to otherwise confined oil pools may take the Australia (Hollingsworth, 1989). Production in 1994
same pathways as previous oil migration. An under- averaged 50,000 bbl of oil per day and 500 mmcf of
standing of such a water drive within a semiconfined gas per day (Hird, 1995). The Eromanga Basin con-
reservoir may significantly alter production strategies. tains common multiple stacked oil pools. Oil pools
In this paper we discuss capillary leakage of oil over sealed by the Callovian Birkhead Formation at the
geological time from several oil pools within the Ero- Bodalla South, Jackson, and Gidgealpa fields, and by
manga Basin, and the subsequent triggering of a prob- the Berriasian Murta Formation at the Murteree Ridge
lem water drive in a very thin reservoir on the fields, are indicated as being full to capillary seal
Murteree Ridge. capacity, and resultant vertical leakage has filled oil
The major geological hypotheses presented are that: pools higher up.
This paper presents some new data on seals
• Shales that occur within sand-rich fluvial/shal- within the Jurassic/Cretaceous Eromanga Basin and
low-lacustrine depositional environments are not briefly discusses migration pathways from the
effective seals due to their lack of lateral continu- underlying Permian/Triassic Cooper Basin into Ero-
ity. The sandstones that they are included in are manga Basin accumulations. It also discusses other
the effective seals, but only if their displacement associated eastern Australian basins, putting the
pressures are less than the reservoir and they Cooper/Eromanga Basin into an overall eastern
have sufficient thickness. Australian context. The locations of all the basins
•Effective seals within the fluvio/lacustrine discussed are shown in Figure 1. The Eromanga,
sequences of the epicratonic Eromanga Basin are Surat, and Carpentaria basins together are often
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 145

Carpentaria Mesozoic Basin Cooper/


Basin Late Paleozoic Age Ma Eromanga Surat
Basin
Winton Formation
Cenomanian

L
20
NT 500 km 100
Mackunda Formation
Griman Creek Formation
Albian
Wallumbilla Formation

Cretaceous
Surat Siltstone
Bowen 106 Coorikiana Sandstone
QLD Basin Aptian
Simpson Wallumbilla Formation
Desert EROMANGA 112 Wyandra Sandstone

Early
Pedirka Basin Barremian Member

Ridge
Basin BASIN 118
A Cooper SURAT Hauterivian
Cadna-owie Formation Bungil Formation
Basin 124
Nebine BASIN Valanginian
B 130
30
Berriasian Murta Formation
SA 136 McKinlay Member Mooga Sandstone
Tithonian Orallo Formation
Namur Sandstone
NSW Gubberamunda
SYDNEY

Late
146 Westbourne Sandstone
VIC Kimmeridgian Formation Westbourne Formation
140 150
150
Adori Sandstone Springbok Sandstone

Jurassic
Oxfordian
Figure 1. A location map of the Mesozoic, epicraton- 157 Birkhead Formation Walloon Coal Measures
ic Eromanga Basin, and foreland Surat Basin, show-
ing the underlying late Paleozoic basins in eastern Early-mid Eurombah Formation
Hutton Sandstone
Australia. Hutton Sandstone
Evergreen Formation
Poolowanna Formation
190 Precipice Sandstone
referred to as the Great Artesian Basin (Habermehl, Cuddapan Formation
1980) or the Great Australian Basin. The Eromanga Bowen Basin
Basin has very similar stratigraphy to the Surat Basin Sequence
Triassic

Early - Late Tinchoo Formation


to its east (Figure 2).

Arrabury Formation
EROMANGA BASIN: OVERVIEW OF
225
RESERVOIRS, SEALS, AND PETROLEUM Late Toolachee Formation
OCCURRENCE 240
Daralingie Beds
The major aquifers/reservoirs of Late Paleozoic to
Permian

Mesozoic basins of eastern Australia are quartz-rich Roseneath Shale


craton-derived (CD) fluvial sediments. Formations Epsilon Formation
containing significant reservoir potential within the Early
Eromanga Basin sequence are the Hutton, Adori, and Murteree Shale
Namur sandstones (Figure 2).
Patchawarra Formation
The uncommon occurrence of source rocks in the
Eromanga Basin has led many workers to attribute Tirrawarra Sandstone
most of the oil located within it to be sourced from the 280
underlying late Paleozoic, fluvio-lacustrine, source Pre-Permian
Pre-Permian

rock-rich Cooper Basin (Heath et al., 1989), as shown Basement


in Figure 3. The major regional seal to the Permian
section of the Cooper Basin is a thick sequence of Tri- Westphalian
310 SEALING
assic red beds, at the base of which is the Arrabury HORIZONS
Formation (Powis, 1989). This formation appears to
have had a profound effect on the distribution of
hydrocarbons in the overlying Eromanga Basin Figure 2. A stratigraphic correlation diagram for the
because where its thickness is >76 m (250 ft), very few Eromanga and Surat basins, northeastern Australia,
hydrocarbons have been discovered above it (Figure including detail of the Permian/Triassic Cooper
4). Many of the oil accumulations of the Eromanga Basin.
Basin occur in the vicinity of the underlying Cooper
Basin Permian and Triassic zero edges. Petroleum has
been produced from most reservoir horizons within Merrimelia Formation through to the Early Creta-
the Cooper/Eromanga Basin, from the Early Permian ceous Cadna-owie Formation (Figure 2). Fields within
146 Boult et al.

STACKED
OIL POOLS
JURASSIC
EROMANGA Arrabury Formation

141° E
100 km
BASIN SEALS zero edge
24° S
= Eromanga Basin
oil production

EROMANGA Permian
BASIN zero edge
Bodalla
South

TRIA Naccowlah 26° S


BASEMENT SSIC SEAL Moorari South

oil migration Jackson

South Australia
pathway COOPER BASIN Gidgealpa

Queensland
PERMIAN COOPER
aquifer BASIN SEALS source
flow

28° S
Figure 3. A simplified oil migration model for the
Cooper/Eromanga Basin. Much of the oil found
within the Eromanga Basin is thought to have Arrabury Formation
been sourced from the Cooper Basin via vertical approximate 250ft
migration near the Permian and Triassic zero edges. Limestone Creek, Dullingari isopach
Jena & Biala
Much of the oil that was originally trapped beneath
the Triassic and Permian Cooper Basin seals
has subsequently been flushed by gas. Oil Figure 4. A location map of fields in the Eromanga
predominantly occurs in the Eromanga Basin Basin showing the location of all gazetted oil
because most gas has been stripped by flowing discoveries, the underlying Permian and Triassic
groundwater in the aquifers (Heath et al., 1989). Arrabury Formation zero edges, and an ~250 ft
Arrabury Formation isopach.

the Eromanga Basin are characterized by multiple, margin to the southwest and an uplifted volcanic arc
vertically stacked pools contained within domes. Typ- to the northeast (Figure 7). The Birkhead Formation
ical oil columns are <20 m (65 ft), irrespective of ver- is the most extensive example of this repetitive
tical structural closure (Heath et al., 1989). The bulk influx of VAD material in the area. Neodymium
of producible hydrocarbons within the Eromanga model ages (Table 1a) have helped in correlation
Basin are crude oils with high API, pour points across the basin, with the VAD sediments giving
between –10° and 40°C, and low gas-to-oil ratios, consistently younger model ages than CD sediments.
typically 50 ft 3 /bbl. Stacked oil pools tend to be The integrity of the Birkhead Formation seal is
more geochemically related than laterally equivalent generally very good, with minimal faulting visible
oil fields. Heath et al. (1989) stated that this was on seismic at this horizon. A large unconformity
probably due to predominantly vertical migration. over the Nebine Ridge at the base of the Namur
They explained that the trend to higher API, lower Sandstone (Power and Devine, 1970) (Figure 5),
pour points, and fractionation of geochemical mark- which cuts down as far as mid-Birkhead Formation,
ers in shallower accumulations was due to the strip- is probably the most extreme example of repeated
ping of water-soluble components from the oils by movement that has occurred at a small scale many
groundwater flow. times in that and other ridge locations, within both
The effective regional seals (type D/E; Sneider the epicratonic Eromanga and foreland Surat basins.
and Stolper, 1991) within the fluvio/lacustrine Juras- Intra- and interbasin tectonics would also have had
sic section of the Eromanga Basin contain significant considerable influence on the distribution of the
quantities of labile, VAD rock fragments with abun- VAD material.
dant pore-filling authigenic clays (He and The interpreted shallow-lacustrine Murta and mar-
Conaghan, 1994). Major regional seals of this type in ginal marine Cadna-owie formations are also major
the Eromanga Basin are the Birkhead and West- seals within the Eromanga Basin. Net sand that does
bourne formations. During the Middle to Late Juras- occur within the Murta Formation has been docu-
sic, tectonic activity within a major volcanic arc to mented (Ambrose et al., 1986; Theologou, 1995) to
the east shed volcanogenic-rich, labile sandstone increase eastward, where large sublacustrine channels
wedges across the Mesozoic basins of northeastern have been interpreted (Zoellner, 1988); it is likely that
Australia (Figures 5, 6). The distance they travelled it is derived from the east. But neodymium model ages
depended, to some extent, on a balance between the (Table 1b) indicate that its provenance is not the same
supply of sediment from a cratonic and uplifted rifting as for the Birkhead Formation seal.
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 147

CRATON DERIVED VOLCANIC ARC DERIVED OTHER SEALS OF


SEDIMENTS/ SEDIMENTS (SEALS) NON VOLCANIC OR
POTENTIAL RESERVOIRS UNCERTAIN ORIGIN
Surat Basin

yyyyyyyyyyy
,,,,,,,,,,,
Eromanga Basin
HooraySandstone Wallumbilla Formation

West Orallo Formation East


Cretaceous marine sediments Gumberamund
a Sandstone

Sp
r in

,,,,,,,,,,,
yyyyyyyyyyy
tion
A ur Sandston e
Cad n
na-owie For m atio a For
M ur
t
ma gb
W o k Sst.
a ll
o on
Nam n n C o a l M e a s ures
atio io
W orm at
est b r n e F r m
A ou
Fo Eastern

,,,,,,,,,,,
yyyyyyyyyyy
Hu
or e ad Nebine Ridge o ne
d
iS t on he
tt o n
S a n d st Highlands
a n ds rk
Bi Ev
e r gre on
Po H e n Fo r m a t i
Pedirka o lo w
ma
tion tt o ne
s to
u

Basin a n na For n S a nd

,,,,,,,,,,,
yyyyyyyyyyy
Bowen Basin
Simpson Desert
Basin Cooper Basin
Pre-Permian Basement

Figure 5. A diagrammatic cross section of line AB (Figure 1) showing the interleaving of VAD (seal) and CD
(reservoir) sediments.

(a) Present Shoreline


(b)
Hutton Time Early Birkhead
Mesozoic Basin
Time
Vo

20 Low energy 20
l ca

500 km Birkhead/Hutton
ni
c
Ar
c

Braided Hutton

yyyy yyyy
,,,, ,,,,
J
G

30 30

,,,, yyyy
yyyy ,,,,
Labile sandstone
Sydney Braided wedge (mostly
Hutton low energy)
G = Gidgealpa
J = Jackson
Cratonic source Cratonic source
area 140
150
40
area 140 150

Figure 6. A paleogeography map showing (a) the probable drainage pattern at Hutton Sandstone time
and (b) the extent of the VAD labile sandstone wedge at the Hutton Sandstone/Birkhead Formation transition
to low-energy time.
148 Boult et al.

Low-energy
fluvial facies Vo
lc

,,yyy
yy,,, ,,,
yyy
High-energy an
ic
braided fluvial Ar
c
facies

s
hland n
Hig Subductio

,,,yy
yyy
,,,,,
yyyyy ,,
rn
u the
So

Rifti Oceanic crust


ng o
f An Volcanic arc derived

,,,,,yy
yyyyy ,,
tarct
ica lithic arenites.
Birkhead
Formation
N Craton derived
quartz arenites.
WESTBOURNE TIME
Pre-Jurassic

Figure 7. A 3-D diagram of the paleogeography during the deposition of the Westbourne Formation, showing
the interleaving of VAD seals and CD reservoir-type sediments.

A marine transgression occurred at the beginning of The Hutton Sandstone/Birkhead Formation


the Albian (106 m.y.), resulting in the deposition of the Reservoir Seal Couplet
offshore marine shales of the Wallumbilla Formation
and equivalents. This forms the youngest and most The Hutton Sandstone and lower Birkhead Forma-
widespread seal (Sneider seal type A/B) of all the east- tion produce the bulk of the oil within the Eromanga
ern Australian basins and, to date, no hydrocarbon accu- Basin; the main seal occurs within the Birkhead For-
mulations have been discovered above this horizon. mation. Production mainly comes from domal struc-
tures. Variable stratigraphic influence on reservoir
geometry and quality is caused by the transitional
CASE STUDIES nature of the Hutton Sandstone to Birkhead Formation
contact. The Hutton Sandstone is a CD braided fluvial
Specific new data are presented on the Jackson, deposit and is regionally consistent in terms of its
Gidgealpa, and Moorari fields of the Eromanga Basin. reservoir quality. The sediments of the Birkhead For-
Bodalla South and Limestone Creek/Biala case studies mation were deposited within a lower-energy mean-
have been published by Boult (1993) and Williams et dering fluvial, swampy to lacustrine environment.
al. (1994), respectively, and will be discussed briefly, Table 3 and Figure 8 display a comparison of the aren-
except where new data have become available since ites discussed in this section. While the Birkhead For-
publication. A summary of the findings of mercury mation is generally conformable over the Hutton
injection capillary pressure (MICP) studies on individ- Sandstone, some erosion of the Hutton Sandstone by
ual fields in this paper is presented in Table 2. In order “Birkhead” channels has been interpreted at the Jack-
to display the sensitivity of the calculated oil column son, Bodalla South (Boult and Theologou, 1993), and
heights to controlling parameters, the maximum and Naccowlah South (Hill, 1985) oil fields, indicating that
minimum values for each parameter are used. From small-scale tectonic uplift and incision of meanders
these, a maximum and minimum oil column height has occurred in the east of the basin toward the Nebine
has been calculated. The development of the seal Ridge (Figures 1, 5).
capacity equation and the determination of displace- The transition from Hutton Sandstone to Birkhead
ment pressure are presented in the Appendix. It Formation is accompanied by a provenance switch
should, however, be noted that minimum values from quartz arenite to a lithic arenite (Watts, 1987).
would be more likely to occur in nature because: (a) a Petrological data (Green and Thomas, 1993; McPhie et
migrating oil column will always take the path of least al., 1993) indicate that the lithic arenite is a sediment
resistance, “weak link,” or lowest displacement pres- with a significant volcanogenic content at Gidgealpa.
sure accessible in any sequence; and (b) the laboratory This may be the same as the one documented by Watts
interfacial tension (Core Lab, 1982) value of 30 (1987) and Whitford et al. (1994) at Bodalla South;
dynes/cm is based on measurements in the United thus, to our knowledge, Gidgealpa represents the fur-
States, and Eromanga Basin oils may contain natural thest west this VAD sediment has been documented.
surfactants. An investigation using initial neodymium-isotopic
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 149

Table 1a. Sm and Nd Isotope Results and Calculated Model Ages for the Hutton Sandstone/Birkhead
Formation Transition in Gidgealpa 32.*

Formation

VAD VAD CD CD CD CD
Birkhead Birkhead Birkhead Birkhead Hutton Hutton
Fm. Fm. Fm. Fm. Sst. Sst.
Sample type Shale Sand Sand Shale Shale Sand
Depth (ft) 5974.5 5989 5997 6005 6012 6020
Nd (ppm) 31.5 35 30.47 41.15 43.55 9.6
Sm (ppm) 6.37 6.95 5.43 9.48 8.26 1.81
143Nd/144Nd 0.512521 0.512502 0.511915 0.51238 0.512019 0.51177
Epsilon Nd –2.3 –2.7 –14.1 –5 –12.1 –16.9
Sm/Nd 0.2022 0.1987 0.1782 0.2304 0.1897 0.188
147Sm/144Nd 0.1223 0.1202 0.1078 0.1394 0.1148 0.1137
Initial 143Nd/144Nd 0.512393 0.512376 0.511802 0.512234 0.511899 0.511651
Initial epsilon Nd –0.8 –1.1 –12.3 –3.9 –10.4 –15.2
Model age (Ga) 0.95 0.96 1.62 1.39 1.58 1.9

*All analyses performed at the University of Adelaide. Model age calculations based on 143Nd/144Nd of present-day depleted mantle =
0.513108, and 146Sm/144Nd = 0.2157. Present-day CHUR = 0.512638, with 147Sm/144Nd = 0.1967. Initial ratios and epsilon values calculated at
0.16 Ma. Analytical techniques are like those described in Foden et al. (1995).

TABLE 1b. Sm and Nd Isotope Results and Calculated Model Ages for the Namur Sandstone and Murta
Formation in the Dullingari Field.*

Formation

Namur Sst.* Murta Fm.* Murta Fm.* Murta Fm.*


Sample type Sand Shale Sand Silt/Sand
Well no. D29 D46 D46 D46
Depth (ft) 4959 4828.8 4828.9 4842
Nd (ppm) 3.69 43.43 8.03 32.03
Sm (ppm) 0.68 7.87 1.82 5.96
143Nd/144Nd 0.511796 0.511859 0.511845 0.511796
Epsilon Nd –16.4 –15.2 –15.5 –16.4
Sm/Nd 0.1844 0.1811 0.2266 0.186
147Sm/144Nd 0.1115 0.1095 0.1371 0.1125
Initial 143Nd/144Nd 0.511679 0.511744 0.511701 0.511678
Initial epsilon Nd –14.7 –13.4 –14.3 –14.7
Model age (Ga) 1.83 1.72 2.27 1.85

*All analyses performed at the University of Adelaide . Model age calculations based on 143Nd/144Nd of present-day depleted mantle =
0.513108, and 147Sm/144Nd = 0.2157. Present-day CHUR = 0.512638, with 147Sm/144Nd = 0.1967. Initial ratios and epsilon values calculated at
0.16 Ma. Analytical techniques are like those described in Foden et al. (1995).

compositions was able to discriminate separate distinc- the high- or low-energy quartz arenite sediments.
tive sediment provenances. At Bodalla South, Whitford Neodymium model ages in clastic sediments date
et al. (1994) assigned neodymium model ages of the mean separation of crustal material from the
800–1000 m.y. to the VAD arenite and 1100–1600 m.y. mantle. The stratigraphic age of the volcanic arc is
to the CD arenite, reflecting a fundamental change in probably much younger than the model age
provenance with an influx of relatively younger crustal assigned to the VAD arenite of the Birkhead Forma-
material derived from the northeast. At Gidgealpa, tion because of a dilution effect in transit from the
neodymium model ages vs. depth are shown in Figure inclusion of older crustal, including older CD, sedi-
9, with the lithic arenite having a value of 950 m.y. and ments (Miller et al., 1986).
the underlying quartz arenite sediments between 1400 The provenance change and energy change at the
and 1900 m.y. The postprovenance-change lithic aren- Hutton Sandstone/Birkhead Formation transition
ite has a considerably younger model age than either are rarely synchronous, and both are probably
150 Boult et al.

Table 2. Comparison of Actual and Calculated Oil Column Heights with Controlling Parameters.

Field
Bodalla Sth Jackson Gidgealpa Moorari Limestone Ck.
Formation Birkhead Birkhead Birkhead Birkhead Murta
Oil Column Height Sensitivity Max. Min. Max. Min. Max. Min. Max. Min. Max. Min.
Calculated oil density (gm/cc)* 0.75 0.75 0.77 0.77 0.77 0.77 0.78 0.78 0.76 0.76
Calculated water density (gm/cc)* 0.97 0.97 0.97 0.97 0.97 0.97 0.96 0.96 1 1
Oil/water interfacial 11.3 7 21.9 21.9 30 21.9 30 21.9 30 21.9
tension (dynes)**
Seal mercury capillary 200 150 250 200 150 100 300 150 125 90
displacement pressure (psi)†
Reservoir mercury capillary 0 5 0 5 0 5 0 5 0 5
displacement pressure (psi)††
Calculated oil column height (ft)‡ 64 29 171 134 140 65 311 110 99 49
Calculated oil column height (m)‡ 19.6 8.8 52.2 40.7 42.7 19.7 94.9 33.5 30 14.9
Actual oil column (m)‡‡ 11.2 36 20 13.7 17
Structural closure height (m)§ 40 30 50 36 53 45 >60 >60 58 34

* Oil and water densities are calculated using standard reservoir engineering principles.
** A value of 21.9 dynes/cm was obtained using the pendant drop method at reservoir conditions on oil from the Jackson field and a syn-
thetic brine. This is used in the minimum case for all other fields except Bodalla South. The values for Bodalla South were obtained using the
Wilhemy plate method on oil and formation brine at laboratory pressure for varying temperatures. Obtained values were then extrapolated to
reservoir temperature for the minimum case and the measurement at the highest temperature, 37°C, is used for the maximum case. Thirty
dynes/cm is used in the maximum case for all other fields where laboratory measurements have not been conducted. It is a commonly used
value by the industry in the absence of any other data. The seal is assumed to be totally water-wet with a contact angle of zero for all cases.
†The displacement pressure for each MICP curve is calculated by extrapolating the curve plateaus back to the pressure axis and taking into

account a closure correction.


††The reservoir mercury displacement pressures in many cases are not measurable and seldom above 5 psi.
‡The interfacial tension and contact angle of mercury are assumed to be 484 dynes/cm and 140° API, respectively (Rootare, 1970)
‡‡Actual oil column heights are the maximum determined from log analysis and may be slightly low if the well used is not at the top of structure.
§Structural closure takes into account any seismic uncertainty due to velocity variations.

diachronous. There is thus some confusion in the liter- The Bodalla South Oil Field
ature as to exactly where the base of the Birkhead For- The Bodalla South oil field (Figure 4) was discov-
mation is. In some areas, the transition from Hutton ered in 1984 (Dolan et al., 1988) and comprises two
Sandstone to Birkhead Formation is predominantly a main reservoirs: the Hutton Sandstone and the Basal
provenance change, with similar-size “Birkhead” Jurassic Sandstone. It is primarily a structural trap
channels replacing “Hutton” channels. In other areas, occurring in the northeast of the Eromanga Basin near
the Hutton Sandstone is conformably overlain by the edge of the underlying Triassic regional seal, the
lower-energy overbank siltstone and shales. For exam- Arrabury Formation. Original oil in place for the Hut-
ple, (Figure 10) at Gidgealpa (McIntyre et al., 1989) ton Sandstone reservoir was proven at 0.47 × 106 m3
and Moorari (Ties, 1983, personal communication), the (3 million bbl). Structural closure at Bodalla South is
boundary is taken as the change between high- and estimated to be 30–40 m (100–130 ft). The maximum oil
low-energy deposits, and the provenance change column encountered in the field is 11.2 m (37 ft) (Table
occurs above this (Figure 10). At Jackson (Dodman 2), which indicates that the reservoir is not full to its
and Rodriques, 1989), the boundary between higher- structural spillpoint.
and lower-energy channels is used to divide between The weak link in the seal, or “automatic valve”
the lower and upper Hutton Sandstone reservoirs; the (Schowalter, 1979), is a VAD, lithic-rich arenite, at the
boundary between the Hutton Sandstone and the base of the Birkhead Formation, which has a sharp
Birkhead Formation is taken as the provenance change contact with underlying CD quartz-rich Hutton Sand-
where recognized. In the Surat Basin, the transition is stone. Analysis of MICP curves of the VAD lithic aren-
known as the separate Eurombah Formation. In this ite and underlying CD quartz-rich reservoir, which
paper we do not intend to redefine the base of the took into account subsurface fluid parameters (Table
Birkhead Formation, although it is recognized that the 2), suggests that the maximum size of oil column that
problem should be addressed. will accumulate beneath this cap rock is 8.8–19.6 m
Of the fields studied and described below, Jackson (29–64 ft). The minimum value is similar to the maxi-
and Bodalla South are more proximal to the volcanic mum oil column encountered in the field at 11.2 m
source area. The Gidgealpa and Moorari fields repre- (37 ft) in Bodalla South 2. The VAD lithic arenite that
sent a more distal position, with Moorari being closer occurs at the base of the Birkhead Formation therefore
to the Eromanga Basin depocenter than Gidgealpa. appears to be a good candidate for a leaky seal and an
1 2

(A)

(B)

(C)

(D)

Figure 8. Thin-section photomicrographs of the CD coarse-grained arenite (A1 and A2), CD fine-grained arenite
(B1 and B2), and VAD arenites (C1, C2, D1, and D2). Photomicrographs on the right were taken by using plane-
polarized light and on the left by using crossed-polarized light. (A1 and A2 field of view (FOV) 4 mm): Gidgealpa
32, 6019.3 ft CD, coarse-grained, well-sorted Hutton Sandstone showing the predominance of monocrystalline
quartz crystals (q) with uniaxial extinction and syntaxial quartz overgrowths (o). Occasional pore-blocking kaolin-
ite (k) secondary porosity (s), caused by grain dissolution, and possible metamorphic rock fragments (m) are evi-
dent. (B1 and B2 FOV 2 mm): Gidgealpa 32, 5990.6 ft. CD, fine-grained, well-sorted lower Birkhead Formation
showing quartz and mica grains enclosed authigenic clay (illite?) (C1 FOV 2 mm; C2 FOV 1 mm ): Gidgealpa 32,
5969.3 ft. VAD, postprovenance-change Birkhead Formation. Medium-grained, poorly sorted lithic arenite/wacke-
stone containing common grains that display volcanic textures. Many grains have a possible micropoikilitic,
devitrification texture that is commonly seen in acid volcanics. This consists of a patchy or zoned quartz ground
mass enclosing feldspar crystals (McPhie et al., 1993). The grain at the center of the picture shows two separate
zones (a and b) of optically continuous quartz that go into extinction at different orientations. Also shown are
grains of biotite (bi), yellow-stained potassium feldspar (Fk), quartz (q) with secondary calcite (c), pore-filling
kaolinite (k), and abundant authigenic quartz occurring as fine-grained aggregates (aq) and syntaxial quartz over-
growths (o). (D1 FOV 2 mm and D2 FOV 1 mm): Jackson 40, 4760 ft 6 in. VAD, postprovenance-change Birkhead
Formation. Medium-grained, poorly sorted lithic arenite/wackestone with highly altered sericitic grains (S), very
fine grained authigenic quartz (aq), pore-filling kaolinite (k), and calcite (c). In the center of the photo are two
basalt grains (b) showing plagioclase laths enclosed in opaque to semi-opaque basaltic glass. The quartz grain (q)
content of this rock is very low, <30%.
152 Boult et al.

DEPTH MODEL AGE (m.y.) explanation for the significant shows encountered
(ft)
600 800 1000 1200 1400 1600 1800 2000
higher up in the section (see Boult, 1993).
5970
LITHIC ARENITE The Jackson Oil Field
The Jackson oil field in the southwestern Queens-
5975 land section of the Eromanga Basin contained about
111 million bbl in three reservoirs at the time of dis-
5980 covery in 1981 (Dodman and Rodriques, 1989): the
Hutton Sandstone with 14.8 × 10 6 m 3 (94.5 million
bbl), the Westbourne Formation with 2.4 × 10 6 m 3
5985 (14.6 million bbl), and the Murta Formation with 0.3 ×
LOW ENERGY

106 m3 (2 million bbl). It is located near the Triassic


PROVENANCE CHANGE Arrabury Formation subcrop limit (Figure 4), and
5990 charge is thought to have come from the Cooper
Basin depocenter to the west (Heath et al., 1989).
5995 A schematic cross section of the Jackson oil field
is presented in Figure 11, and reservoir boundaries
are shown in Figure 12. The lower Hutton Sand-
6000 stone is interpreted to have a mainly vertical water
QUARTZ ARENITE

drive, with coning occurring within two months of


production and requiring the installation of electric
6005
submersible pumps on seven wells. The upper Hut-
ton Sandstone is generally less permeable and more
6010 heterogeneous than the lower Hutton Sandstone.
HIGH ENERGY

The upper Hutton Sandstone shows evidence of an


edge water drive and is not in vertical pressure
6015 communication with the lower Hutton Sandstone
reservoir.
The Jackson top Hutton Sandstone horizon has a
6020
structural closure of ≤50 m (175 ft) (Altman, 1993, per-
Figure 9. Neodymium model ages vs. depth for sonal communication), and the maximum oil column
Gidgealpa 32. The lithic arenite (VAD sediment) encountered is 36 m (120 ft). A cross section of the Hut-
shows significantly lower model ages than the ton Sandstone (Figure 13) shows that the Birkhead For-
quartz arenite (CD sediment), indicating a probable mation is stratigraphically low in Jackson 40 by 11 m
provenance change. (36 ft). A core in Jackson 40 revealed a massive VAD,

Figure 10. The


SOUTH Hutton Sandstone/
QUEENSLAND Birkhead Formation
AUSTRALIA
transition detail
Jackson / Bodalla from across the
Gidgealpa Moorari Naccowlah South Eromanga Basin,
BIRKHEAD FORMATION

showing the varia-


tion in thickness of
the “low-energy
shallow
lacustrine channels
facies” beneath the
provenance change
PROVENANCE between VAD and
CHANGE
meandering
meandering
fluvial ENERGY
CD sediments.
fluvial CHANGE

meandering referred to
braided fluvial as upper
Hutton in
fluvial
SANDSTONE

braided Queensland
braided fluvial section of the
fluvial Eromanga Basin
HUTTON

braided
fluvial
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 153

1 4 2° 2 6 '
depth
subsea ft Sealing
horizons
-3000 NORTH SOUTH
1 km
o

MURTA FORMATION o
-3500 o oo
N
NAMUR SANDSTONE

UPPER WESTBOURNE FORMATION o o


o

-4000
LOWER WESTBOURNE FORMATION 27° 36' A2 1 o
o 27° 36'
BIRKHEAD FORMATION
o o
40 o
HUTTON SANDSTONE 3
-4500
15 o o
Figure 11. A simplified schematic cross section of o
the Jackson field (modified from Dodman and o oo o
13
Rodriques, 1989). o
o
lithic arenite channel (Figure 8D), with an average per- o
o
meability of <0.1 md. o
The lithic arenites at Jackson are very similar to
those at Bodalla South (Boult, 1993) and represent a 21
westward extension of the postprovenance-change
channels recognized there. A correlation based on Bo
o
wireline logs, cuttings, and production data suggests
that the effective seal at Jackson is an overbank facies
that is equivalent to the top of the lithic arenite channel o
cored in Jackson 40. This unit is ~15–18 m (50–60 ft)
thick across the field. It also extends as far as Nac- o
cowlah South oil field (Figure 4), where it has been
cored in Naccowlah South 4. 2 7° 3 8 ' 2 7° 3 8 '
Analysis of MICP curves of the lithic arenite from
the channel, which took into account subsurface fluid
parameters (Table 2), has revealed that the maximum
size of oil column that will accumulate beneath this o
cap rock is 40.7–52.2 m (134–171 ft) (Boult and Theolo- Murta Formation
gou, 1993). These values are similar to the maximum
oil column encountered in the field. They are also sim- Westbourne Formation
ilar to the structural closure of the field. The seal at
Jackson is a good candidate for consideration for being Hutton Sandstone
1 4 2° 2 6 '

leaky, given that the minimum calculated oil column


is less than the maximum and most likely structural
closure, and that chemically similar oil (Heath et al.,
1989) occurs higher up.
Above the overbank deposits of VAD lithic arenites Figure 12. Jackson field oil pool boundaries,
and siltstones, porosity occurs within the Birkhead showing the line of section for Figure 13 (modified
Formation, either as a result of secondary develop- from Dodman and Rodriques, 1989).
ment or, more likely, as a result of decrease in labile
grain content due to incorporation of more CD grains.
Water is the only indicated pore fluid in these sands, trace of oil was recovered. This could be close to the oil
except for Jackson 3, which is near the crest of the migration conduit leaking from the CD Hutton Sand-
structure. This horizon in Jackson 3 was tested, and a stone underneath. If this location does represent the
154 Boult et al.

JACKSON FIELD 21 40 3 15 13 2 0 WELL


380 m 350 m 320 m 400 m 920 m

35 km
A B
NACCOWLAH

Depths
subsea
feet
Westbourne Formation
SOUTH 4

4150

4250
Birkhead Formation
Lithic

pro
channel

ve
Upper

na
Hutton

nc
Oil Pools Lower

e
4350 Hutton

en

ch
Oil Pool

erg

an
ge
y

ch
an
ge
4400
Hutton Sandstone
CRATON-DERIVED VOLCANIC-ARC-DERIVED CORE
OIL POOLS
SEDIMENT SEDIMENT (SEAL) POSITIONS

Figure 13. Structural cross section across the Jackson field Hutton Sandstone oil pools, showing provenance
change (top dashed line) between CD and VAD sediments; energy change (bottom dashed line) between
braided fluvial and meandering fluvial/lacustrine depositional environments; and oil pools. A correlation
with Naccowlah South is also included (see Figures 4 and 12 for locations).

site of a leakage conduit, then the reason that oil is not two domes. Gas is found only within the Permian
pooling downflank must be because of the lack of con- Cooper Basin, which onlaps onto a basement high,
nectivity between intra-Birkhead Formation sand bod- with the major accumulation being 7070 × 106 m3 (250
ies that are diluted with CD quartz. This is as expected billion ft3), occurring in the Toolachee Formation of
within a meandering fluviatile system. There are also the north dome. This is sealed beneath the Arrabury
common thinly bedded shales and siltstones within Formation at the base of the Triassic. A thin, 4-m (12-ft)
the Birkhead Formation. These are not considered lat- oil leg occurs beneath the gas accumulation in the Tir-
erally continuous enough to be effective as fieldwide rawarra Sandstone of the southern dome.
seals, but they may have local effects. The Triassic seal has been eroded away or was not
deposited over part of the southern dome; the gas
The Gidgealpa Oil and Gas Field accumulation in the Toolachee Formation is limited
The first gas in the Cooper/Eromanga Basin was here by a stratigraphic spillpoint (Figure 14). Both oil
discovered in Gidgealpa 2 in 1963 (Figure 4). The field and gas have probably migrated up into the overlying
produces gas from the Permian Tirrawarra Sandstone, Eromanga Basin above the southern dome, but
Patchawarra and Toolachee formations of the Cooper because no gas occurs in the Jurassic at Gidgealpa and
Basin, and oil from the Poolowanna Formation, Hut- the oils have very low gas-to-oil ratios, it is assumed
ton Sandstone, lower Birkhead Formation, and Namur that water-washing has stripped all the gas away.
Sandstone in the Jurassic of the Eromanga Basin Most of the oil that leaked up into the Jurassic at the
(McIntyre et al., 1989). A schematic cross section is crest of the southern dome has tended to migrate ver-
shown in Figure 14, and reservoir boundaries are tically; 1.7 × 106 m3 (10.9 million bbl) of oil have been
shown in Figure 15. The hydrocarbons are located in trapped in the Hutton Sandstone and lower Birkhead
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 155

SW Wallumbilla Fm (marine)
Cadna-owie Fm NE
Depth Murta Fm
subsea
feet
-5000 Namur Sst

SEALS
Birkhead Fm

Hutton Sst
-6000

yyyyyyyyyy
,,,,,,,,,,
c S e al
Poolowanna Fm a s s i
T r i l a c h e e F m
T o o

,,,,,,,,,,
yyyyyyyyyy
-7000

Patchawarra Sst
Tirrawarra Sst 25
3Km PRE-PERMIAN
gas migration path oil migration path GAS OIL PERMIAN

CRATON-DERIVED VOLCANIC-ARC-DERIVED OTHER SEALS OF


SEDIMENTS (SEALS) NON VOLCANIC OR
SEDIMENTS UNCERTAIN ORIGIN

Figure 14. A schematic cross section of the Gidgealpa field, showing possible migration pathways and the con-
trol of the Triassic seal on migration pathways.

Formation by a VAD lithic arenite capillary seal within (1) Hutton Sandstone reservoir: Base—1830.1 m
the Birkhead Formation. About 0.6 × 106 m3 (4.0 mil- (6006 ft) (Figure 8 A1, A2). Braided fluvial, quartz
lion bbl) of oil has leaked past the Birkhead Formation arenite in which most of the labile minerals such
up into the Namur Sandstone, where it is trapped as feldspar have been dissolved, creating sec-
beneath a local mid-Namur Sandstone seal and at the ondary porosity.
base of the Murta Formation. (2) Lower Birkhead Formation reservoir: 1830.1–
The structural closure at the top Hutton Sandstone 1824.9 m (6006–5989 ft) (Figure 8 B1, B2). Mean-
horizon within the southern dome has been mapped to be dering fluvial, quartz arenite. This reservoir is
at least 53 m (175 ft) (Singh, 1990). The maximum oil col- very poorly developed in Gidgealpa 32. Bioturba-
umn encountered to date is 20 m (66 ft) in Gidgealpa 25. tion in the form of sand-filled burrows is com-
Two correlatable cores have been taken through the mon, and there is evidence of rapid deposition
base of the Birkhead Formation and into the Hutton with the preservation of dewatering structures.
Sandstone in Gidgealpa 32 and 26. The arenites from At the top of the lower Birkhead Formation reser-
these have been extensively sampled. Thin-section and voir, a soil with recognizable “a,” “b,” and “c”
MICP data from them are presented in Figure 16 and horizons developed. This shows there was a hia-
plotted against the wireline depth for Gidgealpa 32. tus in deposition before the Birkhead Formation
None of the shales or siltstones within this sequence seal was deposited.
are thicker than half a meter (1.5 ft). Many of the aren- (3) Birkhead Formation seal: 1824.9 m (5989 ft)—top
ites have erosive bases, and the shales contain arenite- (Figure 8 C1, C2). Meandering fluvial lithic aren-
filled defects in the form of locally abundant ite to wackestone. The Birkhead Formation seal
bioturbation and dewatering structures. The shales are represents a continuation of the same facies as
therefore not likely to be effective seals and have not the lower Birkhead Formation reservoir, but
been included in the diagram. The arenites in this sec- with a different provenance. Sedimentation rates
tion can be divided into three types, as described in were often rapid, with ripple-bedded sands
Table 3 and Figure 8. comprising ≤60% of the facies. Throughout the
156 Boult et al.

140°00'E
Permian Toolachee Fm
gas Patchawarra Fm
Tirrawarra Sst
NE

Namur Sst
Jurassic Birkhead Fm
oil Hutton Sst
Poolowanna Fm

Jurassic ,,,,
,,,, Namur Sst
H

,,,,
oil over
Birkhead Fm

,,,,
Permian
gas
H
North
,,,, ,,,
Hutton Sst

,,,
Poolowanna Fm H

,,, 90
0
-6
0
50
-7
0
30
-7 0
10
-7
00
-69
B 28° 00'S

,,,,,
TRDG

53
E

,,,,,
IA E

,,,,,
subsea contour
SS

interval 200ft at top


,,,,,
32
I C

o Toolachee Fm
,,,,,,,
,,,,,
o
o20 o
,,,
,,,
,,,

,,,,,,,
,,,,,,,,,, o

,,,,,,,,,,
o
27
30

A
,,,,,
,,,,, H

,,,,,
,,,,,
5 km
SW
Figure 15. An oil pool map of the Gidgealpa field showing (1) the location of the Triassic zero edge and its
control on the Toolachee Formation gas accumulation, which is only present to the northeast, and (2) the
Triassic zero edge’s control on stacked Jurassic oil pools, which only occur within the southern dome. Cross
section AB shown in Figure 18.

Birkhead Formation seal sequence the dominant extensive enough to form an effective barrier. The dis-
lithology is a very lithic rich arenite with abun- placement pressures from all samples are displayed in
dant recognizable grains of fine-grained vol- Figure 16; it is thought that the interval between 1816.7
canics (Figure 8). Some of these grains have and 1822.8 m (5961.3–5982.0 ft) would be the effective
been subjected to considerable plastic deforma- seal. Analysis of displacement pressures displayed in
tion, especially where the quartz grain content Figure 16 indicates that the oil column retainable by the
falls below a critical percentage, allowing local- effective seal at Gidgealpa is 19.7–42.7 m (65–140 ft)
ized framework collapse (Harris, 1989). A col- (Table 2). The minimum and most probable figure is in
lapsed lithic arenite occurs at the base of the close agreement with the actual oil column of 20 m
Birkhead Formation seal in Gidgealpa 32 (66 ft) that occurs beneath the Birkhead Formation seal
between 1823.8–1824.9 m (5985.5–5989 ft). in Gidgealpa 25, which is near the crest of the struc-
ture. The maximum figure is still less than structural
It is considered by the authors that it is the non- closure.
collapsed VAD lithic arenite that is the “weak link” or Gravestock et al. (1983) noted the common occur-
effective capillary seal at Gidgealpa. Neither the shales rence of calcite within some aquifers of the Eromanga
nor the collapsed lithic arenite would be laterally Basin. Singh (1990) mapped calcite cemented intervals
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 157

Figure 16. Gidgealpa


field, Hutton
Sandstone/Birkhead
Formation mineral-
ogy and displace-
ment pressure log
showing the location
of the Birkhead
Formation seal:
RF = rock fragments
including common,
fine-grained vol-
canic lavas, Clays =
clays that cannot be
identified in thin
section, F-or =
orthoclase feldspar,
F-ab = plagioclase
feldspar with albite
twinning, S-or =
secondary/dissolu-
tion/intragranular
porosity, kaol =
kaolinite or dickite
occurring either as
intergranular pore
filling or grain
replacement,
org = organic matter,
opq = opaques,
Hv = heavy
nonopaque
minerals, mica =
muscovite and
biotite, Qtz = quartz
grains and
overgrowths,
IG-por = primary
or intergranular
porosity, Calc =
predominantly
calcite with minor
siderite, and
Hg DP = mercury
displacement
pressure.
158 Boult et al.

Table 3. Comparison of Arenite Lithologies of the Hutton Sandstone/Birkhead Formation Transition.

Hutton Sandstone/
Lower Birkhead Birkhead Formation
Hutton Sandstone Formation Facies (Post-Provenance Change)
Provenance CD CD VAD
Energy/Facies High Medium to low Medium to low
braided fluvial meandering fluvial meandering fluvial
Composition quartz arenite quartz to subarkosic lithic arenite to lithic
(see Figure 16) arenite wackestone with
abundant recognizable
grains of fine-grained
volcanics
Grain size medium to coarse fine to medium fine to medium
with occasional
pebbles
Sorting medium to good medium medium to poor
Grain shape rounded subangular subangular
Common accessories – muscovite authigenic clays and
calcite
Primary porosity (%) 15–25 10–20 0–5
Secondary porosity 2–10 2–10 2–15
Permeability (md) 500–3000 2–100 <0.2
Hg pore displacement <1–5 5–15 100–250
Pressure (psi)
Photomicrograph A1 and A2 B1 and B2 C1, C2, D1, and D2

within the Namur Sandstone over the Gidgealpa


1 3 9 ° 58'E

1 4 0 ° 00'E
south dome (Figure 17), and several workers (Jensen- 2 8 ° 00'S 75 thickness of 2 8 ° 00'S
-59 calcite, contour
Schmidt, 1989; Schulz-Rojahn, 1993; Townsend 1993) interval = 25 ft
0
have since noted poikilotopic calcite cement within the -6
00 95
0
-5
Namur Sandstone concentrates near the crest of other depth contour
• 31 interval = 25ft
oil-bearing structures. Schulz-Rojahn (1993) pointed
out that the Eromanga Basin is virtually devoid of evi-
-59

• 28
•32 •
25

dence for the microbial alteration of hydrocarbons,


which is thought to have triggered carbonate precipi- • 29
• • 20
tation in other, cooler petroleum-bearing sequences

(e.g., O’Brien and Woods, 1995; Hovland et al., 1987). •5 22
+
•21 •
Schulz-Rojahn (1993) and Ryan-Grigor and Schulz-
-59

• 27
Rojahn (1995, and personal communication) high- • 31
00

• 23
10

lighted the probable role of carbon dioxide in


5
12

2 8 ° 02'S + 2 8 ° 02'S
controlling carbonate cement precipitation near the
crest of petroleum fields. In the Eromanga Basin, avail- • 34
• 16
able geological and geochemical data suggest that the • 33
occurrence of calcite-cemented zones within the +
Namur Sandstone over the Gidgealpa south dome
may be related to chemical reactions involving migra-
1 3 9 ° 58'E

1 4 0 ° 00'E

tion of carbon dioxide from deeper sequences. Carbon 2km


dioxide is thought to be a precursor to oil maturation
(Momper, 1980) and migration, and oil often contains
some dissolved carbon dioxide. Figure 17. Calcite-cemented zone isopachs for the
Both the spatial and statistical relationships Namur Sandstone and top Hutton Sandstone
between calcite-cemented zones in the Namur Sand- (depths in feet) for the Gidgealpa south dome
stone and both over- and underlying traps (Jensen- (Singh, 1990). This shows the focussing of the possi-
Schmidt, 1989) that generally contain hydrocarbons ble migration pathway by the Birkhead Formation
are sufficient to view the cements as migration- seal and the resultant positioning of the MRDZ off
related diagenetic zones (MRDZs). The term differs structure.
from the commonly used “hydrocarbon-related dia-
genetic zone” (HRDZ) because the Eromanga Basin
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 159

27 30 20 32 25 5 3 WELL
A
Depths
750 m 1750 m 500 m 700m 1200 m

B
subsea GIDGEALPA SOUTH DOME
feet Adori
5700 Sandstone

barren
lower
5800 Birkhead Birkhead
oil pool

Birkhead Formation
ch
an
ge
ce

5900
an
en

ch

?
ov

an
pr

ge
y
erg
en

6000
Hutton Oil Pools

Hutton Sandstone
CRATON-DERIVED VOLCANIC-ARC-DERIVED CORE
OIL POOLS POSITIONS
SEDIMENT SEDIMENT

Figure 18. A structural cross section across the Gidgealpa field showing provenance change (top dotted line)
between CD and VAD sediments; energy change (bottom dotted line) between braided fluvial and meander-
ing fluvial/lacustrine depositional environments; and oil pools.

calcites do not strictly represent alteration products flank of the field. Production pressure data in combi-
of hydrocarbons. nation with petrological core data and wireline data
Thus, the MRDZ within the Namur Sandstone, the have been used to create a correlation across the field,
0.6 × 106 m3 (4.0 million bbl) that occur above the Birk- which is presented in Figure 18. This correlation of the
head Formation of a very similar oil to that in the lower Birkhead Formation reservoir is similar to the
reservoir below it, and the abundant shows that are Upper Hutton Sandstone reservoir at Jackson, with a
evident within the VAD lithic arenite of the Birkhead change to meandering fluvial from braided fluvial fol-
Formation, indicate that leakage of carbon dioxide and lowed by a provenance change. We suggest that the
oil has probably occurred or is occurring through lower Birkhead Formation reservoir at Gidgealpa is
these sediments as they act as an “automatic valve.” probably equivalent to the upper Hutton Sandstone
The Hutton Sandstone reservoir at Gidgealpa has a reservoir at Jackson.
strong bottom-water drive. The lower Birkhead Forma-
tion reservoir has limited pressure communication The Moorari Oil and Gas Field
with the Hutton Sandstone reservoir, which is primar- The Moorari field is an oil- and gas-producing field,
ily due to vertical permeability destruction caused by with the main reserves in the Permian Tirrawarra
intervening siltstones. It is interpreted to have an edge Sandstone, near the base of the Cooper Basin. When
water drive, because it is in partial communication discovered, the Tirrawarra Sandstone contained 2.31 ×
with the Hutton Sandstone reservoir on the western 106 m3 (14.7 million bbl) of oil and 1070 × 106 m3 of gas
160 Boult et al.

seal is ~76 m (250 ft) thick, thus limiting connection to


0 GR GAPI 200 120 DT US/F 60
the underlying Cooper Basin hydrocarbon charge. The
70-9 Sand is at least 30 m (100 ft) above the top of the
Namur Sandstone Hutton Sandstone in Moorari 4 (Figure 19). The 70-9
Sand is a quartz to subarkosic arenite, which is locally
Depth clay-rich and micaceous and very similar to the lower
ft Birkhead Formation reservoir at Gidgealpa. Directly
above this sand is a thin, 15-m3 (6-in.) coal overlain by
6950 thinly interbedded, very fine grained sandstones and
siltstones/shales interpreted as lacustrine-to-coal
Formation

swamp facies. The very fine grained sandstones are


considered to be the effective seal. These sands have
7000 a significant labile rock fragment content, which may
correlate with the provenance change. However, the
major influence on displacement pressure does not
appear to be the pore-filling clays, as in the coarser-
grained sediments in Gidgealpa, but the fine grain
7050 size. Cores in other Moorari wells indicate that the
55 m (180 ft) of Birkhead Formation above the 70-9
Sand was deposited in a shallow-lacustrine environ-
ment with interbedded thin shales and fine-grained
sandstones. These have erosive bases that lead to
core

provenance change ? sand-on-sand stratification, which provides the weak


70-9 Sand link in this seal.
A series of samples was taken from the core in
Moorari 4. Mercury injection capillary pressure
7150 (MICP) curves were generated and compared with
petrological data using the same methodology as that
used for Gidgealpa 32. The effective seal to the 70-9
Sand is calculated to be capable of retaining a maxi-
Birkhead

mum oil column of 33.5–94.9 m (110–311 ft) (Table 2).


7200 The maximum oil column that is interpreted for the
70-9 Sand reservoir is ~14 m (45 ft). The interpreted
structural closure is at least 60 m (197 ft). Therefore,
the 70-9 Sand reservoir is neither full to capillary leak-
energy change age or structural spillpoint. The effective seal is a thick
accumulation of very fine grained lithic arenites to silt-
stones. Patchy shows that do occur within the seal are
Hutton Sandstone not thought to be tertiary migrating oil from the 70-9
Sand below, due to high displacement pressures and
7300 small oil column buoyancy. There is evidence of small
quantities of oil generated within the Birkhead Forma-
tion in the deeper parts of the basin such as at Moorari
(Smyth et al., 1984). This may be the case here, rather
Figure 19. Moorari 4 well log showing 37 m (120 ft) than being Permian-sourced oil that has leaked
of low-energy Birkhead Formation facies that occur through the 76-m- (250-ft-) thick Triassic seal.
below the provenance change. This is a well close to
Birkhead Formation depocenter. The seal is the fine- The Namur Sandstone, McKinlay Member, and
grained swamp/lacustrine facies that occurs >2203 m Murta Formation, Murteree Ridge, Eromanga Basin
(7090 ft).
The Murteree Ridge forms an embayment of base-
ment into the southern margin of the Permian Cooper
(Rodda and Paspaliaris, 1989). Minor amounts of gas Basin so that Permian and Jurassic sediments onlap
occur higher up in the Permian, but none has been onto its flank. The lowest Eromanga Basin sequence is
found to date in the Eromanga Basin. Only 0.3 × 106 m3 the Hutton Sandstone. Only minor oil accumulations
(2 million bbl) of oil occur in a thin sand (70-9 Sand) are present within the Hutton Sandstone in this area,
within the Birkhead Formation of the Eromanga Basin. due to the thin nature of the Birkhead Formation.
The Moorari oil and gas field (Figure 4) is located The Namur Sandstone is interpreted as a CD
closer to the depocenter of the Eromanga Basin than braided fluvial deposit similar to the Hutton Sand-
any of the other fields discussed in detail in this paper. stone and is widespread across the Eromanga Basin.
Here the Birkhead Formation is >94.5 m (310 ft) thick The McKinlay Member is conformable on top of the
(Figure 19). The regional Triassic Arrabury Formation Namur Sandstone and is interpreted as a waning
Œ

‡ˆ

Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production

energy, meandering fluvial-to-lacustrine shoreline


deposit similar to the base of the lower Birkhead For-
mation reservoir at Gidgealpa. The McKinlay Member
and Murta Formation are of more limited areal extent
than the Birkhead Formation. The Murta Formation
mainly comprises laminated carbonaceous siltstones
and fine-grained sandstones known as the “low-
permeability reservoir” (LPR) (Figure 20). It is inter-
preted to represent a change to a predominantly lacus-
trine depositional environment, with some fluvial
influence near its geographical extremities. The Murta
Formation is capped by a shaly type A seal. Within the
Murta Formation there are some very thin, very
mature sandy horizons that are known as the “high-

Log Depth
(Feet)
3860

3880
30

10
JENA 11
GR

PEF

,,,,,,,
180

yy
,,
,,
yy
,,
,,
yy
,,,

Interpreted
Facies
Cadna-Owie
Formation
161

{|€
permeability sands” (HPS). These are extremely pro- HPS

,,
yy
3900
ductive and are interpreted to represent lower to
upper shoreface sand bodies.
Together, these horizons form an important reser-

,,
yy
voir/seal/reservoir/seal quadruplet in the more

,,,
3920
prospective zones above the edge of the Triassic

,,
yy
regional seal. In fields with multiple stacked oil pools,
and thus abundant charge from the Permian, oil
pools in the Murta Formation are often present as the 3940

,,
yy
shallowest available oil.
Murta
Formation
Limestone Creek, Biala and Jena Oil Fields

,,
yy
of the Murteree Ridge Area 3960

The HPS and LPR of the Murta Formation in the


Murteree Ridge area are estimated to contain ≤11.6 ×

,,
yy
106 m3 (74 million bbl) of oil. The Namur Sandstone 3980
and McKinlay Member are estimated to contain ~2.5 ×

,,
yy
106 m3(16 million bbl) of oil (N. Williams, 1995, per-

,
sonal communication). The HPS is 1 ft thick and
occurs 36 m (120 ft) above the McKinlay Member 4000

,,
yy
reservoir (Figure 20). Siderite-cemented layers (type B
seals) ≤1 m (3 ft) thick are common within LPRs but

,,
yy
are not acting as effective seals, as indicated by post-
4020
production repeat formation test (RFT) data. They con-


tain numerous open and kaolinite-filled fractures, as
McKinlay
indicated by core logging and formation microscanner Member

,,,
interpretation. The fine sandstones and siltstones 4040
within LPR are interpreted to represent storm events.

,,,
,,,
They have erosive bases that lead to sand-on-sand
,,

stratification and the formation of highly tortuous ver-


,, ,,,
Namur

,,
yy
tical sand pathways, which provide the weak link in 4060 Sandstone
this seal. The sands and siltstones are thus, besides

,, ,,
yy
being a low-permeability reservoir, likely to be the Siderite Cemented Facies Barrier-Bar

,y
effective capillary seal (type E/D seals) to the underly-

,,
,,
ing McKinlay Member accumulations. Shoreface Facies Lagoonal Shales

,,
MICP data (Table 2) indicate that the fine sands and

,, ,,
Nearshore Facies

,,
siltstones of the LPR are capable of acting as a capillary Meandering Fluvial

seal, and the calculated minimum, but most probable,

,,
Lacustrine Shelf Facies
oil columns of 15 m (50 ft) are in good agreement with (Distal to Proximal) Braided Fluvial

actual oil columns within the McKinlay Member/ Offshore Facies


Namur Sandstone. Fracture stimulation of several wells
has resulted in an increased water cut, with the water
being derived from the LPR. This, in combination with
the patchy distribution of oil shows within the LPR, Figure 20. Jena 11 (Murteree Ridge) gamma-ray and
indicates that the oil that is present occurs either as part photoelectric factor (PEF) curves correlated with
of a transition zone or as a series of oil migration con- core-defined lithofacies showing the location of the
duits extending from the McKinlay Member to the HPS. high-permeability sand (HPS).
162 Boult et al.

pressure (psia)
2000

1500

1000

500
0 1 2 3 4 5
cummulative fluid production
(million bbl)

Figure 21. Limestone Creek/Biala (Murteree Ridge)


averaged pressure history data from all wells for the
HPS, showing that the pressure declined with pro- Figure 22. Repeat formation test data from Jena 11
duction within the HPS until just over 0.3 × 106 m3 and 12 in the Murteree Ridge area. All measured
(2 million bbl) of oil had been produced, when it pressures are above that of the theoretical gradient
stabilized. that would be associated with normal linear flow of
water from the McKinlay Member through the
Murta Formation to the HPS.
Under production, the HPS initially shows a rapid
pressure drop. This eventually stabilizes, indicating a
weak pressure support mechanism (Figure 21). Con- (1989) noted a statistical link between hydrocarbon
ventional drive mechanisms have been selectively occurrence within the Eromanga Basin and the occur-
eliminated as potential sources of water (Williams et rence of calcites within the Namur Sandstone. Struc-
al., 1994). Repeat formation test data indicate that a tures with calcite in the Namur Sandstone are
pressure sink extends below the HPS for at least 20 m three-and-a-half times more likely to contain oil than
(65 ft) in Jena 11 (Figure 22). All the RFT pressures are those that do not contain calcite. However, 51% of
above the theoretical pressure gradient that would be structures that have associated calcite are still dry. This
associated with flow of water from the McKinlay last figure reminds us that an MRDZ only shows up in
Member through the rock matrix into the HPS. Con- migration pathways under a given set of chemical con-
duits for flow are not provided by fractures but by the ditions; if there is no trap available, oil will not have
highly tortuous sand pathways previously alluded to accumulated.
in the LPR. Hence, it is suggested that over geological Figure 24 is a schematic representation of the con-
time these conduits have allowed oil (the nonwetting trols on oil migration within a fluvio/lacustrine
phase) to migrate into the HPS, while over the time of sequence such as that of the Jurassic Eromanga Basin
production they allow only water (the more mobile sequence. It assumes that capillary leakage is the
wetting phase) to flow into the HPS, providing a weak dominant sealing mechanism. It also assumes the
pressure support (Figure 23). faults are not active, and thus the most common sce-
nario (Downey, 1993) applies: faults are neither seals
DISCUSSION nor conduits to fluid flow. Figure 24 introduces the
notion of “seal bite,” which describes the vertical off-
The most important aspect of seal evaluation is set of sealing lithologies along a migration pathway
establishing which lithofacies within a formation is the and its control on oil migration. Seal bite may be
weak link and thus the effective seal. Within a sand- caused by either fault offset, minor folding, or facies
rich fluviatile/shallow lacustrine system where over- interdigitation. Seal bite also depends on the fluid
bank and background sedimentation shales are not properties involved in the control of oil migration,
very thick, they are not likely to be the effective seal to such as hydrocarbon buoyancy and capillary pres-
migrating hydrocarbons. The effective seal within such sures generated. Thus, along a migration pathway
a facies must be the intervening sandy facies. This is that occurs beneath a dipping seal, if seal bite is
especially true when fault offset juxtaposes sandy greater than the possible retainable oil column, verti-
interbeds. Shales and siderite layers may, however, cal migration will be initiated; otherwise, lateral
form significant barriers on a production time scale, migration will continue. Figure 24 shows the follow-
creating highly tortuous pathways for fluid flow. ing features that control oil migration.
Another important aspect of seal evaluation is an
assessment of possible migration pathways beyond the • That weak-link lithologies, such as labile or fine-
seal under investigation. The migration-related diage- grained sandstone, control the location of hydro-
netic zones (MRDZs) found within formations above a carbon traps.
seal seem to be good candidates for indicating possible • That faults do not need to be sealing or acting as
leakage of buoyant fluids from below. Jensen-Schmidt conduits to focus oil migration conduits. It is the
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 163

A
A B
HPS
Murta Fm

Top
McKinlay

Expected RFT profiles show that pressure


depletion within the LPR (due to production
McKinlay Mbr

of water) would only extend down to the


Namur Sst.

McKinlay when measured outside the area


of the Mckinlay oil pool. Profile B is an actual
RFT profile

Murta Member - Lacustrine facies


Water Filled Sands

Oil Filled Sands

Migration conduit Pressure support (brine) conduit


Vertical migration of oil through When pressure depletion in HPS
has reached a critical value water
the crest of the structure from the can leak through from the Namur
McKinlay to the HPS aquifer

Figure 23. An oil migration and pressure support model for the Murta Formation. RFT = repeat formation
test, HPS = high-permeability sand, and LPR = low-permeability reservoir.

resultant juxtapositions of weak-link lithologies fluvio/lacustrine depositional environments that are


that do this. supplied with alternating CD and VAD sediments are
• If the seal bite is greater than the maximum hydro- important here; for example, the Surat and Carpen-
carbon column that a seal will hold back, vertical taria basins (McConache et al., 1994).
migration may be initiated before major struc- Where the labile sandstone is the seal, its displace-
tures are accessed and filled. ment pressure depends not only on the existence of
• If the seal bite is less than the maximum petroleum pore-blocking diagenetic clays but also on the amount
column a seal will hold back, the petroleum will of framework collapse that has taken place. It would
migrate predominantly along bedding into, and seem the more proximal this type of seal is to the
begin to fill, major structures. However, they may source of the labile material, the higher its displace-
not fill to structural spillpoint, because once the ment pressure will be. Depth of burial seems to have a
capillary displacement pressure of the seal to the lesser effect on its displacement pressure. The Jackson
structure is exceeded, vertical migration will be Birkhead Formation seal, which is at a depth of 1430 m
initiated. (4700 ft), is shallower and slightly coarser grained than
at Gidgealpa at 1830 m (6000 ft), and we might expect
This kind of control may be the reason why at the seal at Jackson to have a lower displacement pres-
Gidgealpa (Figure 14), the Poolowanna Formation oil sure than at Gidgealpa. But Jackson is also more prox-
accumulation is absent from the crest of the south imal, with a higher labile content. This has had a
dome, but Hutton Sandstone/Birkhead Formation oil greater influence than depth or grain size; thus, the
is present at the crest of the dome. seal at Jackson is capable of retaining ≤36 m (120 ft) of
We have investigated the Birkhead Formation oil, whereas at Gidgealpa it has probably leaked when
first and in most detail because this is the main seal the oil column has reached 20 m (66 ft).
within the Eromanga Basin. It is probable that simi- Where the very fine sandstones and siltstones of a
lar mechanisms of hydrocarbon entrapment will lacustrine facies are the effective seal, as in the Murta
apply, not only for other horizons within the Jurassic Formation on the Murteree Ridge and possibly
of the Eromanga Basin, but for the other Mesozoic above the 70-9 Sand at Moorari, it would appear that
basins in northeast Australia. Those containing burial has a major influence. The base of the Murta
164 Boult et al.

coals, occurred just prior to the major influx of labile


sandstone; this tends to support tectonic control on river
Migration base levels.
ground
water Related
flow Diagenetic CONCLUSIONS
Zone
The location of the effective seal within the Birk-
head Formation, away from its depocenter, is directly
controlled by the influx of volcanic-arc-derived (VAD)
labile sandstone into the depositional system. Within a
fluvio/shallow-lacustrine system where shales are not
laterally very continuous due to either erosion, dewa-
tering, or faulting, they are only likely to trap minor
Seal bite < Th amounts of oil. The effective seal in the examples dis-
cussed, where the lacustrine facies are not well devel-
oped, is caused by the diagenesis and compaction of
the VAD sediment within the Birkhead Formation.
Seal bite Where the shallow-lacustrine facies are well devel-
> Th oped, the fine sands are the effective seal. Their effec-
tiveness may be enhanced by diagenesis.
Important characteristics of the Hutton Sand-
Oil sourced
from Permian stone/Birkhead Formation transition in predicting oil
occurrence and production characteristics are: (1) the
Lithic sandstone - effective seal Oil pools
Discontinuous fluvial shale seals
proximity of the location to the volcanic lithic source
Reservoir sandstones
area (closer to source area leads to better quality seal);
Figure 24. An oil migration and entrapment model (2) the depth of burial of the seal: a greater depth of
for the Eromanga Basin, showing how the interac- burial leads to better quality seal, especially in the
tion of “seal bite” and maximum oil column height finer grained facies; and (3) the stratigraphic thickness
(Th) controls the location of migration pathways. of the upper Hutton Sandstone/lower Birkhead For-
This model assumes that faults are neither sealing mation low-energy facies. A thinner upper Hutton
nor conduits to flow. Sandstone allows closer proximity of the good-quality
lower Hutton Sandstone reservoir to the main regional
VAD sandstone seal within the Birkhead Formation.
Formation of the Murteree horst area is ~1220 m
(4000 ft), and the base of the mid-Birkhead Formation
lacustrine facies at Moorari is at 2152 m (7065 ft). The
APPENDIX
former would be classified as a type D/E seal capable Development of Seal Capacity Equation
of holding a column of 15 m (50 ft), and the latter
would be classified as a type C seal capable of hold- The maximum size of an oil column that can be
ing 60 m (200 ft) of oil. trapped beneath a membrane seal occurs when the
There also appears to be a link between the tectonics buoyancy of that column is equal to the capillary pres-
that produce the periodic influx of labile material and sure generated by the seal. This can be represented by
the change from high-energy to low-energy depositional equation 1:
environments. It may be that the local downwarping of
both cratonic and foreland basins has changed river base Pch/w ( seal ) – Pch/w ( reservoir ) = ( SGw – SGh ) • g • Th (1)
levels and created ephemeral shallow freshwater lakes
and swamps. It also may be that the periodic activation
of the volcanic arc to the east blocks up the generally Converted to field units (feet) and where the parame-
northward and eastward flow of the rivers either ters and units are Pch/w = capillary displacement pressure
directly or by dumping large amounts of sediment in (psia); SGh = subsurface density of hydrocarbon (g/cm3);
their path, as shown in Figure 6. Alternatively, this could SGw = subsurface density of water (g/cm3); g = gravity;
be a combination of these effects. A facies change prior and Th = height of hydrocarbon column (ft), then
to a provenance change at many locations across the
Eromanga Basin is evidence that river base level change Pch/w ( seal ) – Pch/w ( reservoir ) = ( SGw – SGh ) • Th × 0.433 (2)
is then followed by the rivers flowing predominantly
from the volcanic arc across the foreland basin and into Rearranged
the epicratonic basin. This simplified model is compli-
cated by large- and small-scale basin warping, which
was probably associated with episodic activation of the Pch/w ( seal ) – Pch/w ( reservoir )
volcanic arc to the east. There is also evidence that a hia- Th = (3)
tus in sedimentation, in the form of soil horizons and (SGw – SGh )0.433
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 165

Which could be written is smooth and mercury closure around it is good, the
extrapolation of the curve plateau back to the pressure
Pch/w ( seal ) – Pch/w ( reservoir ) axis is used. If the closure is poor, the preferable
Th = (4) method is to use the incremental plot to determine
(gradw – gradh ) where the apparent mercury displacement increases
significantly.
for meters
ACKNOWLEDGMENTS
Pch/w ( seal ) – Pch/w ( reservoir )
Th = (5) We would like to thank the Energy Research and
(SGw – SGh )1.42 Development Corporation, Canberra, Australia, for
funding project 1541, “Quantification of hydrocarbon
Conversion from Lab Data to Field Data migration and entrapment,” which initiated the search
for Seal Capacity into seals within the Eromanga Basin. We would also
like to acknowledge R. Seggie and N. Williams of
Capillary pressure has been defined within basic SANTOS Ltd.; D. Gravestock of the South Australian
physics by equation 6: Department of Mines and Energy; J. Schulz-Rojahn of
the National Centre for Petroleum Geology and Geo-
2. IFT .Cosθ physics; and J. Jago, E. Lanzilli, and D. Folks of the
Pc = (6)
r University of South Australia for valuable discussion
during the preparation of this document.
Where IFT = interfacial tension, and θ = the contact
angle of the two fluids against the solid measured REFERENCES CITED
through the wetting phase.
Rearranged Ambrose, G., R. Suttill, and I. Lavering, 1986, The geol-
ogy and hydrocarbon potential of the Murta Mem-
2 × IFT × Cos θ ber, Mooga Formation, in the southern Eromanga
r= (7) Basin, in D.I. Gravestock, P.S. Moore, and G.M. Pitt,
Pc eds., Contributions to the geology and hydrocarbon
potential of the Eromanga Basin: Geological Society
Therefore, if r is constant of Australia Special Publication 12, p. 71–84.
Boult, P.J., 1993, Membrane seal and tertiary migration
2 × IFTh/w × Cosθ h/w 2 × IFTm/a × Cosθ m/a pathways in the Bodalla South oilfield, Eromanga
= (8) Basin, Australia: Marine and Petroleum Geology,
Pc h/w Pc m/a v. 10, no. 1, p. 3–13.
Boult, P.J., and P.N. Theologou, 1993, Quantification of
Rearranged hydrocarbon migration and entrapment: Unpub-
lished end of grant report, Energy Research and
IFTh/w × Cosθ h/w Development Corporation, project no. 1541.
Pc h/w = × Pc m/a (9) Core Lab, 1982, A course in special core analysis,
IFTm/a × Cosθ m/a
D. Keelan, ed., p. 3–12.
Dodman, A.P., and J.T. Rodriques, 1989, The Jackson
Substitution of equation 9 into 5 for meters gives oil field development, in B.J. O’Neil, ed., The
Cooper and Eromanga Basins, Australia: Proceed-

Th =
{
IFTh/w × Cosθ h/w × Pc m/a( seal ) – Pc m/a( reservoir ) } (10)
ings of the Petroleum Exploration Society of Aus-
tralia, Society of Petroleum Engineers, Australian
IFTm/a Cosθ m/a × ( SGw – SGh )1.42 Society of Exploration Geophysicists (SA Branches),
Adelaide, p. 81–89.
If the seal is water-wet with the contact angle Dolan, P., P.D. Griffiths, and S.R. Welton, 1988, The
assumed to be 0 and accounting for mercury/air IFT of successful recovery of well productivity in the
484 N/m and contact angle of 140°, then Bodalla South Field: The Australian Petroleum
Exploration Association Journal, p. 7–18.

Th =
{
IFTh/w × Pc m/a( seal ) – Pc m/a( reservoir ) } (11)
Downey, M., 1993, Faults, leaks or seals? Presented at
Hedburg Conference, “Seals: a multidisciplinary
(SGw – SGh )526 approach,” Crested Butte Colorado, June 21–23.
Foden, J., J. Mawby, S. Kelley, S. Turner, and D. Bruce,
1995, Metamorphic events in the eastern Arunta
Reading MICP Curves to Inlier, Part 2: Nd-Sr-Ar isotopic constraints: Pre-
Determine Displacement Pressure cambrian Research, v. 71, p. 207–227.
We always plot the incremental and cumulative Gravestock, D.I., M. Griffiths, and A. Hill, 1983, The
plots of mercury intrusion vs. pressure. If the sample Hutton Sandstone—two separate reservoirs in the
166 Boult et al.

Eromanga Basin, South Australia: The Australian Eromanga Basins, Australia: Proceedings of the
Petroleum Exploration Association Journal, Petroleum Exploration Society of Australia, Society
p. 109–119. of Petroleum Engineers, Australian Society of
Green, J.C., and J.F. Thomas III, 1993, Extensive felsic Exploration Geophysicists (SA Branches), Adelaide,
lavas and rheoignimbrites in the Keweenawan Mid- p. 91–101.
continent Rift plateau volcanics, Minnesota: McPhie, J., M. Doyle, and R. Allen, 1993, Volcanic tex-
petrographic and field recognition: Journal of tures: A guide to the interpretation of textures in
Vulcanology and Geothermal Research, v. 54, volcanic rocks: University of Tasmania, Centre for
p. 177–196. Ore Deposit and Exploration Studies, Hobart, 198 p.
Habermehl, M.A., 1980, The Great Artesian Basin, Miller, R.G., R.K. O’Nions, P.J. Hamilton, and E.
Australia: Bureau of Mineral Resources Journal of Welin, 1986, Crustal residence ages of clastic sedi-
Australian Geology and Geophysics, v. 5, p. 9–38. ments, orogeny and continental evolution: Chemi-
Harris, N.B., 1989, Diagenetic quartz arenite and cal Geology, v. 57, p. 87–99.
destruction of secondary porosity: An example Momper, J.A., 1980, Oil expulsion—a consequence of
from the Middle Jurassic Brent Sandstone of NW oil generation: AAPG Distinguished Lecture: New
Europe: Geology, v. 17, p. 361–364. York, Science-Thru-Media, Inc.
Hawlader, H.M., 1990, Diagenesis and reservoir O’Brien, G.W., and E.P. Woods, 1995, Hydrocarbon-
potential of volcanogenic sandstones—Cretaceous related diagenetic zones (HRDZs) in the Vulcan
of the Surat Basin, Australia: Sedimentary Geology, sub-basin, Timor Sea: recognition and exploration
v. 86, p. 181–195. implications: The Australian Petroleum Explor-
He, F., and P.J. Conaghan, 1994, Diagenesis of Jurassic ation Association Journal, p. 220–252.
and Lower Cretaceous sandstones of the Eromanga Power, P.E., and S.B. Devine, 1970, Surat Basin, Aus-
Basin in New South Wales: Australian Geological tralia—sub-surface stratigraphy, history and
Survey Organization Journal of Australian Geology petroleum: AAPG Bulletin, v. 54, p. 2410–2437.
and Geophysics, v. 15, no. 2, p. 191–215. Powis, G.D., 1989, Revision of Triassic stratigraphy
Heath, R., S. McIntyre, and N. Gibbins, 1989, A Per- at the Cooper Basin to Eromanga Basin transition,
mian origin for Jurassic reservoired oil in the Ero- in B.J. O’Neil, ed., The Cooper and Eromanga
manga Basin, in B.J. O’Neil, ed., The Cooper and Basins, Australia: Proceedings of the Petroleum
Eromanga Basins, Australia: Proceedings of the Exploration Society of Australia, Society of
Petroleum Exploration Society of Australia, Society Petroleum Engineers, Australian Society of Explo-
of Petroleum Engineers, Australian Society of ration Geophysicists (SA Branches), Adelaide,
Exploration Geophysicists (SA Branches), Adelaide, p. 265–277.
p. 405–416. Rodda, J.S., and T.G. Paspaliaris, 1989, Tirrawarra and
Hill, L.V., 1985, Environmental analysis of the Hutton Moorari oil fields enhanced oil recovery schemes—
Sandstone to Birkhead Formation transition within further development, in B.J. O’Neil, ed., The
the south-western Eromanga Basin, Queensland: Cooper and Eromanga Basins, Australia: Proceed-
Honor’s thesis, Adelaide University. ings of the Petroleum Exploration Society of
Hird, K., 1995, Oil and gas report, J.B. Wear and Son: Australia, Society of Petroleum Engineers, Aus-
Australian Research Report, March, 72 p. tralian Society of Exploration Geophysicists (SA
Hollingsworth, R.J.S., 1989, The exploration history Branches), Adelaide, p. 121– 129.
and status of the Cooper and Eromanga basins, in Rootare, H.M., 1970, A review of mercury porosime-
B.J. O’Neil, ed., The Cooper and Eromanga basins, try: advanced experimental techniques in powder
Australia: Proceedings of the Petroleum Explo- metallurgy: New York, Plenum Press, p. 225–252.
ration Society of Australia, Society of Petroleum Ryan-Grigor, S., and J.P. Schulz-Rojahn, 1995, Seismic
Engineers, Australian Society of Exploration Geo- delineation of structure-controlled carbonate
physicists (SA Branches), Adelaide, p. 3–9. cement and potential economic implications, Angel
Hovland, M., M.R. Talbot, H. Qvale, S. Olaussen, and Field, North West Shelf: The Australian Petroleum
L. Aasberg, 1987, Methane-related carbonate Exploration Association Journal, p. 280–295.
cements in pockmarks of the North Sea: Journal of Schowalter, T.T., 1979, Mechanics of secondary hydro-
Sedimentary Petrology, v. 57, p. 881–892. carbon migration and entrapment: AAPG Bulletin,
Jensen-Schmidt, B., 1989, Calcite as an oil indicator in v. 63, p. 723–760.
the Eromanga Basin, South Australia: Unpublished Schulz-Rojahn, J.P., 1993, Calcite-cemented zones in
report, Monitoring Western/Central Division the Eromanga Basin: clues to petroleum migration
(ESSO). and entrapment: The Australian Petroleum Explo-
McConache, B.A., P.W. Stainton, M.G. Barlow, and ration Association Journal, p. 63–76.
J.N. Dunster, 1994, The offshore Carpentaria Singh, R., 1990, The seismic response of calcite
Basin—Gulf of Carpentaria, North Queensland: cemented sandstone—Gidgealpa: Master’s thesis,
The Australian Petroleum Exploration Association University of South Australia.
Journal, p. 614–625. Smyth, M., A.C. Cook, and R.P. Philp, 1984, Birkhead
McIntyre, S.M., C.L. Jamal, C.L. Pidcock, and M.I. revisited: petrological and geochemical studies of the
Kabir, 1989, The Gidgealpa oil and gas field: a Birkhead Formation, Eromanga Basin: The Australian
case history, in B.J. O’Neil, ed., The Cooper and Petroleum Exploration Association Journal, p. 230–242.
Capillary Seals Within the Eromanga Basin, Australia: Implications for Exploration and Production 167

Sneider, R.M., and K. Stolper, 1991, Petrophysical Petroleum Exploration Association Journal 1994,
properties of seals (abs.): AAPG Bulletin, v. 75, p. 320–329.
no. 3, p. 673–674. Williams, N.V., P.N. Theologou, P.J. Boult, M.
Theologou, P.N., 1995, Murta Formation/McKinlay Zwigulis, R.P. Schlicting, and R.A. Price, 1994,
Member of the Murteree Ridge Nappagoongee- Unusual lacustrine reservoirs and seals of the
Murteree Block—improved oil recovery project: Murteree Horst area, Eromanga Basin, South Aus-
Ph.D. thesis, University of South Australia. tralia: Society of Petroleum Engineers paper no.
Townsend, G., 1993, Distribution, chemistry, isotopic 28751, presented at the Asia Pacific Oil & Gas Con-
composition and origin of diagenetic carbonates: ference, Melbourne, Australia, November 7–10,
Namur Sandstone, Cooper Eromanga Basin, South p. 99–112.
Australia: Honor’s thesis, University of Melbourne. Wiltshire, M.J., 1989, Mesozoic stratigraphy and
Veevers, J.J., ed., 1984, Phanerozoic Earth History of palaeogeography, eastern Australia, in B.J. O’Neil,
Australia: New York, Oxford University Press, ed., The Cooper and Eromanga Basins, Australia:
418 p. Proceedings of the Petroleum Exploration Society
Watts, K.J., 1987, The Hutton Sandstone–Birkhead of Australia, Society of Petroleum Engineers, Aus-
Formation transition, ATP 269, Eromanga Basin: tralian Society of Exploration Geophysicists (SA
The Australian Petroleum Exploration Association Branches), Adelaide, p. 279–291.
Journal, p. 215–228. Zoellner, E., 1988, Geology of the Early Cretaceous
Whitford, D.J., P.J. Hamilton, and J. Scott, 1994, Murta Member, Mooga Formation, in the Copper
Sedimentary provenance studies in Australian [sic] Basin area, South Australia and Queensland:
basins using neodymium model ages: Australian Ph.D. thesis, Flinders University of South Australia.
Uygur, K., H. Is, and M.A. Yukler, 1997, Reservoir characteriza-
tion of Cretaceous Mardin Group carbonates in Bölükyayla-
Cukurtas and Karakus Oil Fields, SE Turkey: a petro-graphic
and petrophysical comparison of Overthrust and Foreland
Zones, in R.C. Surdam, ed., Seals, traps, and the petroleum
system: AAPG Memoir 67, p. 169–198.

Chapter 11

Reservoir Characterization of Cretaceous


Mardin Group Carbonates in Bölükyayla-
Cukurtas and Karakus Oil Fields, SE Turkey:
A Petrographic and Petrophysical
Comparison of Overthrust and Foreland Zones
Kadir Uygur
Huseyin Is
TPAO Exploration Group
Ankara, Turkey

M. Arif Yükler
IBA Inc.
Dallas, Texas, U.S.A.

ABSTRACT
Approximately 95% of Turkey’s total oil production comes from south-
east Turkey, with 70% of the 95% from Cretaceous Mardin Group carbon-
ates. This study is carried out to evaluate the petrophysical and
petrographic properties of the source-reservoir-seal carbonate intervals of
the Mardin Group oil fields of the foreland area and oil fields of the Upper
Cretaceous overthrust frontal zone of southeast Anatolia. The data include
thin sections, cores and plugs, drill-stem tests, electrical logs, organic geo-
chemistry, and basin analyses results from 65 exploration wells in both
regions.
The Aptian–Lower Campanian Mardin Group is deposited on the shelf-to-
intrashelf part of a passive continental margin of the Arabian plate. Relative
sea level changes in the Cretaceous are responsible for three main shallowing-
upward cycles that produced three reservoir intervals separated from each
other by source and/or seal intervals. Each of the cycles is underlain and
overlain by unconformity surfaces.
Structurally, the oil fields of the overthrust frontal zone and the foreland
area studied are represented by the Cretaceous imbricated structures and the
Miocene wrench system, respectively. Large accumulations of hydrocarbons
have been trapped along both the east-west elongated, narrow, and asymmet-
rical thrusted anticlines of the imbricated zone and the northeast-
southwest–trending en echelon anticlines of the wrench system.

169
170 Uygur et al.

The source, reservoir, and seal rocks are all carbonates. The oil, which is
in fractured dolomites and limestones, is 22˚–35˚API. Porosity, permeabili-
ty, and water saturation of reservoir pay zones are 3–15%, 0.1–7000 md,
and 20–50%, respectively. Secondary porosity is dominant and, in decreas-
ing order of abundance, it occurs as dolo-intercrystalline, vugular, and
fracture types.
The Mardin Group carbonates have undergone both early and late dia-
genesis resulting from the existence of shallowing-upward depositional
cycles and burial-tectonic stresses during the Cretaceous and Tertiary
Periods. The important effects of diagenesis on reservoir quality have been
fracturing and dolomitization. Intrafracture mineralization, mineral
replacement, and dissolution also play an important role in determining
reservoir quality. The reservoir performance, on the field scale, is good at
the structurally high intervals due to increased frequency of fractures and
intensified late dolomitization.
The same and/or similar static hydraulic regimes prevail in the double
porosity system reservoirs in and between the oil fields. There is an increase
in initial reservoir pressure with depth for both foreland and overthrust
frontal zones.
The main factors affecting hydrocarbon generation in source rock for the
overthrust frontal zone in the northwest and the foreland area in the south-
east are thickness of both the Campanian Karadut-Kocali and the Upper
Miocene allochthonous units, the thickness of the Eocene–Miocene carbon-
ates, and clastics.
Due to gravitational segregation, the API gravity of oil decreases with
depth in both regions. The API gravity of oil also decreases gradually from
the overthrust frontal zone to the foreland area.

INTRODUCTION This study, an extension of Uygur and Gürel (1991),


is based on information from 65 exploration and pro-
The study area (Figure 1) is located in the foreland duction wells. Analysis of drill-stem tests, electrical
and Upper Cretaceous overthrust frontal zones of logs, cores and plugs, thin sections, and organic geo-
southeast Turkey. It is near the cities of Adiyaman to chemistry from these wells, including four selected
the west, Urfa to the south, and Diyarbakir to the east. wells modeled using the quantitative basin analysis
It covers ~190 km2 of subsurface area and consists of software of YUKLERPC ® (Yükler and Welte, 1980;
six oil-producing fields: Karakus, K. Karakus, Cen- Yükler, 1987), make up the basis for this study.
dere, and G. Karakus of the foreland, and Bölükyayla
and Cukurtas of the overthrust frontal zone (Figure 1). Stratigraphy
This study is a petrophysical evaluation and com-
parison of the Mardin Group carbonates in these oil The Mardin Group carbonates are Aptian–Early
fields and is the product of four years’ endeavors con- Campanian in age (Figure 2). The group is mainly
ducted by the TPAO Exploration Group. This paper limestone and dolomite in both regions. Approxi-
introduces the stratigraphy, depositional environ- mately 300 m of Mardin Group carbonates were pene-
ments, and structural position of the Mardin Group trated by the wells. The group consists of four
carbonates with respect to reservoir quality. In addi- mappable rock units of the three main cycles (Figure
tion, it explains, interprets, and compares diagenetic 2): the Areban-Sabunsuyu (Aptian–Cenomanian), the
and reservoir characteristics (i.e., porosity, permeabil- Derdere (Cenomanian–Turonian?), and the Karababa
ity, static reservoir pressure, API oil gravity, hydrocar- (Santonian–Lower Campanian) formations. These for-
bon potential) of the Mardin Group carbonates in the mations are separated from each other by unconformi-
oil fields that were studied. These fields produce ties. The contacts at the bottom and top of the Mardin
>50% of total crude production of TPAO. Group are also unconformities (Figure 2).
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 171

Figure 1. Index map of the study area.

DEPOSITIONAL ENVIRONMENT AND STRUCTURAL POSITION AND


RESERVOIR QUALITY RESERVOIR QUALITY
The Mardin Group carbonates are deposited on the The Karakus, K. Karakus-Cendere, and G. Karakus
shelf and intrashelf part of a passive continental mar- oil fields of the foreland area (Figure 4) are located just
gin of the Arabian plate (Horstink, 1971; Uygur and to the south of the frontal part of the Orogenic belt.
Aydemir, 1988; Celikdemir et al., 1991). During the The Bölükyayla and Cukurtas oil fields are in the
Cretaceous, relative sea level changes (Gorur et al., foothills of the Orogenic belt (Figure 4).
1987) caused the formation of three main shallowing- The foreland area is represented by gentle, low-
upward depositional cycles and several subcycles relief anticlines and faulted folds (Akgül and Pasin,
(Figure 3) in the Mardin Group carbonates. Unconfor- 1992, personal communication). The foothills of the
mities between the cycles are also evident as erosional Orogenic belt are partially characterized by the
or nondepositional surfaces by sequence stratigraphic Upper Miocene and Upper Cretaceous imbricated
analysis (Tardu et al., 1990), who state that the Mardin overthrust belt. The foreland is to the south of the
Group consists of at least three different depositional imbricated Cretaceous overthrust frontal zone (Fig-
sequences. Both thickening-upward and coarsening- ure 4). In general, the present tectonic features of the
upward sequences are observable in each of the three study area and adjacent areas are developed as the
main shallowing-upward cycles. result of the collision of the two continental masses:
The three shallowing-upward cycles initially cre- the Arabian and Anatolian plates (Horstink, 1971;
ated the reservoir intervals separated by source Dewey et al., 1973).
and/or seal intervals (Figure 3). In general, the lower Structurally, the oil fields of the foreland are on the
parts of cycles II and III are source and seal rocks, with Miocene wrench system (Çemen and Akgül, 1991, per-
reservoir rocks generally in the upper and middle part sonal communication). This system, known as the
of each cycle. These three reservoir intervals, particu- Adiyaman Fault zone, is shown in Figure 5. The sys-
larly the dolomites of the second cycle (Derdere For- tem, very similar to the classic examples given by
mation), are the main oil producers and are often Wilcox et al. (1973) and Hobbs et al. (1976), is an
interconnected by tectonically formed fractures and oblique-slip type with dominantly sinistral (left-
dissolution pathways. Unconformity surfaces between lateral) displacement (Brown, 1991; Pasin, 1992, per-
the cycles, which introduce additional porosity and sonal communication). It contains a series of north-
permeability, may also play an important role in deter- east-southwest–trending left-lateral strike-slip faults
mining reservoir quality (Wagner and Pehlivan, 1985; and en echelon folds (Figure 5). However, Uysal and
Wagner et al., 1986; Uygur and Aydemir, 1988; Duran Duygu (1992, personal communication) observed that
and Aras, 1990; Celikdemir et al., 1991). left-lateral displacement of the Adiyaman Fault zone is
172 Uygur et al.

Figure 2. General
stratigraphy of the
Mardin Group
(Celikdemir et al., 1991).

not dominant, and the structure of the area has been and acoustic (BHTV) images. They observed that frac-
formed primarily by Upper Cretaceous thrusting and ture densities and orientations are related to local
reactivation of Upper Cretaceous thrusting that structural events such as faults, folds, and unconfor-
occurred during the Miocene. mities.
Large accumulations of hydrocarbons have been The Bölükyayla and Cukurtas oil fields of the
trapped in the northeast-southwest–trending en eche- foothills are located in the Cenomanian–Turonian
lon folds of the wrench system. The system, which Karadut allochthonous complex area, which is mainly
belongs to the East Anatolian Fault System, is believed composed of carbonate flysch. The emplacement of the
to be Middle Miocene (Çemen and Akgül, 1991, per- allochthonous complex at the northern margin of
sonal communication), Middle Eocene–Miocene (Yil- southeast Anatolia and into the Kastel Formation
maz et al., 1991), Miocene (Perinçek and Çemen, 1991) occurred during the Late Cretaceous (Akgul and Erdo-
in age and the result of late Tertiary reactivation gan, 1992). Below the Karadut complex, the Sayindere
(Brown, 1991). The basin modeling results show that Formation, Karabogaz Formation, and Mardin Group
the expulsion and accumulation of oils took place from carbonates are present (Figure 4).
the Miocene to the Pleistocene (Yükler, 1987). There-
fore, if structural wrenching took place during SOURCE, RESERVOIR,
Miocene times, then the earliest that hydrocarbon AND SEAL ROCKS
migration into the traps could have occurred was in
the Miocene. This has been confirmed by the applica- The source, reservoir, and seal rocks of the oil fields
tion of other basin analysis methods in the area. are all carbonates (Table 1). As explained above, gen-
The fractures in these carbonate reservoir rocks erally the lower parts of each shallowing-upward
have significant contribution to the reservoir quality, cycle, with the exception of cycle I, contain source
especially to the productivity (Tore et al., 1992). and/or seal rocks, and the upper and middle parts con-
Ozkanli and Standen (1991) studied the degree of frac- tain reservoirs. The representative ranges of porosity
turing, orientation of fracture systems, and fracture (θ), permeability (K), and total organic carbon (TOC)
porosity in these reservoirs by both electrical (FMS) values are listed in Figure 3.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 173

Figure 3. Generalized sketch of lithostratigraphy, facies, depositional cycles, areal


distribution, tectonic setting, and economic importance of Mardin Group carbonates
(adapted from Wagner and Tuna, 1988).
174 Uygur et al.

Figure 4. Schematic structural cross section of southeast Turkey (modified from Horstink, 1971; Akgul and
Erdogan, 1992).

Source Rocks Karababa-C member are reservoirs of secondary


importance (Table 1). Unconformity surfaces, which
The dominant source rock in the area is a phos-
are the sites for dissolution and karstification features,
phatic mudstone of the Karabogaz Formation (Table
between both the Karabogaz and Karababa formations
1, Figure 6). This source rock is typified by its dark
and between the Karababa and Derdere formations
color, mud-supported texture, TOC content, hydro-
also enhance the reservoir quality of the Mardin
gen index (HI), and Tmax values (Table 2). The values
Group carbonates.
in Table 2 indicate type II kerogen and moderate
The reservoir rocks, mainly fractured dolomites
maturity (Soylu and Gurgey, 1991, personal commu-
and limestones, do not contain significant primary
nication). In general, in a southeast (foreland area) to
porosity. Diagenesis, karstification, and tectonics
northwest (overthrust zone) direction, TOC content
have generated major secondary porosity develop-
decreases, maturity increases and amount of amor-
ment through various stages of dolomitization, dis-
phous organic matter increases (Soylu, 1994, per-
solution, and fracturing.
sonal communication).
In addition, tectonically controlled fractures (Fig-
The lower parts of cycle III (Karababa-A member
ure 8) play a dominant role in the development of
of the Karababa Formation) and cycle II (the lower
well productivity by causing the reservoir intervals
part of the Derdere Formation) also have source rock
potential (Table 1, Figure 3). to be interconnected with each other.

Reservoir Rocks Seal Rocks


As indicated by cores, logs, drill-stem test evalua- Clayey, fossilliferous wackestone of the Sayindere
tions, thin sections, and porosity and permeability data Formation is the dominant seal rock type (Table 1)
for the Mardin Group carbonates, the lower part of the for the Mardin Group reservoirs. This seal is at the
fractured Karabogaz Formation, dolomitic wackestone top of the Karabogaz Formation. In general, the
of the Karababa-C member of the Karababa Formation, source rocks of the area are also effective seal rocks.
and especially the dolomites (Figure 7) of the Derdere The lower part of the Derdere Formation may be a
Formation, are the dominant reservoir rocks (Figure 3, seal for the Sabunsuyu reservoir. The Karababa-A
Table 1) in the area. Dolomites of the Sabunsuyu For- member, which is another potential seal, is above the
mation, grainstones of the upper part of the Derdere Derdere reservoir. All the seals are either mudstones
Formation, and the bioclastic wackestones of the or wackestones (Table 1).
Figure 5. Structure contour map of the K. Karakus-Cendere (upper right), Karakus (center),
G. Karakus (lower right), and Cukurtas and Bölükyayla (upper left) oil fields (modified from
Pasin, 1991).
176 Uygur et al.

Table 1. Source, Reservoir, and Seal Carbonates of the Studied Oil Fields, Southeast Turkey.

Oil Fields Source Rock Reservoir Rock Seal Rock


Karaboğas Fm. Phosphatic Mudstone Wackestone Mudstone/Wackestone
Sayindere Fm. Clayey Fossiliferous
Wackestone
Karababa-A Mudstone/Wackestone Mudstone/Wackestone (?)
Member
Karababa-C Bioclastic Wackestone
Member Dolomitic Wackestone
Derdere Fm. Mudstone/Wackestone Dolomite Mudstone/Wackestone
(lower part) Grainstone (lower part?)
Sabunsuyu Fm. Dolomite
Unconformity(?) Karaboğaz/Karababa
Surfaces Karababa/Derdere

DIAGENESIS AND RESERVOIR These processes include fracturing in both the Creta-
ceous and the Miocene deformations, pressure solu-
QUALITY tion, coarse-crystalline late dolomitization, and
dissolution. The most important of these tectonodiage-
The Mardin Group carbonates in the study area netic events on reservoir quality are fracturing and late
have undergone both early and late diagenesis as the dolomitization.
result of three shallowing-upward carbonate cycles Visual analyses of cores, thin-section analyses, and
(Figure 3) and burial-tectonic stresses. The effects of porosity-permeability-fluid saturation-mercury injec-
these three shallowing-upward cycles on the reservoir tion measurements of both the Karababa-C member
quality are: (1) the creation of three carbonate reser- and Derdere Formation from wells such as Cendere-9
voirs separated from each other by source and/or seal allow the following generalizations for diagenesis,
carbonates, (2) the development carbonate rock types reservoir quality, and burial history: (1) the Karababa-
containing intergranular porosity at the upper parts of C member is generally tight and vertically fractured,
each cycle, (3) subaerial exposures and consequent with some fractures still open. No visible porosity is
karstification, desiccation cracks, and early dolomiti- observed in cores and thin sections. It may be a sec-
zation. Burial-tectonic stresses, which are superim- ondary reservoir, when associated with fault zones. (2)
posed on the previous events, caused the initiation The upper parts of the Derdere reservoir are generally
and development of tectonodiagenetic processes. limestone. Toward the middle part, it changes to

Figure 6.
Photomicrograph of
phosphatic mudstone
from the Karabogaz
formation. Dark colors
are phosphate in dark
matrix, which is rich in
organic matter. Width
of the field: 3.6 mm.
Plane-polarized light.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 177

Table 2. Source Rock Characteristics of the porosity is controlled by dolomitization (Figure 9),
Phosphatic Mudstone of the Karabogaz Formation dissolution (Figure 10), brecciation (Figure 11), and
in the Oil Fields Studied, Southeast Turkey.* fractures (Figure 9). These parameters that affect poros-
ity are very intense near fault zones, at the crest of folds,
TOC HI Tmax and/or on the maximum curvature of folds. The porosity
(wt. %) (mgHC/grock/wt. %) (°C) resulting from the above causes is secondary in type.
Decreasing order of abundance of this secondary poros-
1–2 350–450 440 ity is intercrystalline (Figures 7, 9), vuggy (Figure 10),
and fracture (Figure 9) types. There is dissolution in
*Values indicate type II kerogen and moderate maturity. From dolo-intercrystalline pores, as indicated by embayments,
Soylu and Gurgey, 1991.
irregular shapes, and enlargements along the dolomite
dolomite. Homogeneous distribution of porosity and crystal faces (Figure 10). The product of this dissolution
dolomitization is not observed on interwell scale. is the vuggy type of porosity.
However, there is a definite relationship between The association of late dolomitization with live
fracturing and dolomitization (Figure 9). Most of the hydrocarbon staining and fracture is shown in Figure
Figure 7. Photomicrograph of
the dolomite reservoir of the
Derdere Formation. This is the
main reservoir for the Mardin
Group carbonates and has an
excellent intercrystalline
porosity. Dark colors between
crystals are hydrocarbon
stains. Width of the field: 2.4
mm. Plane-polarized light.

Figure 8. Core photography of


fractures. The interconnection
of the reservoir intervals is
made by these fractures.
The vertical fractures are
dominant. Along these
fractures, some diagenetic
processes, such as dolomitiza-
tion, calcification, and
dissolution, exist. Scale is
5 cm. Karakus-2 well, depth of
2692.5m, Karababa-C
formation.
178 Uygur et al.

Figure 9. Photomicrograph
of dolomite and its association
with fracture. Coarse-
crystalline dolomites are
along the fracture, which
extends from the lower right
to the upper right. Dark areas
in intercrystalline pores are
hydrocarbon stains. Width
of the field: 1.40 mm. Plane-
polarized light. Karakus-14,
Derdere Formation.

12. The fracture in the center of the photomicrograph G. Karakus oil fields of the foreland and the
indicates that the late dolomitization and live hydro- Bölükyayla and Cukurtas oil fields of the overthrust
carbon staining along the fracture are associated with frontal zone are given in Table 3.
each other. The connections of the Mardin Group In the foreland, the oil is produced dominantly
reservoirs with each other are conducted by these from the Upper Cretaceous fractured dolomites and
permeable fractures (Figure 12). secondarily from the fractured limestones at a mini-
mum subsea depth of 1600 m. The combined dimen-
sions of the fields are ~l5 × 7 km and they are
RESERVOIR CHARACTERISTICS OF structurally characterized by northeast-southwest–
THE MARDIN GROUP CARBONATES trending low relief en echelon folds and faulted folds
(Figure 5).
The representative values of the reservoir proper- The maximum height of oil column is ~250 m. The
ties of the Cendere, K. Karakus, Karakus, and approximate oil/water contact is at 1850 m subsea in

Figure 10. Photomicrograph


of dissolution in dolo-
intercrystalline porosity.
Irregular embayments along
the dolomite crystal faces
indicate the dissolution, which
may later produce vuggy-type
porosity. Dark areas in the
embayments are hydrocarbon
stains. Width of the field: 0.70
mm. Plane-polarized light.
Karakus-14, Derdere
Formation.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 179

Figure 11. Photomicrograph


of brecciation in the Derdere
formation. This is a fault-
related brecciation.
Dark spots and lines are
hydrocarbon stains. Width
of the field: 5.40 mm. Plane-
polarized light. Cendere-l.

the Karakus, K. Karakus, and Cendere oil fields. It is gross production of the 40 wells in the fields was
~1870 m subsea for the G. Karakus field. 72,000 bbl/day, with 50,000 bbl/day of net crude oil
The porosity and permeability of the system and production in October 1991, and 27,000 bbl/day in
the water saturation ranges of pay zones of the fore- February 1994.
land oil fields are 3–15%, 0.1–7000 md, and 20–50%, In the overthrust frontal zone, the representative
respectively. The reservoir temperature is between values of the reservoir properties of the Bölükyayla
200° and 250°F (93°–121°C). The initial reservoir pres- and Cukurtas oil fields are different from the oil fields
sure ranges from 2600 psig to 3542 psig, with an aver- of the foreland area (Table 3). Here, the maximum
age of 3056 psig. Oil gravity ranges between 25° and height of the oil column (120 m) is less, the oil-water
32° API, with an average of 29° API. The gross pro- contact (–1970 ft; –2250 m) is deeper, the system’s per-
duction of each well ranges from 200 to 5,000 meability (0.1–65 md) is smaller, the reservoir temper-
bbl/day, with an average water cutoff of 30%. The ature (220°–270°F) is higher, maximum net crude oil

Figure 12. Photomicrograph


of the relationship
between the late dolomite,
hydrocarbon stain, and
fracture. Coarse-crystalline
dolomites and hydrocarbon
stain (dark line in the center)
along the fracture. Width
of the field: 5.40 mm. Plane-
polarized light. Karakus-14,
Derdere Formation.
180 Uygur et al.

Table 3. Reservoir Parameters and Variations Extracted from DSTs and Log Analysis in the Foreland and the
Overthrust Frontal Zone Oil Fields, Southeast Turkey.

Overthrust
Parameters Foreland Frontal Zone
Dimension (km) 15 × 7 8×7
Max. Height of Oil Column (m) 250 20
Oil-Water Contact BSL (m) –1850, –1870 –1970, –2250
Pay Zone Fractured Fractured
Dolomite (LST) Dolomite (LST)
φ (%) 3–15 1–7
K (md) 0.1–7000 0.1–65
Sw (%) 20–50 0.0–30
Reservoir Temperature (˚F)(˚C) 200–250 (93–121) 220–270 (104–132)
Initial Reservoir Pressure (psi) 2600–3542 2900–4050
Oil Gravity (˚API) 25–32 33–40
Max. Net Crude Oil 50,000 1050
Production (bbl/day)

Figure 13. The plot of


depth vs. the average
system’s permeability
for 141 measurements
in the Cendere,
K. Karakus, Karakus,
G. Karakus, Bölükyayla,
and Cukurtas oil fields.
High variability in
permeability values is
observed. Permeability
measurements are from
drill-stem tests of the
Mardin Group carbonates.
BSL = below sea level.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 181

Figure 14. The plot of depth vs. the system’s average permeability for 43 measurements in the Cendere, K.
Karakus, Karakus, and G. Karakus oil fields. Permeability measurements are from drill-stem tests of the
Derdere Formation. BSL = below sea level; R, correlation coefficient, = –0.4533; –x, mean depth BSL, –1815.98 m;
–y, mean permeability, = 12.59 md; and Y = 15.91 – 0.0082X is the regression equation.

Figure 15. Photomicrograph of


coarse dolomite crystals (on
the left) and fine dolomitic
crystals (on the right). Coarse
dolomites are in association
with thin fracture from lower
center toward upper left. Dark
colors along the fracture are
hydrocarbon stains. Width of
the field: 5.40 mm. Plane-
polarized light. K. Karakus-l,
Derdere Formation.
182 Uygur et al.

Figure 16. Photomicrograph


of heterogeneity in the
amount of dolomite.
Dolomite crystals are more
in the left than in the right
side. Intense dolomitization
(late) along the fractures.
Width of the fields: 5.40
mm. Plane-polarized light.
Karakus-14, Derdere
Formation.

Figure 17. Structural cross section through the K. Karakus and Cendere fields. Karakus-2 and -4 are structural-
ly higher than the Cendere-2 and -1 wells (see Figure 5 for location of cross section).
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 183

Table 4. Comparison of Carbonate Textures, Porosity, and Diagenesis for the Cendere and N. Karakus Oil
Fields of the Foreland Area.

Well Name Karababa-C Member Derdere Formation


Cendere-1 Bioclastic Carbonate breccia
wackestone/mudstone Pressure solution
Pressure solution Fractured
Fractured φ**≤1%
φ*≤3%
Cendere-2 Bioclastic- Dolosparite, dolomitic
wackestone/packstone wackestone
Fractured Vertical extension and
φ**≤6% conjugate shear fractures
Intercrystalline, vuggy φ
φ**≤15%
N. Karakus-2 Dolomitic wackestone Dolosparite
Fractured Fractured
φ*≤1% Intercrystalline, vuggy φ
Intense dissolution
φ*≤20%
N. Karakus-1 Bioclastic-wackestone Dolosparite
Fractured Fractured
φ*≤3% Intercrystalline, vuggy φ
Moderate dissolution
φ*≤10%
* φ is from log.
** φ is from core.

Figure 18. Plot of depth below sea level vs. the average system’s permeability of Mardin Group car-
bonates in the Cendere and K. Karakus oil fields, foreland area.
184
Uygur et al.

Figure 19. Drill-stem tests indicated flow rates for the Cendere and K. Karakus fields, foreland area.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 185

Figure 20. Depth reference to sea level (RSL) vs. static reservoir pressure
for the Mardin Group reservoirs in the Karakus, K. Karakus, Cendere,
and G. Karakus fields of the foreland area and fields of the overthrust
frontal zone.
186 Uygur et al.

Figure 21. Static reservoir pressure contour map at the top of the Derdere Formation.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 187

Figure 22. Depth reference to sea level (RSL) vs. bottom-hole temperature
(BHT) for the Mardin Group reservoirs in the foreland and overthrust
frontal zone (OTFZ) oil fields.
188 Uygur et al.

Figure 23. Depth reference to sea level (BSL) vs. API oil gravity. R, correlation coefficient,
= 0.7597; –x, mean depth, = –1772.83 m; –y, mean API oil gravity, = 29.01° API; and Y = 57.477 +
0.016X is the regression equation.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 189

Table 5. Correlation Coefficients for API Oil Gravity vs. Depth in the Oil Fields of Foreland Area, Southeast
Turkey.

Field Name Number of Mean BSL Correlation


Samples Depth (m) Mean (API) Coefficient (R)
Cendere 9 –1759.9 29.73 0.9614
K. Karakus 13 –1772.1 29.07 0.7705
Karakus 44 –1766.8 29.12 0.7170
G. Karakus 5 –1842.2 26.87 0.6733
Total 71 –1772.8 29.01 0.7597

production (1050 bbl/day) is less, and the initial reser- in the left side of the photomicrograph is more than
voir pressure (2900–4050 psi) and API oil gravity in the right side. The intense dolomitization occurs
(33–35) are higher than those of the oil fields in the along the fractures. All these features indicate the
foreland (Table 3). heterogeneity in tectonic and diagenetic textures.

VERTICAL AND LATERAL VARIABILITY Structurally High Intervals


IN RESERVOIR QUALITY The reservoir performance, in the field scale, is gen-
erally good at the structurally high intervals. This is
The most important reservoir parameter of the largely due to increase in fractures and associated
systems is permeability, which ranges from 0.1 to dolomitization, and better permeability and porosity
7000 md. The variable average production rate of at the crest of the folds where the fracturing is intensi-
each well is between 200 bbl/day and 5000 bbl/day. fied. For example, if one compares the reservoir para-
These values indicate that there is a great variability meters of the K. Karakus-2 and -4 wells, which are
in the reservoir quality of the Mardin Group carbon- structurally located at the higher position with respect
ates in the study area. On the other hand, there are to the Cendere-2 and -1, which are in the lower posi-
no main vertical or lateral changes in lithology, tion (Figure 17), one gets better parameters at the
facies, and thicknesses of the three shallowing- structurally higher intervals and/or on the maximum
upward cycles of the Mardin Group carbonates on structural curvature than at the lower intervals and/or
the interwell scale. Therefore, the variations in per- at the flanks of the fold.
meability and production rates are related to the dia- Table 4 shows the comparisons of carbonate tex-
genesis, including fractures of the Mardin carbonate tures, porosity types, and diagenesis for structurally
reservoirs. high (K. Karakus field) and low (Cendere field) areas.
Figure 13, which is the plot of depth below sea The best place in terms of reservoir quality is the
level vs. average system’s permeability of both the location of the K. Karakus-2 well. At this location,
Karababa-C member and the Derdere Formation of which comprises the highest point in the area, the
the foreland and overthrust frontal zone oil fields, intensity of dolomitization, dissolution, and the
shows the very high variability in permeability values. amount of secondary porosity of the Derdere Forma-
The similar plot for the Derdere Formation of the fore- tion are more than the intensity of dolomitization,
land oil fields (Figure 14) also shows a similar result. dissolution, and amount of secondary porosity of the
Although the relationship between depth and perme- Derdere reservoir in Cendere-1 and -2 and K.
ability seems to be clearer for the Derdere Formation, Karakus-1 locations (Table 4).
statistically it is still insignificant, with a correlation Similarly, the depth (below sea level) vs. the aver-
coefficient of –0.4533. age system’s permeability plot shows that the per-
In general, permeability depends upon the Creta- meability values (Figure 18) at the structurally high
ceous and the Miocene fracture systems and dolo- intervals (K. Karakus field), especially in the K.
intercrystalline porosity. The intensity of fracturing Karakus-2 well, are better than those at the struc-
and dolomitization in the area are heterogeneous turally low intervals (Cendere field). Consequently,
and/or heterogeneity in tectonic and diagenetic tex- the flow rates (in barrels per day per meter) of the
tures exists. This heterogeneity, in microscale, is structurally high intervals are better than those of
shown in Figures 15 and 16. Figure 15 shows that dif- structurally low intervals (Figure 19).
ferent sizes of dolomite crystals exist in the left and
right sides of the photomicrograph. Coarse dolomite
Static Reservoir Pressure
crystals in the left side of the photomicrograph are
associated with a thin fracture from the lower left The static reservoir pressures of the Mardin Group
toward the upper right. Figure 16 illustrates the het- carbonates in the foreland and overthrust frontal
erogeneous distribution of dolomitization that was zones, from 257 drill-stem tests, are plotted against
initiated by fractures. The amount of dolomitization depth with reference to sea level (Figure 20). This plot
190 Uygur et al.

Figure 24. Depth reference to sea level (BSL) vs. API oil gravity of foreland and overthrust frontal zone
oil fields.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 191

Figure 25. API oil gravity contour map at the top of the Derdere Formation.
192 Uygur et al.

Figure 26. Computed subsidence history with water depth, Bölükyayla-2 well.

indicates a good correlation between static reservoir Bottom-Hole Temperatures and


pressure and depth: the pressure increases with Geothermal Gradients
increasing depth. Almost all the data points fall on the
The corrected bottom-hole temperatures (BHT) of
same trend, with a correlation coefficient of 0.65. Some
the Mardin Group reservoirs in both the foreland and
discrepancies in the data points are due to the exis-
the overthrust frontal zones are plotted against subsea
tence of more than one production zone, ongoing pro-
depth (Figure 22). This figure shows a positive correla-
duction, and consequent pressure drops or the effects
tion between BHT and subsea depth; that is, BHT
of interference between wells in the fields. If these dis-
increase with increasing depth.
crepancies are omitted, the correlation coefficient
The detailed investigation of the plot, however,
would be >0.65. This indicates that the same and/or
exhibits two geothermal gradients. The geothermal
similar hydraulic regimes prevail in the Mardin Group gradient of the overthrust frontal zone in the north-
reservoirs of both regions due to interconnection west (0.89°F/100 ft; 1.30°C/100 m) is smaller than
between reservoir intervals by means of tectonically the geothermal gradient of the foreland zone
controlled fractures and dissolution voids. (2.15°F/l00 ft; 3.9°C/100 m) in the southeast. This
The static pressure contour map at the top of the means that the Mardin Group reservoir in the over-
Derdere Formation (Figure 21) indicates lower pres- thrust frontal zone is cooler than the Mardin Group
sure values at the structurally high intervals than at reservoir in the foreland area. This is due to the pres-
the structurally low intervals. This is obtained by com- ence of thicker high-thermal-conductivity carbon-
paring the structural contour map in Figure 5 with the ates in the overthrust frontal zone and the movement
static pressure contour map in Figure 21. The static of the formation water from the deeper intervals in
pressure contours of the Karakus, K. Karakus, Cen- the overthrust zone to the shallower depths in the
dere, and G. Karakus oil fields of the foreland and the foreland zone. There are also areas with meteoric
Bölükyayla and Cukurtas oil fields of the overthrust water recharge.
frontal zone are almost a mirror image of the structure Since the oil fields of the foreland area have rela-
contours expected in any hydrostatic to slightly sub- tively higher geothermal gradients than the oil
normally pressured system. fields of the overthrust frontal zone, the foreland
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 193

Figure 27. Computed subsidence history with water depth, Karakus-3 well.

area is expected to be a better site for hydrocarbon The API gravity of oil also decreases gradually from
accumulation due to active water drive mecha- the overthrust frontal zone in the northwest to the
nism. However, this hypothesis must be checked foreland area in the southeast (Figure 25). The cause of
by other parameters, such as salinity and hydraulic this, however, is still being studied, and it may be that
head distributions within the Mardin Group reser- either “heavy oils are generated away from the Thrust
voirs. Belt (Upper Cretaceous) and lighter oils are generated
close to the Thrust Belt” (Yükler, 1991) or short-
API Oil Gravity distance migrations occur from the adjacent low areas
toward the structurally high areas.
In the foreland area, the API gravity of oil
decreases with increasing depth (Figure 23 ) in each
of the Cendere, K. Karakus, Karakus, and G. Karakus BASIN MODELING AND RESULTS
oil fields. In other words, oil gets heavier with
increasing depth. The correlation coefficients of API The deterministic dynamic basin modeling software
gravity vs. depth for each of the fields studied range developed by Yükler (Yükler and Welte, 1980; Yükler,
from 0.6733 to 0.9614 (Table 5). The correlation coef- 1987), YUKLERPC®, is used to quantify the geologic
ficient for the total samples (n = 71) is 0.7597. This (lithofacies, chronostratigraphy, type of event, rate of
points to a gravitational segregation of oil within the sedimentation, and/or erosion and tectonic evolution),
reservoir and/or migration of hydrocarbons with temperature, pressure, and hydrocarbon generation,
strong water flow from the deeper overthrust frontal migration, and accumulation histories. Figures 1 and 5
zone where hydrocarbons are generated to the shal- show the study area and the wells that are modeled. In
lower zones in the foreland zone. this chapter, we present only the subsidence and
There is no relationship between the API gravity hydrocarbon generation histories at two wells:
of oil and depth in the oil fields of the overthrust Bölükyayla-2 in the overthrust frontal zone and
frontal zone. This is due to the scarce data points in Karakus-3 in the foreland area (Figure 5).
this area. Figure 24 shows the relationships of the In the basin modeling, the sedimentary sequences
API gravity of oil and depth in both regions. at the wells are subdivided into 40 layers from the
194 Uygur et al.

Figure 28. Computed temperature and hydrocarbon generation (type II) at present, Bölükyayla-2 well.

Aptian (125 Ma) to the present and have vertical conti- temperature data are used, and there is good agree-
nuity in lithology and lateral continuity in time. The ment between the computed and measured values.
vertical continuity is essential to correctly compute The computed hydrocarbon generation vs. depth
pressure and temperature histories. The lateral conti- trends (Figures 28, 29) are then compared against all
nuity in time is needed to accurately define chrono- available pyrolysis data and hydrocarbon occurrences
stratigraphy and to plot the results at any time during at the wells. At the Bölükyayla-2 well, the organic mat-
the geologic history (Yükler and Welte, 1980). ter is immature down to 1186 m, in early oil generation
Figures 26 and 27 illustrate the subsidence histories phase (10°–14° API) down to 1828 m, at main oil phase
at the Bölükyayla-2 and Karakus-3 wells, respectively. (15°–28° API) down to 3015 m (peak oil generation at
The sedimentation rates were slow to moderate until 3015 m) and at condensate phase (30°–50° API) down
the late Campanian (75 Ma). With the thrusting during to T.D. (Figure 28). At the Karakus-3 well, the organic
the latest Campanian–earliest Maastrichtian (75–73 matter is immature down to 1805 m, at early oil gener-
Ma), the subsidence rate was very high at the ation phase down to 2689 m, and at main oil phase
Bölükyayla-2 well (Güven et al., 1991), followed by down to T.D. (Figure 29).
slow to moderate rates until the early Eocene (54 Ma) Figure 30 shows the hydrocarbon generation his-
and higher rates until the Oligocene (35 Ma) (Figure tory for the two potential source rocks, Derdere and
26). At the Karakus-3 well, the subsidence rate was Karabogaz formations, at the Bölükyayla-2 well. The
high from the earliest Campanian to the Oligocene, Derdere Formation reached peak oil generation at
which is typical of a foreland area (Figure 27). The 43.6 Ma and the Karabogaz Formation at 37.4 Ma.
study area was then subjected to erosion from the Since oil generation rates (and heating rates) were
Oligocene to Middle Miocene (35–16 Ma), followed by low within the area, expulsion efficiency for oil was
another deposition until the end of Miocene (5 Ma) also low. Therefore, oil expulsion took place when
(Figures 26, 27). During the last 5 Ma, the study area the source rocks entered the light oil/condensate
was uplifted and the Selmo Formation was partly or generation phase. Consequently, the oils generated
totally eroded. and expelled in the overthrust frontal zone found
In the optimization of the computed pressure and within the area have API gravities that are ≥30˚ API.
temperature values, all the available maturity and The modeling results show that expulsion of oils in
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 195

Figure 29. Computed temperature and hydrocarbon generation (type II) at present, Karakus-3 well.

the overthrust frontal zone started as early as 45 Ma Fracturing and late dolomitization, and the rela-
and continued until the Late Miocene (6 Ma). The tionships of depth vs. static reservoir pressure, API
Miocene thrusting also had a major influence on oil gravity variations, and bottom-hole tempera-
generation and migration of oils. Both of the poten- tures, indicate that the Cendere, K. Karakus,
tial source rocks at the Karakus-3 well are still at Karakus, G. Karakus, Bölükyayla, and Cukurtas
early oil generation phase (Figure 31). No effective Mardin Group reservoirs behave as a single, contin-
oil expulsion and migration can take place at this uous reservoir. A similar increase in static reservoir
location. pressure with depth is also an indication that similar
hydraulic regimes exist in Mardin Group reservoirs
CONCLUSIONS of both the foreland area and the overthrust frontal
zone. Increased intensity of fracturing, dolomitiza-
The petrophysical and petrographic properties of tion, dissolution, and better permeability and poros-
the three carbonate reservoir intervals within the ity at structurally high intervals exert the main
Mardin Group are evaluated to improve the explo- control over the reservoir quality. Heterogeneity in
ration success in southeast Turkey. fracturing and dolomitization must be studied in
Three carbonate reservoir intervals were initially detail by means of oriented cores, thin sections, elec-
separated by source and/or seal carbonate rocks. tron scanning images, and interference tests for
Those reservoir intervals may now be interconnected future prospect areas.
with each other by tectonically controlled fractures The computed geothermal gradients show that
and dissolutions. The present configuration of the gradients in the overthrust frontal zone
source reservoir and seal intervals is the result of the (0.89°F/100 ft; 1.30°C/100 m) is lower than the fore-
combination of carbonate shallowing-upward cycles land area (2.15°F/100 ft; 3.9°C/100 m). This suggests
and burial-tectonics stresses, which initiated and meteoric water recharge in the former and discharge
developed the tectonodiagenetic processes and in the latter zones. The active water circulation
superimposed on those carbonate shallowing- makes the foreland area a better site for hydrocarbon
upward cycles. accumulation.
196 Uygur et al.

Figure 30. Computed hydrocarbon generation history (type II) for the two potential source rocks, Derdere and
Karabogaz formations, Bölükyayla-2 well.

The basin modeling results show that oil generation TPAO for the discussions we had regarding this study.
from the two potential source rocks (Derdere and Ihsan Donmez and Cem Pasin of TPAO Drafting
Karabogaz formations) in the overthrust frontal zone Department deserve special thanks for drawing the
started with the late Cretaceous thrusting. Expulsion figures.
of oils, however, started toward the mid-late Eocene
time. Oil expulsion occurred at peak oil and/or con-
densate phase. Hydrocarbon generation in the fore- REFERENCES CITED
land area is directly related to the thickness of the Akgul, A., and L.T. Erdogan, 1992, Factors affecting
Paleocene and Eocene sediments. Some of the sub- hydrocarbon generation in Southeast Turkey, with
basins are still at early oil generation phase, whereas examples from the Bölükyayla and Karakus oil
others are approaching peak oil generation. fields, in A.M. Spencer, ed., Generation, accumula-
Both the type of hydrocarbons generated in these tion and production of Europe’s hydrocarbons: Spe-
two zones and the active water from the overthrust cial Publication of the European Association of
frontal zone to the foreland area resulted in a gradual Petroleum Geoscientists 2: Oxford, Oxford Univer-
decrease in API gravities of oils in the traps from the sity Press, p. 231–236.
overthrust frontal zone to the foreland area. Brown, S.A., 1991, New structural insight in southeast
Turkey (abs.): Ozan Sungurlu Symposium, Novem-
ACKNOWLEDGMENTS ber 26–28, 1991, Ankara, Turkey, p. 29–30.
Celikdemir, E.M., S. Dulger, N. Gorur, C. Wagner, and K.
We thank the TPAO Exploration Group for support Uygur, 1991, Stratigraphy, sedimentology and hydro-
and permission to publish this work. The help of carbon potential of the Mardin Group, Southeast
Vedat Aydemir and Hamdi Gungor of the TPAO Turkey, in A.M. Spencer, ed., Generation, accumu-
Petrophysical Evaluation Unit in extracting the reser- lation and production of Europe’s hydrocarbons:
voir parameters and in drawing some of the diagrams Special Publication of the European Association of
is appreciated. We would like to thank Ahmet Dinçer, Petroleum Geoscientists 1: Oxford, Oxford Univer-
Bayram Saritas, Samim Tatli, and Adnan Cetin of sity Press, p. 439–454.
Reservoir Characterization of Cretaceous Mardin Group Carbonates in Bölökyayla-Cukurtas and Karakus Oil Fields 197

Figure 31. Computed hydrocarbon generation history (type II) for the two potential source rocks, Derdere and
Karabogaz formations, Karakus-3 well.

Dewey, J.F., W.C. Pitmann III, W.B.F. Ryan, and J. Bon- Ozan Sungurlu Symposium, November 26–28,
nin, 1973, The plate tectonics and the evolution of 1991, Ankara, Turkey, p. 91–92.
the Alpine system: Geological Society of America Pasin, C., 1991, Karakus ve Cendere sahalari Mardin
Bulletin, v. 84, no. 10, p. 3137–3180. Grubu üstü yapy kontur haritasi: TPAO Arama
Duran, O., and M. Aras, 1990, Yeniköy petrol sahasi Grubu.
(G.D. Anadolu) Mardin Grubu karbonatlari cokel Perinçek, D., and I. Çemen, 1991, Structural highs in
fasiyes ve rezervuar özellikleri: Türkiye 8, Petrol the forelands of the southeastern Anatolia fold-
Kongresi Jeoloji Bildirileri cildi, p. 272–280. thrust belt—their age, geometry and tectonic im-
Gorur, N., E. Celikdemir, and S. ve Dulger, 1987, plication (abs.): Ozan Sungurlu Symposium,
Guneydogu Anadolu X, XI ve XII. Petrol bol- November 26–28, 1991, Ankara, Turkey, p. 50.
gelerinde Mardin Grubu karbonatlarinin sedi- Soylu, C., and K. ve Gurgey, 1991, Adiyaman bol-
mentolojisi, yayilimi, fasiyes, cokelme ortami ve gesinde Kretase yasli birimlerin petrol turum
paleocografya: TPAO Report no. 2321. potansiyelleri: TPAO Report no 1945.
Güven, A., A. Dinçer, M.E. Tuna, and T. Çoruh, 1991, Tardu, T., Y. Akcay, and L. Is, 1990, GD Anadoluda
Guneydogu Anadolu Kampaniyen Paleosen Otok- secilmis bazi stratigrafi birim ve birliklerinin sismik
ton istifinin stratigrafisi: TPAO Report no. 2828. stratigrafik analizi (abs.): Turkey 8, Petrol Kongresi,
Hobbs, B.E., W.D. Means, and P.F. Williams, 1976, An Genisletilmis Bildiri Ozetleri, p. 51–52.
outline of structural geology: New York, John Wiley Torre, J., E. Memioglu, and S.C. Bakiler, 1992, Injection
and Sons, p. 383–387. revives Turkey’s low-energy wells: Middle East
Horstink, J., 1971, The Late Cretaceous and Tertiary Well Evaluation Review, no. 12, p. 44–48.
geological evolution of eastern Turkey: Türkiye l, Uygur, K., and M. Gurel, 1991, Petrophysical evalua-
Petrol Kongresi, p. 25–41. tion and comparison of Mardin Group carbonates
Ozkanli, M., and E. Standen, 1991, A study of frac- in Karakus, K. Karakus-Cendere, and G. Karakus
ture morphology from borehole image data in oil fields of SE Turkey: Ozan Sungurlu Symposium,
Karakus, Cendere, Ozan Sungurlu fields (abs.): November 26–28, 1991, Ankara, Turkey, p. 1–28.
198 Uygur et al.

Uygur, K., and V. Aydemir, 1988, Bölükyayla- wrench tectonics: AAPG Bulletin, v. 57, p. 74–96.
Cukurtas sahalarinda (XII. Bolge) Derdere, Yilmaz, Y., E. Yigitbas, and C. Genc, 1991, A new
Karababa, Karabogaz ve Sayindere formasyonlar- approach on the geological evolution of SE Anato-
inin yeralti jeolojisi; Petrografi, sedimentoloji, ortam lian orogenic belt (abs.): Ozan Sungurlu Sympo-
analizi, petrofizik ve goreceli olay kronolojisi: sium, November 26–28, 1991, Ankara, Turkey,
TPAO Report, p. 2554–259. p. 42
Wagner, C., and M. Pehlivan, 1985, Karst-geological Yükler, M.A., 1987, How essential is quantitative basin
interpretation of the Mardin carbonates in the Cem- modelling in petroleum exploration?: 7th
berlitas field: A pilot study: Unpublished TPAO Petroleum Congress of Turkey, Proceeding: Geol-
Report no. 2051. ogy, p. 392–404.
Wagner, C., and M.E. Tuna, 1988, Campanian cycle IV Yükler, M.A., 1991, The effects of quantitative basin
carbonates in southeast Turkey: depositional envi- modelling application in petroleum exploration in
ronment and paleogeography: Unpublished TPAO Turkey (abs.): Ozan Sungurlu Symposium, Novem-
Report no. 2528, 11 p. ber 26–28, 1991, Ankara, Turkey, p. 21–22.
Wagner, C., M. Pehlivan, and C. Soylu, 1986, Oil habi- Yükler, M.A., and D.H. Welte, 1980, A three-dimen-
tat of the Adiyaman area—southeast Turkey: A sional deterministic dynamic model to determine
joint geological geochemical study: TPAO Report geologic history and hydrocarbon generation,
no. 2139. migration and accumulation in a sedimentary
Wilcox, R.E., T.P. Harding, and D.R. Seely, 1973, Basic basin: Fossil Fuels, Editions Technip, p. 267–285.
Surdam, R.C., Z.S. Jiao, and H.P. Heasler, 1997, Anomalously
pressured gas compartments in Cretaceous rocks of the
Laramide Basins of Wyoming: a new class of hydrocarbon
accumulation, in R.C. Surdam, ed., Seals, traps, and the
petroleum system: AAPG Memoir 67, p. 199–222.
Chapter 12

Anomalously Pressured Gas Compartments


in Cretaceous Rocks of the Laramide Basins
of Wyoming: A New Class of Hydrocarbon
Accumulation
R.C. Surdam
Z.S. Jiao
H.P. Heasler
Institute for Energy Research, University of Wyoming
Laramie, Wyoming, U.S.A.

ABSTRACT
Cretaceous shales in the Laramide basins of Wyoming (LBW) below
~8000–9000 ft (2440–2740 m) typically are anomalously pressured. In the
basin centers, the top 1000–2000 ft (305–610 m) of the anomalously pressured
zone is transitional and typically occurs within upper Cretaceous shales
(Steele, Cody, or Lewis). The overpressured zone [~2000 ft (610 m) thick]
persists down to the lowermost organic-rich Cretaceous shale. Typically, the
rocks below these shales in the Powder River and Wind River basins are nor-
mally pressured. The top of the anomalously pressured zone is identified by
marked increases in sonic transit time, hydrocarbon production index, clay
diagenesis (smectite to illite), and vitrinite reflectance. Many of the affected
organic-rich shales are characterized by bitumen-filled microfractures.
In the LBW, the major difference between pressure compartmentalization in
Cretaceous sandstones and shales is one of scale. The overpressured
Cretaceous shales in each of the basins comprise a basinwide, dynamic pres-
sure compartment. In contrast, the Cretaceous sandstones within each basin are
subdivided stratigraphically and diagenetically into relatively small, isolated
pressure or fluid-flow compartments [largest dimension 1–10 mi (1.6–16 km)]
within the shale section.
The driving mechanism of pressure compartmentalization in both the shales
and sandstones is the generation and storage of liquid hydrocarbons that sub-
sequently partially react to gas, converting the fluid-flow system from a single-
phase regime to a multiphase regime in which capillarity controls permeability.
In a single-phase, water-dominated system, internal and external stratigraphic
elements (ranging from paleosols along unconformities to transgressive shales)
act as low-permeability rocks with finite leak rates. These elements evolve dia-
genetically during progressive burial (smectite altering to illite; kaolinite to

199
200 Surdam et al.

chlorite). As more liquid hydrocarbons are generated and the oil-to-gas reac-
tion proceeds, the system becomes saturated with hydrocarbons. This results
in the expulsion of free water and greatly increased displacement pressures,
which cause the low-permeability elements acting as fluid-flow barriers to
form capillary seals. Three-dimensional closure of the seals results in fluid
compartmentalization and anomalous pressure and, thus, the formation of
an anomalously pressured gas accumulation. In the sandstones, three-
dimensional closure of capillary seals above, below, and within a sandstone
results in isolated fluid-flow or pressure compartments within the sand-
stone. In a few cases, faulting bounds sandstone compartments by emplacing
low-permeability carbonate- or quartz-cemented rocks adjacent to them.

INTRODUCTION trend of sonic transit time occurs at ~8000–9000 ft


(2440–2740 m) depth in the Cretaceous shales in the
The aim of this chapter is to document the pressure central portions of the Powder River, Bighorn, and
regimes characterizing Cretaceous shales and sand- Wind River basins (Figure 3A–C). This increase in
stones in the Laramide basins of Wyoming (LBW) sonic transit time marks the top of a continuous, over-
(Figures 1, 2). This paper consists of (1) evaluation of pressured shale section that begins at 8000–9000 ft
the regional pressure regime in Cretaceous shales, (2440–2740 m) depth in these three basins and persists
with emphasis on the causes and timing of overpres- down to 10,000–11,500 ft (3050–3505 m) depth in the
suring; (2) evaluation of the pressure regime within Powder River and Wind River basins (Figure 3A, C).
Cretaceous sandstones, with emphasis on the mecha- Due to drilling patterns, the base of the regionally
nisms causing anomalous pressures within reservoir overpressured shale section in the Bighorn Basin has
sandstones; (3) discussion of the origin of anomalous not been determined (Figure 3B). In the Washakie
pressures in the LBW; and (4) discussion of the evolu- Basin (Figure 3D), the top and bottom of the overpres-
tion of the pressure regimes and the formation of sured shale section occur at depths of 10,000 and
anomalously pressured gas accumulations during 15,000 ft (3050 and 4570 m), respectively.
either continuous subsidence or subsidence and subse- Mud weights and measured pressures of thin
quent uplift and erosion. sandstones interbedded with the Cretaceous shale
section also indicate that this section in the LBW is
typically overpressured. During drilling, mud
REGIONAL ANOMALOUS PRESSURE weights were typically adjusted from 8.4 to ~9.0 ppg
REGIME IN THE UPPER CRETACEOUS at the top of the Cretaceous shale section, and from
SHALES 9.0 to ≤15 ppg within the section. A mud weight of
8.4 ppg corresponds to a pressure gradient of 0.436
Delineation of Anomalous Pressure psi/ft, and a mud weight of 14.8 ppg corresponds to
Using Sonic Logs a pressure gradient of 0.77 psi/ft; a gradient of
~0.433 is considered normal. Rocks occurring strati-
Determining trends in acoustic transit time (µs/ft) graphically above and below the overpressured Cre-
on sonic logs is a conventional method of detecting taceous shale section in the LBW are generally
anomalous pressures in shales (Hottman and John- characterized by normal pressure (i.e., they have a
son, 1965; Magara, 1976; Powley, 1982). Powley (1982, pressure gradient of ~0.433 psi/ft). Shale displace-
personal communication) cites many examples of ment pressures measured in the laboratory follow a
pressure compartments delineated on the basis of gradient parallel to the regional lithostatic gradient,
analyzing trends in sonic transit time from sonic logs. but offset to lower pressures (Figure 4A, B).
Because the velocity of a compressional wave in sedi- The sonic anomaly panels shown in Figure 5A–D
mentary rock is dependent on the effective rock stress, are a more detailed representation of the pressure
an increase in transit time is interpreted as indicating regime of the Cretaceous shales in the LBW. They
a lower effective rock stress and, therefore, overpres- were constructed from decompacted sonic transit-
suring. Almost all sonic logs from the LBW exhibit a time profiles through the LBW in two steps: (1) pan-
marked increase in transit time (decrease in velocity) els of sonic logs along cross sections across each
in the deeper portions of the basins (Figure 3A–D). basin were filtered using gamma-ray logs so that
Downhole trends in digitized sonic logs from the only fine-grained clastic lithologies were repre-
LBW (Figure 3A–D) indicate that a reversal in the sented, and the panels were used to construct sonic
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 201

Figure 1. Index map of the


WYOMING Laramide basins of
45
Wyoming (LBW). The solid
lines are the lines of the
sonic log panels in Figure 3.
Powder
Bighorn River
44 Basin
Basin

Wind River
Latitude

43 Basin

Green
42 River Washakie
Basin Basin

50 miles

41
111 110 109 108 107 106 105
Longitude

velocity profiles; and (2) the velocities from the sonic ANOMALOUS PRESSURE REGIMES
log panels were corrected for compaction (i.e., they IN THE UPPER CRETACEOUS
were decompacted) by using an exponential function
constrained by the lowest observed transit-time SANDSTONES
value in the sonic logs from each basin (i.e., typi-
cally~60 µs/ft). After this two-step operation, the As indicated by measured pressures from drill-stem
sonic transit-time anomaly or anomalies were iso- tests (DSTs) and repeat formation tests (RFTs), the
lated for each basin. Cretaceous sandstones of the LBW at a present-day
Several observations can be made from the sonic depth of ~8000–9000 ft (2440–2740 m) are anomalously
anomaly panels (Figure 5A–D). First, except at the pressured, tight gas sandstones (Figure 6). The pres-
basin margins of the LBW, the onset of overpressuring sure regime of Cretaceous sandstones in the LBW dif-
is marked by a transitional zone (slightly overpres- fers from the pressure regime characterizing the shales
sured rocks; mud weights >8.4 ppg) ~1000–2000 ft mainly in terms of scale. The Cretaceous sandstones
(305–610 m) thick; the top of the transitional zone in within the overpressured shale section are not part of a
the basin center is approximately horizontal. Sec- large, basinwide pressure compartment like the
ond, the highest pressures are found in the lower shales; rather, the fluid-flow and pressure systems of
half of the overpressured shale section; where individual sandstones are separated spatially, both
observed, the bottom of this section is regionally vertically and horizontally (Figure 6) (Heasler et al.,
coincident with the lowermost organic-rich shales in 1994). Even within specific sandstones (e.g., Dakota,
the Cretaceous. Finally, overpressuring within these Muddy, or Frontier), rocks are subdivided into rela-
shales develops in each basin as one large and fairly tively small, isolated compartments (Figure 6).
continuous volume of rock, or pressure compart- The pressure anomalies shown in Figure 7 are simi-
ment. The only exception occurs along the margins lar to a multitude of anomalies found in the Cretaceous
of the basins, where the Cretaceous section sandstones in the LBW and described in Heasler et al.
approaches the surface and where erosion (unroof- (1994). In each case, the flow anomalies on a potentio-
ing) has been greatest; the overpressured shale sec- metric surface map are characterized by steep sides and
tion at the basin margins is wedge shaped (Figures relatively flat tops. Heasler et al. (1994) interpret these
3, 5A, C). In the Powder River Basin, the base of the anomalies as pressure compartments isolated from sur-
Cretaceous shale section is parallel to the Fuson rounding rocks by the three-dimensional closure of
Shale, but the top of it cuts across stratigraphy (Fig- seals and characterized by insignificant fluid flow
ure 5A). within the compartments. The mechanism responsible
202 Surdam et al.

Figure 2. Stratigraphic nomenclature chart for the LBW. (Modified from the Wyoming
Geological Association Annual Field Conference Guidebook, 1993).

for such compartmentalization is discussed in the next gradient (0.9–1.0 psi/ft) parallel to the regional litho-
section. static gradient, but offset to much lower pressures
The Cretaceous sandstones of the LBW above and (Figure 4A, B). For the present discussion, this gradi-
below the pressure boundary exhibit very different fluid- ent is designated the “sandstone gradient.” These
flow characteristics (Figure 4A, B). Sandstones above the sandstones support the weight of the fluid column
pressure boundary typically follow a path parallel or down to the boundary, plus the weight of the rock and
subparallel to a hydrostatic gradient (~0.433 psi/ft); at fluid column from the pressure boundary down to the
any specific depth, these sandstones support only the depth of the sandstone. Figure 4A, B demonstrates
weight of the overlying fluid column. In contrast, sand- that the sandstones within the overpressured shale
stone reservoirs below the pressure boundary follow a section are not only isolated from the sandstones
Figure 3. Typical acoustic transit time vs. depth plots for the Powder River, Bighorn, Wind River, and Washakie basins.
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming
203
204 Surdam et al.

(a) Pressure (psi) (b) Pressure (psi)


0 2000 4000 6000 8000 10000
0 2000 4000 6000 8000 10000
0 0
"L
Reservoir Pressure ith Reservoir Pressure
os
Displacement Pressure ta Displacement Pressure
tic
"H y d

2000

0.4
"Li "
G
rost

33
tho ra
sta di
atic"

ps
tic en
"G t,

i/ft
rad 1.
Gra

4000 0
ien ps
di

t, 1 i/f
ent,

5000 .0 t
ps
0.

i/ft
433

6000

Depth (ft)
ps
Depth (ft)

i/ft

Top of Pressure Compartment "Gas"


Gradient Top of Pressure
Sh 8000
ale Compartment
"D Sh
isp Pr ale
lac Sa
10000 em es "
Sa en 10000 nd su Dis
nd tP st re pl
sto res on " ac
ne su e G e
re" Pr ra m
"F Gr es di en
rac su en t
tur ad 12000
e" ien re t
Gr t G
ad ra
ien di
t en
14000 t

15000

Figure 4. (A) Pressure profile for the major Cretaceous gas reservoirs in the Washakie Basin. (B) Pressure pro-
file for the major Cretaceous gas reservoirs in the Powder River Basin. Note that at the depth that compart-
mentalization is complete, the evolution of the pressure regime is significantly altered. Below the top of
overpressure, the change of reservoir pressures parallels the lithostatic gradient, but offset to lower pressures
than regional lithostatic gradient.

above the pressure boundary, but also from the adja- al., 1988; Martinsen, 1994). The Muddy Sandstone
cent shales below the boundary. along the west-east cross section shown in Figure 9 is
Sandstones within the overpressured Cretaceous externally sealed above and below by the Mowry and
shale section can be overpressured, normally pres- Skull Creek shales (transgressive shales), and is inter-
sured, or even underpressured. The pressure regime nally separated into flow compartments by imperme-
of the sandstones is dependent on whether, and how, able horizons within the sandstone (Figure 7;
the sandstones are connected to rocks outside the vol- Moncure, 1992). Note that along the cross section
ume of overpressured shales. Note that when under- shown in Figures 7 and 9, the inclination of the rocks is
pressured sandstone reservoirs are present in these 2° regionally, and there is no structural closure along
basins, they occur in the upper portion of the over- the section.
pressured shale section (Figures 5, 8). In Figure 8, the Figure 10 summarizes the relationship between
underpressured compartment is revealed by severe sandstone pressure compartments and the regional
lost circulation during drilling and the anomalous overpressuring of Cretaceous shales. The distribution
behavior of the acoustic transit time and resistivity. of sandstones and shales in the Wind River Basin in
Figure 9 illustrates detailed compartmentalization Figure 10 was determined from gamma-ray logs. Shal-
within an individual sandstone reservoir facies, in this lower, anomalously pressured sandstones are under-
case the Muddy Sandstone along a west-east cross sec- pressured.
tion from the Amos Draw to the Kitty to the Ryan field
to a wildcat well in the Powder River Basin. Along
this section, the Muddy Sandstone is part of a ORIGIN OF ANOMALOUS PRESSURES
shoreface/valley-fill depositional system in which the IN THE LARAMIDE BASINS
younger, valley-fill elements (Recluse, Cyclone, and OF WYOMING
Ute members) are separated from the older, shoreface
sandstones (Rozet/Lazy B members) by a paleosol Role of Disequilibrium Compaction
developed along a regionally prominent lowstand
unconformity, with intensive clay infiltration into the Many factors could have caused the marked increases
underlying sandstone (Odland et al., 1988; Wheeler et in sonic transit time observed in the Cretaceous shale
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 205

Powder River Basin Bighorn Basin


(A) W E (B) W E
2000

4000 le
ee 4000
St

a
ar
br
io
6000 6000
m

N
F

n
nce

so

Depth, ft
La

Fu
8000
Depth, ft

8000
10000

M
10000

or
12000

ris
on
hale
14000 ry S

Fm
12000 M ow
16000
14000
110 100 90 80 70 60 50 40 30 20 10 0 5 15 25 35 45

Distance, mi Distance, mi

(C) SW Wind River Basin NE (D) W Washakie Basin


E

M 4000
4000
or La
ris nc
on e 6000

AL
MO
8000

Depth, ft

ND
Depth, ft

8000

12000
10000

12000
16000

14000
20000
0 6 12 18 24 30 36 42 48 54 60 66 72 0 5 10 15 20 25 30 35 40 45 50 55 60 65
Distance, mi Distance, mi

Increasing Anomaly

Figure 5. Velocity/gas profiles after decompaction correction across the LBW, showing basinwide overpressure
compartmentalization in the Cretaceous shale section. Note that the basinwide overpressure compartments
have a wedge shape in the Powder River and Wind River basins. Selected formation boundaries are shown for
structural reference.

section in each basin of the LBW. These include poros- Basin, the total dissolved solids (TDS) in produced
ity, fluid type, effective rock stress, lithology, matrix waters average 10,000 ppm. The relatively constant
composition, salinity, isotherm, and bed thickness, and dilute TDS content of these waters indicates that
although variations in porosity and fluid properties (1) marine rocks in the Cretaceous section have been
are generally considered to be the most significant fac- strongly influenced by meteoric water influx from the
tors. Dramatic increases in sonic transit time like those basin margins and (2) the upper 8000–9000 ft
observed on sonic logs from the LBW (Figure 3A–D) (2440–2740 m) of section in each basin of the LBW is
are usually attributed to undercompaction, which is likely in hydrologic communication. The rapid and
associated with increased porosity. However, there is significant increase in TDS (to averages ~35,000 ppm)
no observed increase in shale or sandstone porosity in produced waters from below the pressure boundary
below the top of the overpressured Cretaceous shale strongly suggests that fluid compartmentalization has
section in the LBW (Table 1). In fact, the measured occurred within the basin, and that there is poor fluid
porosities of shales in the overpressured section in the communication between rocks above and below a
LBW range from only 2%–5%. Thus, as the Cretaceous present-day depth of 8000–9000 ft (2440–2740 m)
shale section in the LBW cannot be undercompacted, (MacGowan et al., 1994). In the Washakie Basin, the
some other factor must be responsible for the observed transition occurs at greater depth.
overpressuring. It is evident from petrographic, as well as geochem-
Variations in fluid properties are observed in the ical, studies that the overpressured Cretaceous shale
LBW, as is evidenced by the significant change in com- section is hydrocarbon saturated, while rocks above
position of the produced waters at the top of the over- and below it are not. For example, ultrathin sections of
pressured Cretaceous section (MacGowan et al., 1994). organic-rich Mowry Shale from the Powder River
For example, above this depth in the Powder River Basin have abundant microfractures and laminations
206 Surdam et al.

Figure 6. Contoured
pseudopotentiometric
surface computed from
drill-stem test data from
the Frontier Formation in
the Laramide basins of
Wyoming (see Figure 1
for basin names). The
sandstone pressured
compartments (red spots)
are isolated from
each other.

filled with bitumen (Figure 11) below the pressure reflectance/depth gradient (Figure 14) (Hunt, 1979;
boundary (≈8000 ft, 2440 m), but only a few thin, Law et al., 1989; Jiao, 1992; MacGowan et al., 1994);
widely separated fractures contain bitumen at shal- (4) a transition from 20% illite to 85% illite in mixed-
lower depths (≈5000 ft; 1520 m) (Figure 11). layer illite/smectite clays in the shale, and ordering of
Because the Cretaceous shale section below the the clay structure (Figure 15) (Jiao and Surdam, 1993);
sonic transit-time reversal in the LBW (1) does not and, most importantly, (5) a marked increase in the
show an increase in porosity, (2) is saturated with displacement pressure and sealing capacity of the Cre-
hydrocarbons and overpressured on a regional scale, taceous shales in a multiphase fluid-flow system (Fig-
and (3) exhibits regional trends in hydrologic commu- ure 16) (Schowalter, 1979; Jiao and Surdam, 1993).
nication (as determined from the composition of Thermal maturation—including generation but
produced waters), it is clear that disequilibrium com- only partial expulsion of oil and conversion of this
paction is not the dominant cause of the anomalous residual stored oil to gas—is likely responsible for the
pressures observed in the LBW. In fact, the gas content variations in fluid properties and other changes in
of fine-grained rocks appears to have the greatest properties of the Cretaceous shales observed in the
impact on compressional wave velocity in these basins LBW. These changes and the continued addition of
(Figure 12). Thus, variations in fluid composition, par- hydrocarbons to the system led to a marked increase
ticularly at gas saturations >15%, are the primary (from 400 to 4000 psi for a gas/water system) in the mea-
cause of the sonic anomalies observed in the LBW, sured displacement pressures of the shales (Table 1).
although lithologic factors also may have played a sig- Preliminary investigation of the coincidence of over-
nificant role. For the remainder of this chapter, sonic pressuring with thermal maturation and hydrocarbon
anomalies are equated to the onset of anomalous pres- accumulation in the LBW has led to a combined analy-
sure (i.e., the pressure boundary) and gas saturation in sis of the relationship of overpressuring to the thermal
the Cretaceous shales. history of the Cretaceous shales in these basins.

Thermal Maturation of the Direct Measurements


Upper Cretaceous Shales in the LBW Production Index
In the central portions of the LBW, significant Vertical trends in production index (PI) from the
changes in the fundamental properties of the Creta- Cretaceous shales in the LBW indicate that more rem-
ceous shales coincide with the sonic transit-time rever- nant hydrocarbons (liquid hydrocarbon + bitumen)
sal, or pressure regime boundary. Overpressuring of are present in the overpressured Cretaceous shale sec-
the Cretaceous shales below the pressure boundary tion than above or below the section (Figure 17; Sur-
can be correlated with (1) an increase in production dam et al., 1994), and that the section is indeed
index [PI = S1/(S1 + S2)], or observed remnant hydro- saturated with hydrocarbons. For instance, above the
carbon, as determined from anhydrous pyrolysis top pressure boundary marking the Cretaceous shale
(Figure 13); (2) variations in kerogen structure and section in the LBW, PI is typically <0.1, and very few
carbon aromaticity; (3) variations in the vitrinite stored hydrocarbons are present. More than 50% of the
Figure 7. Three-dimensional representation of the Muddy Sandstone pseudopotentiometric surface in the vicinity of the Amos Draw and Kitty
fields. These two nearly contiguous overpressured compartments are separated by the Rozet unconformity within the Muddy Sandstone. Heads are
calculated from the repeat formation test data collected by Moncure (1992).
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming
207
208 Surdam et al.

Fuller Reservoir II ( W RB ( wide depth range are shown in Figure 18. The major
0 peak to the right on the spectra is the resonance sig-
nal from carbon in aliphatic functional groups, and
the major peak to the left is carbon in aromatic func-
2000 tional groups. The NMR spectra indicate that with
depth, the peak area for aliphatic carbon decreases.
This decrease reflects changes in kerogen structure
due to thermal maturation: aliphatic functional
4000 groups are lost as a result of the generation of liquid
and gaseous hydrocarbons. At the pressure bound-
ary, the aliphatic carbon peak is either completely
6000 gone or greatly reduced, and kerogen has little
remaining capacity to generate hydrocarbons, either
Depth (ft)

Transition
liquid or gas.
8000 Vitrinite Reflectance
Underpressured
Vitrinite reflectance (Ro), a measure of the thermal
maturation of organic matter with depth, shows pro-
10000 (Severe lost circulation) gressive increase with depth in each basin of the LBW.
Vitrinite reflectance increases slowly with depth, from
Transition 0.4 at 3000 ft (910 m) to 1.0 at 9000 ft (2740 m) (Figure
12000 14). Below 9000 ft (2740 m) depth, Ro increases rapidly,
Overpressured from 1.0 to 1.6 over the 9000 to 11,500 ft (2740–3505 m)
depth interval (Figure 14).
14000 DST 0.64 psi/ft Clay Diagenesis
Transition The diagenesis of interstratified illite/smectite clays
DST 0.48 psi/ft during the progressive burial of a sedimentary
16000 sequence is considered to be an important empirical
170 130 90 50
∆ t (µs/ft) geothermometer (Hower et al., 1976; Pytte and
Reynolds, 1989). It has been demonstrated that the
Figure 8. A typical acoustic transit time vs. depth diagenetic trend in illite/smectite clay is temperature-
plot for the Wind River Basin. Underpressured sec- dependent and may be related to regional hydrocar-
tions are determined by severe lost circulation, mud bon generation (Bruce, 1984; Hagen and Surdam,
weight, and transit time anomalie. Drill-stem test 1984).
data indicate that the formation pressure becomes The alteration of smectite to illite during clay diage-
normal under the overpressured section. nesis was concurrent with the addition of hydrocar-
bons to the fluid-flow system in the Cretaceous shale
section in the LBW. This is evidenced by the fact that
fields in rocks above the boundary follow a normal the percentage of illite in mixed-layer clays increases
(hydrostatic) pressure gradient, are contained within a from 20% to 85% illite, and the structure of the clays
regional, single-phase fluid-flow system, and are changes from random to ordered at ~8000 ±1000 ft
under water drive. In contrast, PI is >0.1 at the top of (2440 ±305 m) (Figure 15), the same depth at which sig-
the overpressured shale section and increases to as nificant increases in the displacement pressure and
much as 0.3 within the section (MacGowan et al., sealing capacity occur in the shales (Figure 15) (Jiao
1994). The difference between PI and the transforma- and Surdam, 1993).
tion ratio (TR) calculated for the reaction of type II
kerogen to liquid hydrocarbon (using kinetic parame- Predictive Kinetic Modeling: Transformation Ratio
ters from Tissot and Welte, 1984) suggests that approx-
imately half of the generated liquid hydrocarbons In order to evaluate the thermal maturation level
have been retained in the source rocks within the over- of the Upper Cretaceous shales in the LBW using
pressured Cretaceous shale section in the LBW (Figure thermal modeling, burial histories were constructed
17; Surdam et al., 1994). The fact that there typically for each basin (Figure 19). Figure 4 shows the time-
are no water legs or hydrocarbon/ water contacts temperature profiles for each basin. Transformation
involved in hydrocarbon production from sandstones ratios (TRs) for the kerogen-to-gas and the liquid-
within the overpressured shale sections in the LBW hydrocarbon-to-gas reactions were calculated with a
(Surdam et al., 1994) seems to support this. program written by Heasler (1995); results are shown
in Figure 20. Conventionally, a TR of 0.4 indicates that
Nuclear Magnetic Resonance liquid hydrocarbons have been generated (Hagen and
13Cnuclear magnetic resonance (NMR) spectra for Surdam, 1984). The kerogen-to-gas reaction has not
organic-rich shale samples from the LBW over a progressed significantly (i.e., reached a TR >0.1) in
Figure 9. East-west cross section of the lower Cretaceous stratigraphic sequence through the Amos Draw, Kitty, and Ryan fields and a wildcat well
in the Powder River Basin. Muddy Sandstone reservoirs along this cross section are oil/water- or gas/water-dominated fluid-flow systems external-
ly sealed above and below by the Mowry and Skull Creek shales. The Muddy Sandstone is internally separated into flow compartments by low-
permeability paleosols and clay-infiltrated sandstones associated with the Rozet unconformity. The Rozet and Skull Creek unconformities are
shown by wavy lines on the cross section. Stippling represents shoreface sandstone.
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming
209
210 Surdam et al.

Figure 10. Southwest-northeast anomalous pressure/gas saturation panel through


the Wind River Basin (Figure 1). The lithology designations are based on gamma-
ray responses. Formation pressure and gas saturation characteristics are based on
acoustic transit time edited for either sandstone or shale DST data. The tops of the
Lance and Morrison formations are shown for structural reference. The normal
pressure regime in the Cretaceous shales is shown in blue. Normally pressured
sandstone reservoirs are represented by white circles. Underpressured sandstone
reservoirs near but below the top of the overpressuring in the Cretaceous shales
are represented by ellipses, are compartmentalized, and probably resulted from
significant post-Laramide erosion. Overpressured sandstone reservoirs are repre-
sented by purple circles and are compartmentalized and gas saturated.

either the centers or the flanks of the LBW (Figure central portions of the LBW was calculated using the
20). The transformation of liquid hydrocarbon to gas kinetic parameters (i.e., activation energy and
for the overpressured Cretaceous section in the frequency factor) of Mackenzie and Quigley (1988).

Table 1. Displacement Pressures and Sealing Capacities Calculated from High-Pressure Mercury Injection
Test Data for the Cretaceous Shale Samples from the Laramide Basins.*

Sample No. Depth (ft) Porosity (%) K (md) Pd (psi) (gas/brine) Sealing Capacity (ft)
Powder River Basin (Mowry Shale)
1 5430 11.7 0.012 400 1000
2 7906 5.2 0.059 1300 3300
3 7909 5.6 0.013 1500 3800
4 8790 4.9 0.012 1800 4600
5 9906 2.7 – 2800 7100
6 10,008 2.2 0.004 3000 7600
7 13,388 2.2 0.004 4000 10,000

Washakie Basin (Lewis Shale)


1 10,037 1.3 0.007 2200 5800
2 10,443 5.2 0.035 2000 5300
3 11,322 2.1 0.008 2200 5800
4 14,392 3.0 0.009 2000 5300

*The porosity and permeability for each sample are also listed. The porosity is helium porosity using Boyle’s Law technique, and permeability
is to gas (sleeve pressure, 400 psi).
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 211

Figure 11. (A) Microphoto-


graph showing a few thin,
separated fractures in the
Mowry Shale at 5000 ft.
Scale bar is 76 µm. (B)
Below 8000 ft, where sonic
logs show a marked
increase in transit time,
abundant wide, linked frac-
tures filled with bitumen
are present in the Mowry
Shale. Scale bar is 38 µm.

In the basin centers of the LBW, liquid hydrocarbons Dominant Mechanism Driving the
in the lower Cretaceous stratigraphic section began Overpressuring in the LBW
converting to gas ~40–50 Ma (Figure 20). The reac-
tion in the basin centers of the LBW reached comple- It is clear from the data presented above that
tion at the time of maximum burial, ~10 Ma, with the there are fundamental changes in organic and inor-
exception of the Washakie Basin, in which all oil ganic geochemical and geophysical properties in the
reacted to gas by 30 Ma. On the flanks of the LBW, rock-fluid system of the Upper Cretaceous shale sec-
according to calculations for the Upper and Middle tion at the pressure boundary. The changes coin-
Cretaceous sections, the reactions of liquid hydrocar- cided with a change in the fluid-flow regime from
bon to gas have reached 25% and 50% completion, water to gas saturation and from normally pres-
respectively. sured to overpressured.
212 Surdam et al.

7.0 0

Powder River
Velocity (v) (ft/sec x 103)

6.0 2000 Bighorn


Wind River
5.0 Washakie
4000
Measured (vp)
4.0
6000

Dept h (ft)
3.0
8000

2.0
0 0.2 0.4 0.6 0.8 1.0 10000
Water Saturation (SW)
Figure 12. Compressional-wave velocity vs.
saturation for a sand pack (from Timur, 1987). 12000

These observations suggest that the generation


and storage of liquid hydrocarbons in the organic- 14000
rich Cretaceous shale section and the subsequent
reaction of this oil to gas were the dominant forces
driving the overpressuring in the LBW. Concur- 16000
rently, clay diagenesis, specifically the alteration of 0.0 0.1 0.2 0.3 0.4 0.5
smectite to illite (dehydration reactions), provided
water to the fluid-flow system, ensuring the multi- PI, fraction
phase fluid character of the shale. As generated but
Figure 13. Plot of the production index (PI) vs. present-
retained liquid hydrocarbons began to react progres-
day depth for the Cretaceous shales in the LBW.
sively to gas, the displacement pressures of the shales
increased by an additional order of magnitude (Fig-
ure 16; Table 1) (Surdam et al., 1994). The fluid-flow
system began to change from a single-phase, water- pressured compartment in each basin of the LBW
dominated regime to a multiphase (oil/gas/water) (Figure 5). This conclusion is illustrated by compar-
hydrocarbon-dominated regime, and low-permeabil- ing Figure 20: the distribution of anomalous pressure
ity lithologies (fluid-flow barriers) began to evolve in the shale-rich Cretaceous section correlates nicely
into capillary seals. with increasing transformation ratios for the oil-to-
Regardless of how low the absolute permeability gas reaction.
is, the low-permeability paleosol or shale horizons In the LBW, the primary difference between pres-
that form flow barriers in a single-phase fluid-flow sure compartmentalization in Cretaceous sandstones
system have finite leak rates, and anomalous pres- and shales is one of scale. The overpressured compart-
sures cannot be maintained (Jiao and Surdam, 1994). ments within the Cretaceous sandstones are typically
However, in a multiphase fluid-flow system relatively small (longest dimension 1–10 mi) (Figure
(gas/oil/water) dominated by capillarity, fluid 6), whereas the surrounding shale is overpressured on
crosses the capillary seal only when the displace- a basinwide scale (Figures 5A, 6).
ment pressure, or lithostatic gradient (fracture gradi- As in the shales, the transition of the fluid-flow sys-
ent), is exceeded. Thus, the relative permeability of tem from single-phase to multiphase in the sandstones
rocks in a multiphase fluid-flow system has much was driven by the storage of liquid hydrocarbons and
more impact on the maintenance of anomalous pres- their subsequent reaction to gas. In fact, the fluid-flow
sures than their absolute permeability (Berg, 1975); systems of the Cretaceous sandstones and shales in the
in the LBW, the overpressured Cretaceous shales are LBW were closely related during the initial stages of
orders of magnitude more effective in retaining overpressuring in the shales (in which primary migra-
anomalous fluid pressure than a conventional strati- tion of hydrocarbons into microfractures occurred).
graphic trap. As a result, the Cretaceous shale-rich This close relationship persisted until pressure buildup
system became a large, seemingly basinwide over- in the shales resulted in hydrocarbon expulsion and
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 213

0 0

Powder River
2000 Bighorn 2000
Wind River
Washakie
4000 4000

6000 6000
Depth (ft)

Depth (ft)
8000 8000

10000 10000

12000 12000
Powder River
Bighorn
14000 14000
Wind River
Washakie
16000 16000
0.0 0.4 0.8 1.2 1.6 2.0 0 20 40 60 80 100
R o (%) Illite in the I/S (%)
Figure 14. Plot of vitrinite reflectance Ro vs. Figure 15. Plot of percent illite in the mixed-layer
present-day depth for the Cretaceous shales in smectite-illite clays vs. present-day depth for the
the LBW. Cretaceous shales in the LBW.

migration into sandstone conduits and traps, which exposure, liquid hydrocarbons in the trap began react-
may have occurred once or many times, depending on ing to gas and bitumen. At this time, if the trap was
the nature of the source rocks and how often the dis- completely surrounded by capillary seals (3-D clo-
placement pressure, or fracture gradient, in the shales sure), a pressure compartment would have formed
was exceeded by internal pressure buildup. that was completely isolated from the shales above
As hydrocarbons migrated from the shales into and below, and from any sandstones updip and
stratigraphic/diagenetic traps in the sandstones, free downdip. The capillary seals observed in these sand-
water was expelled from the bottom of the traps, stones are typically capable of withstanding a 1000-psi
where the low-permeability horizons were still in con- pressure differential at the top and a 4000-psi pressure
tact with a dominantly single-phase flow system and differential at the bottom of the overpressured Creta-
had finite leak rates. As long as free water was a signif- ceous shale section (Figure 7).
icant component of the fluid phase, the traps remained
under water drive, even though the seals in contact Sealing Capacity of the Cretaceous Shales
with liquid hydrocarbon accumulations were domi-
nated by capillarity. When hydrocarbons finally satu- Sealing capacity is a function of pore radius and
rated a trap and the fluid-flow regime changed from fluid characteristics, and is expressed as the height of
single-phase to multiphase (Berg, 1975; Iverson et al., the hydrocarbon column that can be supported by a
1994), low-permeability zones in the sandstones (e.g., seal. According to the classification of Sneider et al.
lowstand unconformities, transgressive shales, and (1991), the typical increase in sealing capacity
paleosols) converted to capillary seals (Figure 16; Jiao observed in the LBW across the pressure regime
and Surdam, 1993; Martinsen, 1994; Surdam et al., boundary is equivalent to a change from a class C
1994). With additional burial and increased thermal seal (capable of supporting an oil column of >100 but
214 Surdam et al.

4000 and 6150 ft (1870 m) at 13,000 ft (3960 m) depth. The


paleosol horizon has a similar trend.
Gas Column
Timing of Overpressuring
It is possible to place an approximate time on the
6000 initiation and subsequent development of overpres-
suring in the basins by (1) identifying the depths at
which the generation of liquid hydrocarbons began
and their subsequent cracking to gas became impor-
tant in the LBW (when TR >0.1) (Figure 19) and (2)
8000
Depth (ft)

locating the corresponding depths in burial history


diagrams for these basins (Figure 20). In the Creta-
ceous shale sections in the central portions of the
basins, the potential for overpressuring began at ~70
Ma (TR of 0.1–0.4 for the kerogen-to-oil reaction) and
10000 probably was firmly developed at approximately 40 Ma
(TR >0.1 for the oil-to-gas reaction; Barker, 1990). Signif-
icant progress in the reaction of liquid hydrocarbon-to-
gas during the last 40 m.y. suggests that substantial
overpressuring has been present in the Cretaceous
12000 shale section during this period of time, leading to the
conclusion that in the centers of the LBW, the reaction
of liquid hydrocarbon to gas had progressed signifi-
cantly by the time of maximum burial at 10 Ma (Fig-
ures 19, 20). Thus, compartmentalized overpressuring
14000 in these basins is not an ephemeral geological phe-
0 4000 8000 12000
nomenon, but is rather a long-lived geological process
Gas Column (ft) intimately related to hydrocarbon generation and
reaction within a rock sequence dominated by
Figure 16. Plot of the displacement pressure vs. organic-rich shales.
present-day depth for the Cretaceous shales in
the LBW.
EVOLUTION OF VARIOUS PRESSURE
REGIMES IN THE CRETACEOUS
<500 ft) to a class A seal [capable of supporting an oil SANDSTONES IN THE LBW
column of >1000 ft (305 m)] (Table 1).
Smith’s equation was used in this study to calculate As indicated, the Cretaceous sandstones contained
the sealing capacity of potential hydrocarbon seals: in the overpressured Cretaceous shale section in the
LBW can be normally pressured, underpressured, or

H=
(PdB − PdR ) overpressured. This section describes possible evolu-
0.433(ρ w − ρ h )
(1) tion pathways for the Cretaceous sandstones.
At any specific depth, the “shale displacement
pressure” shown in Figure 4 is significantly greater
where H is the maximum hydrocarbon column in feet than the pressure along the “sandstone gradient”;
that a seal can support, PdB is the subsurface displace- both gradients are parallel to the regional lithostatic
ment pressure in pounds per square inch (psi) of the gradient, but offset to lower pressures. Note that even
seal rock, PdR is the subsurface displacement in psi of though the reservoir sandstones follow a gradient
the reservoir rock, ρ w is the subsurface density in parallel to the regional lithostatic gradient, there is no
g/cm3 of water, ρh is the subsurface density of the evidence of natural hydraulic fracturing, or poly-
hydrocarbon, and 0.433 is a units conversion factor. directional fracturing, in these sandstones, even those
The displacement pressure, the minimum pressure subjected to the highest overpressuring. One explana-
required to start mercury moving through a sample, tion for this observation has been offered by Lorenz et
was determined by extrapolating the injection pres- al. (1991), who note that natural hydraulic fracturing
sure curve (plateau portion) to 0% mercury saturation. of overpressured rocks is extremely rare because “as
This pressure was then converted to displacement the pore pressure increases, the horizontal in-situ
pressure for the subsurface gas/water and oil/water stresses increase proportionately until all stresses are
system using Schowalter’s (1979) nomograms. Results nearly equal, at which point the development of an
are shown in Table 1. For a gas/water system, the seal- oriented, large-scale systematic fracture system is
ing capacity of the Cretaceous shales is 600 ft (180 m) unlikely.” In some reservoir sandstones within the
(gas column) at a depth of 5000 ft (1520 m) and rapidly Cretaceous shale compartment, small-scale unidirec-
increases to 2150 ft (655 m) at 8000 ft (2440 m) depth tional and nearly vertical extensional fractures are
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 215

TR/PI Figure 17. Plot of


production index (PI) and
0.0 0.2 0.4 0.6 0.8 1.0 transformation ratio (TR)
0 vs. depth for the Mowry
Transformation Ratio Shale, Powder River Basin.
Sample depths for the PI
Production Index have been corrected for
2000 0
post-Laramide erosion.
The difference between
PI and TR <9000 ft suggests
4000 2000 that a significant portion
of the generated and
expelled hydrocarbon
has remained in the
Maximum Burial Depth (ft)

6000 4000 organic-rich shale.

Present-Day Depth (ft)


8000 6000
Measured Remaining
Liquid Hydrocarbon
10000 8000
Calculated Liquid
Hydrocarbon Generation
12000 10000

14000 Hydrocarbon
Stored in
12000
Source Rocks

16000 14000

observed; the small-scale fractures typically termi- compartment at constant volume. However, at some
nate within the sandstones. depth, the liquid-hydrocarbon-to-gas reaction will
Laboratory measurements indicate that if the capil- reach completion, and will no longer provide a mecha-
lary displacement pressure is greater than the pressure nism for increasing pressure. Thereafter, the possibil-
differential needed for microfracture, the measured ity exists that the pressure/depth gradient of a
shale displacement values (Figure 4) may represent sandstone compartment might decrease with progres-
the fracture gradient of the shales, or at least be paral- sive burial.
lel to this gradient. If this is the case, the measured However, as all basins of the LBW have experi-
capillary displacement pressures are minimal values, enced at least some uplift and erosion (unroofing),
much less than the actual capillary displacement pres- deviation from the sandstone gradient due to great
sures of the shales. Therefore, as pressure increases, burial is not presently observed in the basin areas
the fracture gradient of these will always be inter- studied. The LBW were uplifted during and subse-
sected before the capillary displacement pressure is quent to the Laramide Orogeny; the exact amount of
exceeded. uplift has not been established, and the amount of
It is possible that with progressive burial to depths unroofing is highly variable. For example, in the Pow-
greater than those shown in Figure 4, the hydrocarbon- der River Basin, the estimated erosion is 1000 ft in the
saturated sandstone compartments will deviate from basin center and 4000 ft or more along the east margin
the sandstone gradient, proceeding along some differ- of the basin (Jiao, 1992). Generally, the amount of ero-
ent pressure/depth gradient. With progressive burial, sion and uplift in the LBW varies from ~1000–2000 ft
these isolated hydrocarbon accumulations initially (305–610 m) in the basin centers to more than 5000 ft
would be characterized by increasing pressure due to (1520 m) along the basin margins (Jiao, 1992;
at least two causes: (1) at constant volume, the liquid- Kharitonova, 1995). For these basins, determining
hydrocarbon-to-gas reaction continually increases the what happened to the isolated gas-saturated compart-
pressure of these isolated sandstone compartments ments during unroofing is of great importance.
and (2) as the depth of burial increases, the tempera- Certain compartments, particularly those that were
ture increases, leading to elevated pressure in the in the lower portion of the overpressured shale section
216 Surdam et al.

Figure 18. 13C nuclear


Powder Washakie magnetic resonance
spectra of the Mowry
River Basin Shale samples from
Basin depths of 3000–18,000 ft
in the LBW. Note the
4900 ft diminishing or
3085 ft disappearance of the
7441 ft
aliphatic carbon peak
4010 ft (the right-hand major
9159 ft
peak) <10,000 ft.
5720 ft
10663 ft
8377 ft
13659 ft
9230 ft
18420 ft

9695 ft
300 200 100 0 -100
11425 ft ppm

Bighorn
300 200 100 0 -100
Basin
ppm
Wind River 3015 ft
Basin 4100 ft
4530 ft
10350 ft 10500 ft
12440 ft 11010 ft

14280 ft 14060 ft

300 200 100 0 -100 300 200 100 0 -100


ppm ppm

during unroofing, appear to proceed up along the intersect the upper boundary of the overpressured
sandstone gradient of Figure 4 (point B’, Figure 21B). shale section during unroofing. However, it is specu-
Those compartments that were in the upper portion lated that the compartment seals are breached at the
of the overpressured shale section during unroofing upper boundary, where the pressure contrast between
and that intersected the boundary between normally the compartment and its external environment is
pressured and overpressured rock have a different greatest, resulting in gas being released into the over-
pressure/depth pathway (point A’, Figure 21B). Typi- lying normally pressured rock system. Subsequently,
cally at this boundary, pressure decreases in the gas- the fluid in the compartment equilibrates with the
saturated compartments; with further unroofing, these overlying gas-charged column, and the fluid gradient
rocks begin to follow an approximate gas gradient becomes a function of the weight of the gas-charged
(0.2–0.3 psi/ft) upward until the hydrostatic gradient is fluid column. There is considerable literature to sup-
intersected (Figure 21B). Because they follow a gas port the hypothesis that fluid expulsion from over-
gradient, these compartments are conventionally clas- pressured compartments is episodic and short-lived;
sified as underpressured. It should be noted that these any fractures opened during such expulsion are
gas-saturated compartments all plot below the shale rapidly sealed by silica or carbonate cements resulting
displacement pressure gradient in Figure 4. from the drop in pressure and temperature accompa-
It is uncertain what mechanism or process causes nying the expulsion (Bruton and Helgeson, 1983; Low-
the shift in pressure regimes as these compartments ell et al., 1993; Roberts and Nunn, 1995). Thus, if these
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 217

Figure 19. Burial history diagrams for the central Powder River, Bighorn, Wind River, and Washakie basins.

compartments are breached by fracture, it appears that of the anomalous pressure section during the uplift
they are quickly reestablished, albeit at significantly and erosion of the LBW.
lower pressure. With further unroofing, no internal
mechanism exists to increase pressure, so the compart-
ments remain isolated but underpressured. These iso-
A NEW CLASS OF HYDROCARBON
lated, underpressured, gas-saturated compartments ACCUMULATION
are observed to exist at the margins of the Laramide
basins of Wyoming (LBW) in the upper portion of and The anomalously pressured, gas-saturated com-
above the overpressured shale section; they remain partments associated with Cretaceous sandstones in
underpressured until they intersect the hydrostatic the deeper portions of the Laramide basins of
gradient (Figure 4). Wyoming (LBW) are dependent neither on structural
Note that some of the largest anomalously pres- closure, faults, nor stratigraphic pinchout. Structural
sured gas accumulations in the LBW are underpres- and stratigraphic factors can be important to these
sured. For example, the Wattenberg field in the accumulations, but are independent of the time of
Denver Basin, and the fields known collectively as hydrocarbon generation. In contrast, the compart-
the Elmworth area and the Hoadley field of the mentalized, anomalously pressured gas accumula-
Western Canada Basin—all giant gas fields—are tions are dependent on the basinwide stratigraphic
underpressured. Moreover, the most significant por- fabric (e.g., a relatively thick shale and/or coal rich
tion of all anomalously pressured gas production in section); the conversion of the fluid-flow system from
the LBW is derived from the stratigraphic interval single-phase to multiphase (e.g., conversion of low-
between the pressure boundary down to 2000 ft permeability flow barriers to capillary seals) due to
below the boundary (Figure 22). In other words, it hydrocarbon saturation; diagenesis (e.g., reducing
appears that gas was concentrated in the upper part permeability in the fine-grained and low-permeability
218 Surdam et al.

Figure 20. Computed transformation ratios of liquid hydrocarbon to gas for the central (A) Powder River, (B)
Bighorn, (C) Wind River, and (D) Washakie basins.

lithologies, and in some cases enhancing porosity in the key to unlocking the immense gas resources of
the reservoir facies); and the nature of hydrocarbon the LBW.
generation, storage, reaction, and expulsion patterns
(Figure 23). Because these anomalously pressured and
gas-saturated compartments depend on a new CONCLUSIONS
assemblage of determinative processes, we feel justi-
fied in recommending that they be treated as a new Conventional hydrologic theory suggests that over-
and unconventional class of hydrocarbon accumula- pressuring is a function of disequilibrium compaction
tion (Figure 24). driven by the imbalance between the rates of sedimen-
Regardless of classification, compartmentalized tation and fluid flux (escape) or artesian flow.
gas accumulations will play a major role in future Although disequilibrium compaction, or water expul-
exploration for the huge energy resources contained sion, may have originally played a role in the develop-
in the deeper portions [>8000–9000 ft (2440–2740 m)] ment of overpressuring of the Laramide basins of
of the LBW. Innovative technologies must be devel- Wyoming (LBW), it clearly has played little or no role
oped to expedite this exploration, as exploration in maintaining these anomalous pressure regimes for
techniques designed to locate conventional strati- millions of years.
graphic and structural traps tend to be blind to these The pressure compartments documented and char-
unconventional, subtle traps. The key to expediting acterized in these basins are the result of multiphase
this exploration is to develop techniques to detect fluid flow coupled with three-dimensional closure of
storage capacity within the anomalously pressured capillary seals. Thus, pressure compartmentalization
and gas-saturated shale section. Commercial gas in the LBW is genetically related to hydrocarbon gen-
storage capacity could range in type from intergran- eration, storage, migration, and reaction (kerogen-to-
ular porosity in sandstone to fracture porosity in liquid-hydrocarbon and liquid-hydrocarbon-to-gas +
shales. The first priority in finding new and large “bitumen”). The addition of hydrocarbons to the fluid-
commercial gas accumulations in these basins will be flow system was essential to pressure compartment
to determine storage capacity and deliverability— evolution in these basins. First, the appearance of
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 219

(a) Pressure Evolution During Burial (b) Pressure Evolution During Uplift
Pressure (psi) Pressure (psi)
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
0 0
"L
ith
os
ta
2000 tic

0.4
2000
"Hy

"
"Li G

33
d ro s

tho ra
di

ps
sta
tatic

tic en

i/ft
"G t,
" Gr

4000 rad 4000 1.


ien 0
adie

t, 1
ps
i/f
nt, 0

.0 t
ps
i/ft

Depth (ft)
.433

6000 6000
A'
p si /
Depth (ft)

ft

"Gas"
Top of Pressure Compartment Gradient Top of Pressure
8000 8000
Sh Compartment
ale
"D Sh
A isp Pr ale
lac es "
10000 em 10000 B' an
S su Dis
en re pl
tP ds " ac
B Sa res
su
to
ne
G e
ra m
nd re" di en
sto Gr 12000 Pr en t
12000 ne ad es
"F ien su t
rac t re
tur G
e" ra
Gr di
14000 ad 14000 en
ien
t t

Figure 21. Present evolution pathways during burial (a) and subsequent uplift and unroofing (b).

hydrocarbons in the system gradually drove the tran- pressure compartments characterizing the Creta-
sition from single-phase (water) to multiphase fluid ceous sandstones of the LBW cannot be under water
flow (gas/oil/water), thereby activating capillary seals. drive—they must be under gas-depletion drive; (2)
Second, as hydrocarbons completely saturate the com- hydrocarbon generation, storage, expulsion, and
partment, they ensure the integrity of the three-dimen- migration are essential to the formation of the type
sional aspect of the bounding capillary seals. In of pressure compartments found in these basins;
addition, hydrocarbon saturation of the compartment and, most importantly, (3) although the pressure/fluid
provides a mechanism for expelling free water from compartmentalization found in these sandstones
the compartment system. begins with an infusion of hydrocarbons from
Clearly, hydrologic overpressuring and compart- associated shales and evolves as the result of the
mentalization of anomalous pressure regimes are fluid-flow system undergoing a transition from
fundamentally different processes and have very dif- single-phase to multiphase flow, a sandstone com-
ferent implications for the exploitation of gas partment, once formed, is isolated from the rest of
resources. Multiphase pressure compartmentaliza- the rock system. A significant implication, particu-
tion is long-lived; it develops midway to late in a larly with respect to drilling procedures, is that the
burial history; it is intimately related to hydrocarbon sandstone pressure compartments need not be at the
generation and reaction; and it can be found over a same pressure as the surrounding overpressured
wide depth interval. Most importantly, gas resources shales.
associated with multiphase fluid-flow compartments In summary, the concept and understanding of
can occur in rocks of any age; they will not spill dur- multiphase fluid flow as it relates to three-dimen-
ing structural reorientation (excluding rupture by sional pressure compartmentalization and fracture
fracturing); and their integrity or appearance can gradients will greatly expedite the search for, the
only be destroyed by significant uplift when the discovery of, and the exploitation of new, unconven-
pressure differential exceeds the displacement pres- tional gas resources. One of the most exciting aspects
sure of the capillary seals, or the fracture gradient is of this work is that the type of sonic log contrast rep-
exceeded, or by deep burial when the pressure dif- resented by the pressure compartments shown in
ferential is eliminated, even though the compart- Figures 7 and 9 should be detectable with modern
ments remain intact. seismic techniques and new pattern-recognition
Three points should be emphasized: (1) the hydro- technology. Because these pressure anomalies are
carbon accumulations associated with the type of the direct result of gas accumulation, their detection
220 Surdam et al.

Distance to Top of Anomalous Pressure (ft)

Distance to Top of Anomalous Pressure (ft)


Cretaceous Major Gas Reservoir, Bighorn Basin Cretaceous Major Gas Reservoirs, Powder River Basin
-5000 -5000
-4000 -4000
-3000 -3000
-2000 -2000
-1000 -1000
0 0
1000 1000
2000 2000
3000 3000
Reservoir: 22 Reservoir: 58
4000 CGP: 776 bcf 4000
CGP: 1456 bcf
5000 5000
0 100 200 300 400 500 0 200 400 600 800
Cumulative Gas Production (bcf) Cumulative Gas Production (bcf)
Distance to Top of Anomalous Pressure (ft)

Distance to Top of Anomalous Pressure (ft)


Cretaceous Major Gas Reservoir, Washakie Basin Cretaceous Major Gas Reservoir, Wind River Basin
-5000 -5000
-4000 -4000
-3000 -3000
-2000 -2000
-1000 -1000
0 0
1000 1000
2000 2000
3000 3000
Reservoir: 26 Reservoir: 32
4000 CGP: 1642 bcf 4000
CGP: 863 bcf
5000 5000
0 200 400 600 800 0 200 400 600 800
Cumulative Gas Production (bcf) Cumulative Gas Production (bcf)

Figure 22. Cumulative gas production diagrams for the (A) Powder River, (B) Wind River, (C) Bighorn, and (D)
Washakie basins, showing production in the stratigraphic interval between the top pressure boundary down
to 2000 ft below the boundary. Gas was probably concentrated in the upper part of the anomalous pressure
section during uplift and erosion of the LBW. bcf = billion cubic feet; CGP= cumulative gas production.

Figure 23. Comparison of a


Conventional trap Anomalously Pressured conventional hydrocarbon
trap (e.g., structural closure)
(Water Drive) Gas Accumulation under water drive with an
(Gas-Depletion Drive) unconventional, anomalous-
ly pressured gas accumula-
tion under gas-depletion
Trap drive. Delineating storage
Gas capacity within the regional
Oil gas-saturated stratigraphic
section is vital to the explo-
Anomalous Gas ration for unconventional
Pressure Saturated hydrocarbon accumulations
Regime Volume in the LBW.
Water Flux

Water Flux
Anomalously Pressured Gas Compartments in Cretaceous Rocks of the Laramide Basins of Wyoming 221

Figure 24. Schematic


UNCONVENTIONAL CONVENTIONAL
diagram showing
unconventional gas
accumulations in
anomalously pressured
compartments. The
sandstones contained
in the Cretaceous shale
section are anomalously
al
pressured, compartmental-
rm ure ized, and gas-saturated.
no ss
"Pressure Seal" pr
e These accumulations
are not formed as a result
of structural or stratigraphic
closure, but rely instead
s on capillary seals to
ou form traps.
al re
om su
an res
p "Sweetspots"

should result in the discovery of new hydrocarbon v. 283-A, p. 540–588.


resources. Hagen, E.S., and R.C. Surdam, 1984, Maturation his-
tory and thermal evolution of Cretaceous source
rocks of the Big Horn Basin, Wyoming and Mon-
ACKNOWLEDGMENTS tana, in J. Woodward, F. Meisser, and J. Clayton,
This study was funded by the Gas Research Insti- eds., Hydrocarbon source rocks of the Greater
tute under contract 5089-260-1894 and the NSF- Rocky Mountain Region: Rocky Mountain Associa-
EPSCoR-ADP Fossil Energy program under contract tion of Geologists, p. 321–338.
EHR-910-8774. The original manuscript was reviewed Heasler, H.P., R.C. Surdam, and J.H. George, 1994,
by Kathy Kirkaldie. In addition, F. Meisner, J. Lorenz, Pressure compartments in the Powder River Basin,
and W. Tyrrell reviewed and improved an earlier draft Wyoming and Montana, as determined from drill-
of the manuscript. R.C. Surdam acknowledges useful stem test data, in P. Ortoleva, ed., Basin compart-
discussions with Bob Siebert of Conoco and Alex Mar- ments and seals: AAPG Memoir 61, p. 235–262.
shall and John Bates of VICO. We also acknowledge Hottman, C.E., and R.K. Johnson, 1965, Estimation of
the significant contributions made to this paper by formation pressures from log-derived shale proper-
Nick Boyd, William Iverson, Donald MacGowan, ties: Journal of Petroleum Technology, v. 17,
Randi Martinsen, Debi Maucione, Francis Miknis, p. 717–772.
Rebecca Moncure, Vladimir Serebryakov, Scott Smith- Hower, J., E.F. Eslinger, M.E. Hower, and E.A. Perry,
son, and Yue Wang. 1976, Mechanism of burial and metamorphism of
argillaceous sediments: 1. Mineralogical and chemi-
cal evidence: Geological Society of America Bul-
REFERENCES CITED letin, v. 87, p. 725–737.
Hunt, J.M., 1979, Petroleum Geochemistry and Geol-
Barker, C., 1990, Calculated volume and pressure ogy: San Francisco, Freeman and Co., 617 p.
change during the thermal cracking of oil to gas in Hunt, J.M., 1990, Generation and migration of
reservoirs: AAPG Bulletin, v. 74, p. 1254–1261. petroleum from abnormally pressured fluid com-
Berg, R.R., 1975, Capillary pressure in stratigraphic partments: AAPG Bulletin, v. 74, p. 1–12.
traps: AAPG Bulletin, v. 59, p. 939–956. Iverson, W.P., R.S. Martinsen, and R.C. Surdam, 1994,
Bruce, C.H., 1984, Smectite dehydration and its rela- Pressure seal permeability and two-phase flow, in
tion to structural development and hydrocarbon P. Ortoleva, ed., Basin compartments and seals:
accumulation in Northern Gulf of Mexico Basin: AAPG Memoir 61, p. 313–319.
AAPG Bulletin, v. 68, p. 673–683. Jiao, Z.S., 1992, Thermal maturation/diagenetic
Bruton, D.J., and H.C. Helgeson, 1983, Calculation of aspects of the abnormal pressure in cretaceous
the chemical and thermodynamic consequences of ahales and sandstones, Powder River Basin,
differences between fluid and geostatic pressure in Wyoming: Ph.D. dissertation, University of Wyom-
hydrothermal systems: American Journal of Science, ing, Laramie, 242 p.
222 Surdam et al.

Jiao, Z.S., and R.C. Surdam, 1993, Low-permeability Moncure, R.S., 1992, Diagenesis and hydrodynamics
rocks, capillary seals, and pressure compartment of the lower Cretaceous Muddy Sandstone in the
boundaries in the Cretaceous section of the Powder vicinity of Amos Draw, Elk Draw and Kitty fields,
River Basin: Wyoming Geological Association Powder River Basin: Ph.D. dissertation, University
Jubilee Anniversary Field Conference Guidebook, of Wyoming, Laramie, 242 p.
p. 297–310. Odland, S.K., P.E. Patterson, and E.D. Gustason,
Kharitonova, N.A., 1995, Analysis of the sonic well 1988, Amos Draw field: A diagenetic trap related
logs applied to erosion calculation in the Bighorn to an intraformation unconformity in the Muddy
Basin, Wyoming: M.S. thesis, University of Sandstone, Powder River Basin, Wyoming, in
Wyoming, Laramie, 105 p. R. Diedrich, M. Dyka, and W. Miller, eds.: Wyom-
Law, B.E., C.W. Spencer, R.A. Carpenter, R.F. Crovelli, ing Geological Association 39th Annual Field Con-
R.F. Mast, G.L. Dolton, and C.J. Wandrey, 1989, ference Guidebook, p. 147–160.
Estimates of gas resources in overpressured low- Powley, D.E., 1982, Pressures, normal and abnormal:
permeability Cretaceous and Tertiary sandstone AAPG Advanced Exploration Schools Unpublished
reservoirs, Greater Green River Basin, Wyoming, Lecture Notes, 38 p.
Colorado, and Utah, in J. Eisert, ed., Gas Resources Pytte, A.M., and R.C. Reynolds, Jr., 1989, The thermal
of Wyoming: Wyoming Geological Association 40th transformation of smectite to illite, in N. Naeser and
Field Conference Guidebook, p. 39–61. T. McCullock, eds., Thermal history of sedimentary
Lorenz, J.C., L.W. Teufel, and N.R. Warpinski, 1991, basins: New York, Springer-Verlag, p. 133–140.
Regional fractures I: a mechanism for the formation Roberts, S.J., and J.A. Nunn, 1995, Episodic fluid
of regional fractures at depth in flat-lying reser- expulsion from geopressured sediments: Marine
voirs: AAPG Bulletin, v. 75, p. 1714–1737. and Petroleum Geology, v. 12, p. 195–204.
Lowell, R.P., P.V. Cappellen, and L.N. Germanovich, Schowalter, T.T., 1979, Mechanics of secondary hydro-
1993, Silica precipitation in fractures and the evolu- carbon migration and entrapment: AAPG Bulletin,
tion of permeability in hydrothermal upflow zones: v. 63, p. 723–776.
Science, v. 260, p. 192–194. Sneider, R.M., K. Stolper, and J.S. Sneider, 1991, Petro-
MacGowan, D.B., Z.S. Jiao, R.C. Surdam, and F.P. physical properties of seals: AAPG Bulletin, v. 75,
Miknis, 1994, Formation water chemistry of the p. 673–674.
Muddy Sandstone and organic geochemistry of the Surdam, R.C., Z.S. Jiao, and R.S. Martinsen, 1994, The
Mowry Shale, Powder River Basin, Wyoming and regional pressure regime in Cretaceous sandstones
Montana: evidence for mechanism of pressure com- and shales in the Powder River Basin, in P. Ortol-
partment formation, in P. Ortoleva, ed., Basin com- eva, ed., Basin compartments and seals: AAPG
partments and seals: AAPG Memoir 61, p. 321–331. Memoir 61, p. 213–233.
MacKenzie, A.S., and T.M. Quigley, 1988, Principles of Tissot, B.P., and D.H. Welte, 1984, Petroleum Forma-
geochemical prospect appraisal: AAPG Bulletin, tion and Occurrence; 2nd ed.: Berlin, Springer-
v. 72, p. 399–415. Verlag, 699 p.
Magara, K., 1976, Thickness of removed sedimentary Timur, A., 1987, Acoustic logging, in H. Bradley, ed.,
rocks, paleopressure, and paleotemperature, south- Petroleum Engineering Handbook: Dallas Society
western part of Western Canada Basin: AAPG Bul- of Petroleum Engineers, p. 51-1–51-52.
letin, v. 60, p. 554–565. Wheeler, D.M., E.R. Gustason, and M.J. Furst, 1988,
Martinsen, R.S., 1994, Stratigraphic compartmentation The distribution of reservoir sandstone in the
of reservoir sandstones: examples from the Muddy Lower Cretaceous Muddy Sandstone, Hilight field,
Sandstone, Powder River Basin, Wyoming, in P. Powder River Basin, Wyoming, in A. Lomando and
Ortoleva, ed., Basin compartments and seals: AAPG P. Harris, eds., Giant oil and gas fields: Society of
Memoir 61, p. 273–296. Economic Paleontologists and Mineralogists Core
Martinsen, R.S., 1995, Stratigraphic compartmentaliza- Workshop 12, p. 179–228.
tion of reservoir sandstones: Examples from the Wyoming Geological Association, 1993, Wyoming
Muddy Sandstone, Powder River Basin, Wyoming, stratigraphic nomenclature chart, in B. Stroock and
in P. Ortoleva, ed., Basin compartments and seals: S. Andrew, eds.: Casper, Wyoming, Wyoming Geo-
AAPG Memoir 61, p. 273–296. logical Association.
Martinsen, R.S., 1997, Stratigraphic controls on the devel-
opment and distribution of fluid-pressure compart-
ments, in R.C. Surdam, ed., Seals, traps, and the
petroleum system: AAPG Memoir 67, p. 223–241.
Chapter 13

Stratigraphic Controls on the Development


and Distribution of Fluid-Pressure
Compartments
R.S. Martinsen
Department of Geology and Geophysics, University of Wyoming
Laramie, Wyoming, U.S.A.

ABSTRACT
Studies in the Powder River Basin of Wyoming indicate that the boundaries
of individual fluid pressure compartments commonly correspond to the
boundaries of various stratigraphic elements (e.g., lithofacies and unconformi-
ties). This finding leads to the question: “What, then, is the difference between
a stratigraphic trap and a fluid pressure compartment?” The fundamental dif-
ference is that conventional stratigraphic and structural traps have incomplete
capillary seal closure, whereas fluid pressure compartments (of the type
observed in the Powder River Basin) have complete capillary seal closure.
Most structural traps and many stratigraphic traps are not completely bound-
ed by low-permeability rocks and have discrete spillpoints. They cannot
achieve complete capillary seal closure and are incapable of becoming pres-
sure compartments. Even the many stratigraphic traps (and fewer structural
traps) that are enclosed by low-permeability rocks do not all comprise pres-
sure compartments. In order for a hydrocarbon reservoir to be enclosed by
capillary seals, not only must it be enclosed by low-permeability rocks, it must
be either completely filled (lack a hydrocarbon-free water contact) or enclosed
by low-permeability rocks that contain multiple fluid phases. These conditions
are more likely met in reservoirs completely enclosed within, or closely associ-
ated with, mature source rocks. The distribution and characteristics of fluid
pressure compartments observed in the Powder River Basin of Wyoming are
discussed within the framework of these types of stratigraphic variables.

INTRODUCTION the generation and maturation of hydrocarbons, espe-


cially the cracking of oil to gas. Overpressuring occurs
Studies in the Powder River Basin of Wyoming have in normally compacted rocks and has been present in
led to the identification of numerous reservoir-scale those rocks for at least 20 m.y. (Jiao, 1992). Boundaries
fluid-pressure compartments (Heasler et al., 1995). of these fluid-pressure compartments typically coincide
Overpressuring in these compartments is attributed to with various types of stratigraphic discontinuities or

223
224 Martinsen

facies changes (Martinsen, 1995). In fact, all the fluid- the trap cannot have a spillpoint. The majority of traps
pressure compartments identified by Heasler et al. that meet this condition are stratigraphic traps or com-
(1995) are also classified as stratigraphic traps. How- bined stratigraphic-structural-diagenetic traps (Figure
ever, not all stratigraphic traps in the basin meet the 1). Not all stratigraphic traps, however, lack spill-
criteria for being fluid-pressure compartments (e.g., points. Therefore, not all stratigraphic traps are capa-
they are not externally closed to fluid movement and ble of becoming fluid-pressure compartments.
they appear to be in hydrodynamic continuity with Furthermore, not all traps that are enclosed by low-
the surface) (Heasler et al., 1995). permeability rocks have total capillary seal closure.
Containment of abnormally high pressure fluid com- If the low-permeability rocks enclosing the trap are
partments over geologic time is a significant problem. In filled with a single fluid phase (water), then the reser-
fact, many geologists believe that overpressuring is time voir must be completely filled and lack a free-water
transient, no matter what its origin, and that pore pres- leg. A contact between water-filled reservoir rock and
sures always attain equilibrium within a few million water-filled low-permeability rock is not a seal, it is an
years. This appears to be true in many areas, such as por- aquitard. Therefore, anomalous pressures generated
tions of the U.S. Gulf Coast, where overpressuring in an incompletely filled trap can be equilibrated by
occurs in both aquifers and hydrocarbon reservoirs; is movement of water out from the water-leg portion of
often associated with thick, undercompacted shale sec- the reservoir into and through the water-saturated
tions; and is commonly the result of disequilibrium com- pores of the low-permeability rocks adjacent to or
paction (Bethke et al., 1988). However, in other areas, below the reservoir. An example of this situation
such as the Powder River Basin, where overpressuring is would involve an incompletely filled reservoir that
dominantly due to hydrocarbon generation and matura- becomes temporarily overpressured because of an
tion, overpressuring may be maintained over much increase in hydrocarbon fluids associated with the
longer time spans because multiphase fluid conditions cracking of oil to gas. In such a reservoir, overpressur-
decrease effective permeabilities. ing could be normalized by the downward displace-
Multiphase fluid conditions exist when two or more ment of the hydrocarbon/water contact within the
gaseous or liquid components occupy the void spaces reservoir and consequent movement of water out of
within a given volume of rock. The addition of hydro- the reservoir into surrounding low-permeability
carbons, either by generation from organic matter or rocks. Whether or not a trap becomes completely filled
migration, into the pores of a water-saturated rock is is a function of (1) the amount of hydrocarbons avail-
the dominant mechanism by which multiphase fluid able to fill the reservoir relative to the volume of reser-
conditions are created in the subsurface. Under condi- voir and (2) the holding capacity of the top and lateral
tions of multiphase fluid flow, low-permeability rocks seals relative to the amount of closure in the trap.
can become seals to fluid flow because each fluid Because many stratigraphic traps are elongated, the
phase added to a system results in decreased effective orientation of a trap’s maximum dimension relative to
permeabilities for all phases present (see Berg, 1975; structure has an important influence on whether the
Schowalter, 1979). Such “capillary seals” may be capa- holding capacity of the top or lateral seal will be
ble of indefinitely preventing fluid flow as long as the exceeded and, thus, on whether the trap will become
capillary threshold displacement pressure is not enclosed by capillary seals (Figure 2).
exceeded (Iverson et al., 1995). Whereas all hydrocar- Another way for a reservoir to achieve complete
bon accumulations are the result of capillary sealing capillary seal closure is for the low-permeability rocks
(Smith, 1966; Berg, 1975; Schowalter, 1979; Downey, surrounding it to contain multiphase pore fluids; cap-
1984), only those completely enclosed by capillary illary forces would then cause the low-permeability
seals fit the definition of a fluid-pressure compartment rocks to become impermeable. This is most likely to
(i.e., externally closed, internally open fluid system) occur in mature source rocks. Reservoirs that are
(Heasler and Surdam, 1992). closely associated with mature source rocks also prob-
ably have more than enough hydrocarbons available
to fill even very large (enclosed) reservoirs. Traps not
CONDITIONS THAT PROMOTE TOTAL closely associated with mature source rocks require
some form of long-distance secondary migration and
CAPILLARY SEAL CLOSURE AND THE have more of a supply problem. Reservoirs enclosed
DEVELOPMENT OF FLUID-PRESSURE within low-permeability rocks are, for the most part,
COMPARTMENTS excluded from long-distance secondary migration
pathways because the buoyant forces associated with
In order for a hydrocarbon trap enclosed by capil- secondary migration are generally insufficient to
lary seals to form, several conditions must be met. enable hydrocarbons to enter into and migrate
Capillary pressures exist in all multiphase fluid-filled through the low-permeability rocks around them.
rocks, but require the presence of low-permeability Because hydrocarbons follow the path of least resis-
(seal-type) rocks in order to trap significant amounts tance, once they begin migrating within the confines of
of hydrocarbons or allow anomalous pressures to a porous and permeable carrier bed, they usually con-
develop. Therefore, as a first condition, the trap must tinue to migrate within it until they either accumulate
be completely enclosed by low-permeability rocks in in conventional traps along the migration pathway or
order for total capillary seal closure to occur. That is, reach the surface.
,,,,,,,
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 225

COMPLETE CLOSURE INCOMPLETE CLOSURE Figure 1. Schematic


diagrams showing complete

,,,,,,,
vs. incomplete closure in
SECTION
CROSS- various types of traps.
VIEW
Contours are of reservoir-
FACIES TRAPS

type rock thickness in feet.

,,,,,,,
0
20
VIEW
MAP

40 30 20 10 0

,,,,
,,
,,,,
,,,,
UNCONFORMITY TRAPS

SECTION
CROSS-

VIEW

,,
,,,,
0 20
20 40
VIEW
MAP

,,,,,,,
,,,,
,,,,
,,
STRUCTURAL TRAPS

SECTION
CROSS-

VIEW

,,,
,,,,
0
VIEW
MAP

100 100

,,
50 75 50 45 45 50

Fault termination or facies change

Porous & permeable Low permeability


(reservoir-type rock) (seal-type rock)

Leakage of hydrocarbons from an underlying accu- associated with thick, high-quality source rocks. An
mulation, and subsequent migration through a poor exception to this generalization is, of course, stratigraph-
trap rock, may be the mechanism wherein some reser- ically enclosed reservoirs that are filled by long-distance
voirs enclosed within low-permeability rock become migration along either faults or fractures, and subse-
filled with hydrocarbons. For many reservoirs enclosed quently have been healed or diagenetically isolated.
in very low permeability rocks, however, a pressure in
excess of that due to buoyancy alone, such as that cre-
ated during hydrocarbon generation, may be necessary POWDER RIVER BASIN, WYOMING
to force the entry of hydrocarbons through the low-
General Setting
permeability rocks into the reservoirs (Figure 3). There-
fore, for a variety of reasons, complete capillary seal clo- The Powder River Basin of Wyoming and Montana
sure is more likely to occur in reservoirs closely contains >6000 m (>20,000 ft) of dominantly clastic
zzzz
,,,,,

yyyyy
226 Martinsen

R 76 W R 75 W R 74 W Figure 2. Schematic
cross sections of a “bar”
geometry sandstone

-3
20
,,,,,



yyyyy
zzzz
{{{{{
||||
reservoir oriented with

0
its long axis (A) parallel
T to structural strike and
46 (B) parallel to structural

-3
40
N dip. Changing the

0
,,,,,



yyyyy
zzzz
{{{{{
||||
orientation of the sand
body relative to structural
dip can change the
maximum feet of strati-
graphic closure from
~350 to >1600 ft.

,,,,,



yyyyy
zzzz
{{{{{
||||
-360
T

0
45

-3800
N

,,,,,



yyyyy
zzzz
{{{{{
|||| -40
00

,,,,,



yyyyy
zzzz
{{{{{
||||
44
N

B A
-4

T
20
-4

43
0
40
-46
-48

N
0
-50

00
00
-52

00
00

8000 FT 0 1 2 3 4 MI

rocks, ranging in age from Middle Cambrian to fluid compartments have been recognized. Nearly all
Oligocene. Figure 4 is a generalized stratigraphic of the overpressured fluid compartments occur within
nomenclature chart for the basin. This chart indicates the Cretaceous section; are restricted to the deeper,
names and ages of formations within the basin, as well more central portions of the basin; are associated with
as the major unconformities. Zones that produce mature to supermature source rocks; are normally
hydrocarbons, zones considered potential source compacted; and are hydrocarbon saturated. That is,
rocks, and zones in which anomalous pressures have they do not contain overpressured aquifers or even
been documented are also indicated. Approximately reservoirs with discrete hydrocarbon/water contacts.
2400 m (8000 ft) of this section, including the Permian– Although discrete fluid-pressure compartments have
Pennsylvanian Tensleep Sandstone through the Upper been identified in the Permo-Pennsylvanian section
Cretaceous Mesaverde Formation, is highly produc- (McBane, 1984; Sheppy, 1986), only a few, questionable
tive of hydrocarbons and has been extensively stud- anomalous pressures have been recorded (Heasler et
ied. Except at its margins, very little structural al., 1995). This apparent lack of fluid-pressure com-
deformation occurs within the basin; the majority of partments in the Paleozoic section is surprising
the hydrocarbon fields are stratigraphic in origin. because the Paleozoic rocks contain numerous strati-
Numerous fluid-pressure compartments have been graphically trapped hydrocarbons and should be at
mapped within the basin (Heasler et al., 1995). Over- higher temperatures and pressures than the Cretaceous
pressured, underpressured, and normally pressured rocks because they are deeper. Fluid compartments
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 227

Conventional
Traps

TOP OF
OVERPRESSURING

Fluid Pressure
Compartment Traps

direction of HC migration

immature organic-rich shale


porous & permeable reservoir
(single phase pore fluids)
mature organic-rich shale
HC accumulations
(multi phase pore fluids)
Figure 3. Diagram showing how sandstones enclosed in shale are more likely to become charged with hydro-
carbons if the shale that encloses them has generated hydrocarbons.

within the Minnelusa typically are identified by vary- The Lower Cretaceous section also consists mostly
ing pressure draw-down histories within fields, by the of shale, sandstone, and some conglomerate. Histori-
presence of virgin pressures in in-fill wells, and by cally it is the major hydrocarbon-producing interval in
varying water chemistries. the basin (Dolson et al., 1991), in part because the
Lower Cretaceous organic-rich Mowry Shale is
believed to be a major hydrocarbon source rock. Depo-
CRETACEOUS SECTION sition of the Lower Cretaceous rocks occurred in allu-
Basinwide Zone of Overpressuring vial plain to shallow-marine environments. Both the
Muddy and the Dakota are highly heterolithic reser-
The Upper Cretaceous section consists of ~1500 m voirs containing shale through pebble-size sediments.
(~5000 ft) of shale and subsidiary amounts of sand- The Lakota is lithologically more uniform and gener-
stone (Figure 5). For the most part, these rocks were ally consists of sandstone with some conglomerate.
deposited in a series of very broad, clastic shelf sys- Major shale units include the Mowry-Shell Creek
tems that intermittently prograded across the overlying the Muddy, the Skull Creek between the
Interior Seaway (Asquith, 1970). The Teapot, Park- Muddy and the Dakota, and the Fuson, between the
man, and parts of the Frontier are mostly shoreline- Dakota and the Lakota. Powley (1982, 1990) indicates
associated deposits, whereas the Sussex, Shannon, the Fuson is the basal seal to the Cretaceous overpres-
and other parts of the Frontier have been interpreted sured regime.
as shelf sand ridges (Gill and Cobban, 1973; Overpressuring in the Powder River Basin is
Merewether el al., 1979; Hobson et al., 1982; Cough- believed to be the result of both hydrocarbon genera-
lin and Steidtmann, 1984; Tillman and Martinsen, tion and maturation, especially the cracking of oil to
1984, 1987). The shoreline-associated sandstones gas (Jiao, 1992). Figure 6A is a west-to-east pressure
typically have wedge-shaped geometries that profile, and 6B is a gross lithofacies cross section of the
interfinger seaward with marine shales and land- Powder River Basin. The top of the overpressured
ward with various coastal plain deposits, and thus regime (yellow and blue region) more closely follows
are not generally enclosed within low-permeability stratigraphic boundaries and is deeper between
rocks. The ridge sandstones are typically encased in ranges 70W and 76W, where the section above the
marine shales. Steele Shale contains significantly more sandstone
228 Martinsen

WEST EAST WEST EAST

POPO
PLIOCENE AGIE FM

CHUGWATER GP
CROW MTN
ALCOVA LS

MIOCENE TRIASSIC
RED PEAK FM

OLIGOCENE
TERTIARY

WHITE RIVER FM WHITE RIVER FM SPEARFISH FM

GOOSE EGG FM

GOOSE EGG FM
ERVAY PINE SALT

EOCENE FORELLE LS
GLENDO SH
FORELLE LS
GLENDO SH
MONCRIEF

WAS
MINNEKAHTA LS MINNEKAHTA LS

FM
KINGSBURY CGL
WASATCH FM PERMIAN OPECHE SH OPECHE SH
FT UNION

FT UNION
TONGUE R TONGUE R
CONVERSE
PALEOCENE
FM

FM
LEBO SH LEBO SH LEO
TULLOCK TULLOCK
TENSLEEP SS
LANCE FM LANCE FM HELL CR
MINNELUSA FM
FOX HILLS SS FOX HILLS SS PENNSYLVANIAN
MESAVERDE

BEARPAW SH
TEAPOT SS AMSDEN FM
BELL SS

PARKMAN SS PIERRE
SUSSEX SS SHALE
MISSISSIPPIAN MADISON LS PAHSAPA LS

UPPER * SHANNON SS

STEELE SH
ENGLEWOOD FM
CODY SHALE

CRETACEOUS JEFFERSON
DEVONIAN FM

NIOBRARA SH NIOBRARA SH

CARLILE SH SAGE BREAKS SH


SILURIAN
* WALL CR TURNER SS
CARLILE SH
GREENHORN LS

FRONTIER FM BIGHORN DOL WHITEWOOD DOL


BELLE FOURCHE SH
ORDOVICIAN
UNNAMED SS ROUGHLOCK,
MOWRY SH MOWRY SH ICE BOX, ALADDIN

SHELL CREEK SH NEFSY SH

LOWER * MUDDY SS NEWCASTLE SS

GALLATIN
DEADWOOD
THERMOPOLIS SH SKULL CREEK SH
UPPER LS
FM

CRETACEOUS DAKOTA SILT DAKOTA SILT GROS VENTRE


CAMBRIAN

FM

*
KARA GP

KARA GP

FALL RIVER SS FALL RIVER SS


INYAN

INYAN

SS
FUSON SH FUSON SH D
EA
LAKOTA CGL LAKOTA CGL ATH
FL
MIDDLE
MORRISON FM MORRISON FM
SUNDANCE FM

SUNDANCE FM

UPR SUNDANCE REDWATER SH


JURASSIC

UPPER LAK MBR LOWER


HULETT SS
LWR SUNDANCE STOCKADE BEAVER
CANYON SPRINGS PRECAMBRIAN
GYPSUM SPR GYPSUM SPR
MIDDLE
LOWER

= Hydrocarbon producing zone


= Known source rock
= Contains overpressured reservoirs
*
Figure 4. Stratigraphic nomenclature chart for the Powder River Basin. Modified from
Wyoming Geological Association (1991).

than elsewhere. Probably considerably less gas and and overpressuring that may have developed is more
pressure were generated where sandstones and silt- likely to have been dissipated by flow within the thick,
stones are abundant because they contain less organic laterally continuous sandstone zones. Thus, although
material than marine shales. In addition, whatever gas the top of overpressured zone does not strictly follow
zzzz
 ,,,
yyy ,,
,,,,,,,,,,
yyyyyyyyyy
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 229

WEST

,,,,,,,,,,
yyyyyyyyyy
,,,,,,,,,,
yyyyyyyyyy
EAST



zzz
|||

zz

|| ,,
yy
{{
,
y


{ {
|
{ 
Lance Fm Lance Fm

,,,,,
Fox Hills SS Fox Hills SS

MESAVERDE Lewis Sh


|||
{ {

Teapot SS

Parkman SS
UPPER CRETACEOUS

Pierre Sh
Sussex SS
STEELE SH

Shannon SS

,,,,,
|{
{ 
|
{ 
|
{ 
Steele Sh

,,,,,
,,,,,,
Niobrara Sh

yyy
,,,
Niobrara Sh

,, ÀÀÀ
€€€
@@@
,,, ,,,
yyy

CARLILE SH
| ,,
|
{

y
z
y
 ,

zz
y

,,
z z

,
y
Carlile Sh Sage Brush Mbr

,,,,,,
Wall Ck SS(1st Frontier) Turner SS

,,
FRONTIER FM

Emigrant Gap Mbr Pool Creek Mbr


Greenhorn LS

,,,,,,,
Belle Fourche Mbr
Belle Fourche Mbr

,
zz,
yy

,,
y
,

z z,
y
,
y

||
,,,,,,,
,

z
y

{
Mowry Sh Mowry Sh
CLOVERLY GP THERMOPOLIS SH

Shell Creek Sh Netsy Sh


LOWER CRETACEOUS

Muddy SS Muddy SS

,,,,,,,
z|,,y,{,,,,,,,
|{ÀÀ€€@@,,ÀÀ€€@@,,,,{|
Thermopolis Sh Skull Creek Sh

,,,,,,,
, ,,,,,,
,,,,,,, CLOVERLY GP
Fall River SS Fall River SS
(Dakota SS) (Dakota SS)

,,,,,,,
Fuson Sh Fuson Sh
Lakota Cgl Lakota Cgl

yy
,,
,,
yy
dominantly sandstone dominantly shale

,,
variable clastic lithologies limestone and/or
calcareous shales
dominantly sandstone
and/or conglomerate hiatus
unconformity
Figure 5. Chronostratigraphic nomenclature chart for the Cretaceous Powder River
Basin, with unconformities and lithofacies indicated.

stratigraphy, it appears to be strongly influenced by it. Furthermore, the base of the overpressured sec-
The higher, nearly horizontal top of the anomaly in the tion follows stratigraphy throughout the basin and
eastern and extreme western parts of the basin occurs appears to be bounded by the Fuson Shale. Whereas
within shale and may represent the top of the gas- the section between the Fuson and the top of the
generating zone. Steele is comprised dominantly of shale (with only
230 Martinsen

(A) Figure 6. (A) West-to-east


W 79 78 77 76 75 74 73 72 71 70 69 68 67 66 65 64 63 62 E cross section through
the Powder River Basin

le
4000 showing the distribution

ee
and relative magnitudes of

St

a
overpressuring. (From

ar
br
6000 Maucione et al., 1995, their

io
figure 16.) (B) West-to-east

N
Depth (ft)

n cross section through the


8000
uso Powder River Basin
F showing the gross
lithofacies (dominantly
10000 sand/dominantly shale)
and the outline of
overpressuring as indicated
12000 in Figure 6A. Note that the
section between the top of
Steele down to top of Fuson
14000 is dominantly shale with
110 100 90 80 70 60 50 40 30 20 10 0
zones of relatively thin,
Distance (mi) isolated sand, and the
section above the Steele is
much more sand rich in the
western part of the basin.
Increasing Anomaly
(B) (Overpressure)
W 79 78 77 76 75 74 73 72 71 70 69 68 67 66 65 64 le 63 62 E
4000
ee
St

a
ar
br

6000
io
N
Depth (ft)

n
8000
uso
F

10000
Outline of
Overpressuring
12000

14000
110 100 90 80 70 60 50 40 30 20 10 0
Distance (mi)

minor amounts of sandstone beds), the section below Reservoir-Scale Fluid-Pressure Compartments
the Fuson is dominated by sandstone and siltstone.
Although Lakota sandstones and conglomerates are Within the regional zone of overpressuring, numer-
sometimes enclosed in shales, the shales tend to be ous reservoir-scale fluid-pressure compartments have
red or green (J. Dolson, 1994, personal communica- been identified. The boundaries of those fluid-
tion), and are probably not organic-rich. Therefore, pressure compartments that have been studied coin-
overpressuring is less likely to occur below the cide with lithofacies boundaries.
Fuson because it contains fewer organic-rich rocks in Examples from the Muddy Sandstone, one of the
which to generate excess pressures and an abun- most highly fluid compartmentalized formations in the
dance of sandstone conduits to equilibrate excess basin, demonstrate stratigraphic controls on the devel-
pressures. opment of reservoir-scale fluid-pressure compartments.
,,,,,,,,
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 231

Figure 7. Schematic

,,,,,,,,
representation of
three scales of stratigraphic
A compartmentalization
within the Muddy Sandstone.

,,,,,,,,
(From Martinsen, 1995.)
LATERAL
SEALS

,,
VERTICAL SEAL

LATERAL

,,,,,,,,
Rozet-unconformity surface

paleosol.

,,,,,,,,
B

,,,,,,,,
,,,,,,,,
VERTICAL SEALS

,,,,,,,,
Transgressive shales

,,,,,,,,,,,
yyyyyyyyyyy ,,,,,,,,
yyyyyyyy
yyyyyy
,,,,,, ,,,,,,,
yyyyyyy ,,,,,,,,,,,
yyyyyyyyyyy ,,,,,,,,
yyyyyyyy
,,,,,, yyyyyyy
yyyyyy ,,,,,,, yyyyyyyyyyy
,,,,,,,,,,,,,,,,,
yyyyyy
,,,,,,,,
yyyyyyyy

,,,,,,,,
,,,,,,
yyyyyy
C
,,,,,
yyyyy ,,,,,
yyyyy ,,,,,,,,,,,,
yyyyyyyyyyyy ,,,,,,
yyyyyy
,,,,,yyyyy
yyyyy ,,,,, ,,,,,,,,,,,,
yyyyyyyyyyyy
,,,,,,,,,,,
yyyyyyyyyyy ,,,,,,
yyyyyy
,,,,,, yyyyyy
,,,,,,,,,,,yyyyyy
,,,,,,
,,,,,yyyyyyyyyyy
yyyyy ,,,,,,
yyyyyy

,,,,,,,,
,,,,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyyyyy
,,,,,,
yyyyyy
,,,,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyyyyy ,,,,,,,
yyyyyyy
,,,,,,
yyyyyy
,,,,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyyyyy
,,,,,,
yyyyyy ,,,,,,,
yyyyyyy ,,,,,,,
yyyyyyy ,,,,,,,
yyyyyyy
,,,,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyyyyy
,,,,,,
yyyyyy ,,,,,,,
yyyyyyy ,,,,,,,
yyyyyyy
,,,,,,,
yyyyyyy ,,,,,,,
yyyyyyy
,,,,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyyyyy
,,,,,,
yyyyyy
,,,,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyyyyy ,,,
yyy,,,
yyy ,,,,,,
yyyyyy
,,,,,,,,,,,,,,,,,,yyy
yyyyyyyyyyyyyyyyyy ,,,
yyy ,,,,,,,,,,,
yyyyyyyyyyy
,,,yyyyyyyyyyy
,,,,,,,,,,,
,,,,,
yyyyy
,,,,,,
yyyyyy
,,,,,
yyyyy
,,,,,,
yyyyyy
,,,,,,,,,,,,,,,,,,yyy
yyyyyyyyyyyyyyyyyy ,,,yyyyyyyyyyy
,,,,,,,,,,,
,,,,,,,,,,,
yyyyyyyyyyy ,,,,,
yyyyy
,,,,,
yyyyy
,,,,, Various Ss lithologies
yyyyy
,,,,,
yyyyy

For an in-depth understanding of Muddy Sandstone unconformity, which divides the Muddy into two sepa-
stratigraphy and depositional history in the Powder rate stratigraphic sequences. Diagenesis beneath the
River Basin, see Gustason (1988), Gustason et al. (1988a, unconformity created a rock with excellent sealing
b), Dolson et al. (1991), and Martinsen (1995). capacity, even in areas where sand is in contact with
At least three scales of stratigraphic compartmen- sand across the unconformity (Jiao, 1992). The overall
talization occur within the Muddy Sandstone (Figure geometries of the compartments formed by the Rozet
7). Each of these three scales appears to coincide with unconformity are defined by the intersection of the
a different scale of fluid-pressure compartmentaliza- unconformity (or the associated paleosol) with either
tion observed in the formation. The largest scale of the underlying Skull Creek Shale or the overlying Shell
compartmentalization is provided by the Rozet Creek Shale. The middle scale of compartmentalization
232 Martinsen

Figure 8. Map of the


120 Powder River Basin
90 Hammond showing the location of
BC T
Muddy Sandstone oil and
9 gas fields, thickness of the
90 MONTANA Fence Creek Bell Creek
WYOMING Sandbar T
S Muddy, and the location
120 Ute LM 57 of mapped lineaments.
90 R64W
N Isopach values taken from
Recluse Dolson et al. (1991, their
90
Oedekoven Collums SR figure 11); lineament data
RZ taken from Slack (1981, his

60
Gas Draw
figure 4) and Martinsen and
30 Marrs (1985, their figure 1).
Amos Draw Kitty
Springen Ranch
SC
60

T
50
R64W
GB
N
Lazy B S-Bar Coyote Creek
Slattery 30
60
FC
Fiddler Creek
Trend
30
Buff CT R61W
T
45
Hilight N
Ha Creek
Clareton Area
Porcupine Finn 60

60

Teapot
Steinle Ranch
BD

WH

T
30 Powell 38
R61W N
GI

Cole Creek
60

30
60

Glenrock WYOMING
T
31
N
R82W R71W

is provided by the shales that overlie the transgressive capillary seal closure. The proximity of the Muddy
surfaces of erosion and further divide the stratigraphic Sandstone to potential source rocks (the Mowry Shale
sequences into individual parasequences or parase- and possibly the Skull Creek Shale) and the generally
quence sets. The geometries of these compartments are limited size of Muddy Sandstone reservoirs provide
defined by the intersection of the shales (either by favorable conditions for the reservoirs to completely
onlap or truncation) with the Rozet unconformity or fill with hydrocarbons. Stratigraphically, therefore, the
paleosol. The third and smallest scale of coincident Muddy Sandstone is strongly predisposed to the for-
stratigraphic and fluid compartmentalization is pro- mation of internal reservoir-scale fluid-pressure com-
vided by the distribution, geometry, and stacking of partments. Observations indicate that with one
individual lithosomes within each of the parase- exception, all of the Muddy Sandstone fields in the
quences sets. Lithosome geometries are highly deep part of the Powder River Basin have the charac-
variable and reflect variations in depositional environ- teristics of fluid-pressure compartments; that is, they
ment and sediment type and supply. are anomalously pressured, depletion-drive reservoirs
As a result of these three scales of stratigraphic com- that lack free-water legs.
partmentalization, nearly all sandstones within the The one exception is Lazy B Field (T49N R74W;
Muddy Sandstone are enclosed by low-permeability Figure 8). It is interesting to note that this field is in
rocks, and therefore meet the first condition for total close proximity to the Amos Draw field (T51N R75W,
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 233

AMOS DRAW WILD HORSE CREEK LAZY B


AMOS DRAW #3 1-27 1-10P LOWERY
29-51N-75W 27-50N-74W 10-49N-74W
~5600 psi 2237 psi 3416 psi
NW SE

|||||||

,,,,,,,,

zzzzzzz
yyyyyyyy

{{{{{{{{
||
{{

 {


|{

9500

yyyyyyyyyyyyyyy
,,,,,,,,,,,,,,,
,,,,,,,,,,,,,,,
,yyyyyyyyyyyyyyyy


{{{{{{{{
|||||||
TOP OF MUDDY
UPPER Kmd

,,,,,,,,,,
yyyyyyyyyy
ROZET
10,000 UNCONF

,,,,,,,,,,
yyyyyyyyyy ,,,,,,,,
yyyyyyyy
,,,,,,,,,,,,,,,
yyyyyyyyyyyyyyy
ROZET UNCONF 9600
LB UNCONF

,,,,,,,,,,
yyyyyyyyyy ,,,,,,,,
yyyyyyyy
LOWER MUDDY
SK UNCONF 9800
LB UNCONF

,,,,,,,,
yyyyyyyy
,,
yy
|,
z{
y

? ?
overpressured SK UNCONF

Gr Sp
,,,,,,,,
yyyyyyyy
,,,,,,,,
yyyyyyyy
,,
yy
?

,,
yy
10,100
SOIL ZONE normally

,,
yy
pressured
SUB-UNCONFORMITY MARINE SS
9900
VALLEY-FILL/TRANSGRESSIVE DEPOSITS
PERFORATED INTERVAL

Figure 9. Cross section showing the stratigraphic relationships of the Muddy Sandstone between the Amos
Draw and Lazy B fields.

Figure 8), a highly overpressured gas condensate or overpressured, as is Amos Draw field. It is less rea-
reservoir at the same depth that produces from the sonable to assume that, unlike all the other Muddy
same stratigraphic zone within the Muddy Sandstone Sandstone fields in the “deep” Powder River Basin,
(Figure 9). Lazy B, however, is normally pressured to there were insufficient hydrocarbons available to fill
underpressured, produces high-gravity oil (51° API, the Lazy B field reservoir.
1314 GOR), and contains both a gas cap and an Studies throughout the basin indicate that, like the
oil/water contact (Wyoming Geological Association, Muddy reservoirs, many of the other Cretaceous reser-
1981). Mercury injection capillary pressure data from voirs identified as fluid-pressure compartments are
Amos Draw suggest that the seal (a paleosol) can trap enclosed by low-permeability rocks; these low-perme-
an ~630-m (~2100-ft) gas column or a 540-m (1800-ft) ability rocks are also potential source rocks, and the
oil column (Jiao, 1992). Because the sandstone reser- structural closure of the reservoirs is limited. For more
voir in Amos Draw field has ~255 m (850 ft) of closure, on stratigraphic and fluid-pressure compartmentaliza-
it is well below the maximum calculated holding tion in the Muddy, see Martinsen (1995).
capacity of 540–630 m (1800–2100 ft) for the seal. The
reservoir in Lazy B, in contrast, has ~660 m (~2200 ft)
of closure and exceeds the maximum calculated hold- PERMO-PENNSYLVANIAN SECTION
ing capacity of the seal. Even though the Lazy B reser-
voir is volumetrically smaller than the Amos Draw The Permo-Pennsylvanian Minnelusa is currently
reservoir, it has more closure because it is linearly the largest producer of hydrocarbons in the Powder
shaped and oriented with its long axis perpendicular River Basin, and is second only to the Tensleep Forma-
to structure. It is also possible that the holding capac- tion (its partly chronostratigraphic equivalent) in
ity of the top seal in the area of Lazy B field has been terms of hydrocarbon production for the state of
reduced. Lazy B field coincides with the Springen Wyoming. Cumulative production is >200 million bbl
Ranch lineament (Figure 8). Subtle basement block of oil (Moore, 1983; Clayton and Ryder, 1984) and
movement along this lineament is believed to have comes mainly from oil fields in the northern Powder
controlled reservoir rock deposition and preservation River Basin (Campbell, Crook, and Johnson counties).
(Gustason, 1988). Any later movements along this Nearly all the Minnelusa fields are partly to entirely
basement block fault may have caused fracturing in stratigraphic in origin. Although discrete fluid
the overlying seal and reduced its holding capacity. compartments have been identified in the Permo-
Whatever the reason, the holding capacity of the seal Pennsylvanian section (McBane, 1984; Sheppy, 1986;
in the Lazy B field apparently is not sufficient to Martinsen, 1992), there is no zone of regional over-
allow the field to be either filled with hydrocarbons pressuring, and only a few questionable anomalous
234 Martinsen

(A) Figure 10. (A) Structure


contour map of the Powder

y,{{y, 
40 45 50 55 60 5
9 MONTANA
WYOMING
River Basin on the top of
0' the Niobrara Formation

-2
00
10 (modified from Maughn,

0
55
A 1990). Also displayed are the


,,,

20
0' Upper distribution of oil fields that
Minnelusa produce from the Minnelusa,
Production
the thickness of the upper
5 Minnelusa, the regressive
SC
50 phase outline of the Lusk
-60
embayment during upper



,,,

00
Minnelusa time, and
1 lineament data (Clayton and

+2
Ryder, 1984, their figure 3).

00
0
0' 1
40 Isopach values are taken
45
-2 0

from Trotter (1984, his figure




,,,
00

9); lineament data are from


5
Slack (1981) and Martinsen
Tensleep
Sandstone and Marrs (1985); Lusk
Lusk
Production
Embayment
embayment boundaries are
40
based on Fryberger (1984).


,,,
zz
yyy 

,,,

0
00
+2

10
-80
00

35 S. DAKOTA
NEBRASKA



||
{{{ 
10
,0 00
-

Middle A'
Minnelusa
30 Production 30

{| 

0
00
+2

80 75 70 65 60

000 Structural elevation, in feet, top of the Niobrara Formation


+2
' Thickness, in feet, upper Minnelusa member
200
Precambrian Lusk Embayment

SC
South Coyote Creek lineament

pressures have been recorded (Heasler et al., 1995). Fluid compartments within the Minnelusa typi-
Furthermore, although the Minnelusa is significantly cally are identified by varying pressure draw-down
lower in the stratigraphic section than Cretaceous histories within fields, by the presence of virgin
rocks, hydrocarbons within it typically are low-gravity pressures in in-fill wells, and by varying water
(average <30° API), low-GOR (<100) oils, and thus do chemistries (McBane, 1984; Sheppy, 1986; Tyler and
not appear to be more mature than the Cretaceous Modroo, 1986). It is possible, however, that fluid
hydrocarbons (See MacGowan et al., 1993). Minnelusa compartments identified in this manner only appear
fields typically have mappable oil/water contacts, lack to be isolated because not enough time has elapsed
basal seals, and commonly have water drives that for pressure variations formed as a result of fluid
result in high per-well recovery of oil in place. removal during production to equilibrate. This pos-
In several fields having depletion-drive reservoirs, sibility is supported by longer term data, which indi-
a tar mat has been recognized and proposed as the cate at least some fields (primarily lower “B” sand)
basal seal (Trotter, 1984). The presence of a tar mat that experienced rapid initial pressure drops are in
indicates degradation of oils, either by water-washing pressure continuity with other fields (P. Slack, 1993,
or bacteria, and suggests that water below the oil col- personal communication). Therefore, although the
umn in those fields is probably meteoric in origin and Minnelusa, like the Cretaceous section, is strati-
thus in hydrodynamic continuity with the surface graphically complex and contains numerous
water system. stratigraphically trapped hydrocarbon accumulations,
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 235

{y,{y,,,, 
(B) Figure 10. (B) Structure
40 45 50 55 60 5 contour map of the Powder
MONTANA River Basin showing the
WYOMING
distribution of oil fields

-2
00
10 that produce from the

,,,



0'

55 Minnelusa, the thickness


20

-4
0 Upper of the middle Minnelusa,
00 Minnelusa the regressive phase outline
Production
of the Lusk embayment
0'
30 ' 5 during middle Minnelusa
300
time, and the line of cross

{ ,,,


50
-60 section shown in Figure 11
00
(Clayton and Ryder, 1984,
their figure 3). Isopach

+2
Lusk 1

00
values are from Trotter

0
Embayment
1
(1984, his figure 9); Lusk

,,,



45
-2 0

embayment boundaries
00

are based on Desmond


5
et al. (1984).
Tensleep
Sandstone
Production

,,,



40


yyy
zz
0
00
+2

10
-80
00

35 S. DAKOTA

{,y{{{
,,,

yyy


zz
||
NEBRASKA

10
,0 00
-

Middle
Minnelusa
30 Production

,yz{| ,
30
0
00
+2

80 75 70 65 60

000 Structural elevation, in feet, top of the Niobrara Formation


+2
' Thickness, in feet, middle Minnelusa member
300

Precambrian Lusk Embayment

the reservoir and production characteristics of Min- The Minnelusa is composed of several lithologies,
nelusa fields are very different from those of Creta- including sandstone, carbonate, and evaporite, and is
ceous fields. the product of numerous transgressive-regressive
cycles of alternating shallow-marine to eolian deposi-
Regional Stratigraphy and Depositional Setting tion. Lithofacies within the upper and middle members
are very similar, except that the middle member con-
The Minnelusa varies from 195 m to >375 m tains numerous thin but widespread radioactive, car-
(650 ft–>1250 ft) in thickness, and has been informally bonaceous black shales. These black shales can be
divided into three members simply referred to as traced throughout much of the southeastern two-thirds
upper, middle, and lower Minnelusa (Foster, 1958). of the Powder River Basin, and have been identified as
The contacts between the three members are uncon- possible hydrocarbon source beds for the Minnelusa
formable, and numerous disconformities are believed (Tromp, 1981; Tromp et al., 1981; Clayton and Ryder,
to exist within each of the members (Moore, 1983; 1984). Several authors (Moore, 1983; Desmond et al.,
Desmond et al., 1984; Fryberger, 1984; Trotter, 1984). 1984; Fryberger, 1984) have suggested that the Persian
The majority of the oil fields produce from the upper Gulf tidal flat (sabkha) deposits described by Shinn
member, although significant production also comes (1983) are a modern analog for the Minnelusa.
from the middle member (Figure 10A, B). The lower Gross lithofacies distributions within both the
member is relatively barren of oil. upper and middle Minnelusa were influenced by a
236 Martinsen

paleogeographic feature called the Lusk embayment relief; however, although numerous beds pinch out
(Figure 10A, B). Because of the very low topographic beneath the Opeche unconformity, these pinchouts
relief of the area, small changes in relative sea level commonly occur parallel to structural dip, and thus
produced large changes in the lateral extent of the have little closure. Relatively few terminations occur
embayment. During transgressive periods, this perpendicular to structural dip.
embayment extended across most of what is now the Although less extensive than vertical seals, several
Powder River Basin and was the site of extensive sub- kinds of lateral seals are present in the Minnelusa. In
tidal carbonate deposition. During regressive periods, the southern and eastern portions of the basin (area of
it contracted extensively to form a smaller, more lin- middle Minnelusa production; Figure 10B), where
ear, northwest-trending embayment, and sand dunes sand supply was more limited, dunes are often iso-
migrated out across the now-exposed carbonate flats. lated (similar to “starved” ripples), and there is a
The Lusk embayment alternately expanded and con- higher proportion of low-permeability interdunal and
tracted many times during Minnelusa deposition and shallow-marine sediment. Thus, reservoirs in this area
produced a high-periodicity alternation of sandstone, are more likely to be enclosed in low-permeability
carbonate, and evaporite that resulted in vertical parti- rocks. Major lithofacies changes along the western
tioning of the formation. edges of dune complexes provide good lateral seals in
For the most part, the carbonates and evaporites the northeastern portion of the basin. In the northern
deposited within the Lusk embayment comprise seals, portion of the basin, relief along the Opeche unconfor-
whereas the dominantly aeolian-deposited sandstones mity shows locally isolated dune deposits in the
comprise the majority of the reservoirs. Beyond the uppermost Minnelusa, especially in areas where the
maximum transgressive extent of the Lusk embay- unconformity cuts down through beds of carbonate or
ment, the vertical lithologic succession of the evaporite. Lateral seals also occur where cementation
Minnelusa is very sand rich. Wherever the Lusk has severely reduced primary porosity and permeabil-
embayment was continuously present, the entire verti- ity within depositionally continuous sandstones. Such
cal lithologic succession consists of low-permeability diagenetically formed seals have been recognized as
rocks that effectively separate sandstone reservoirs providing traps for several Minnelusa oil fields (Fry-
present to the east of the embayment from those to the berger, 1984; George, 1984). Although both calcite and
west. During upper Minnelusa time, the northern anhydrite cementation are pervasive within the Min-
boundaries of the Lusk embayment during regressive nelusa, most of the known upper Minnelusa oil fields
phases are located in the north-central Powder River lack bottom seals (J. Dolson, 1994, personal communi-
Basin, and appear to coincide with the Springen Ranch cation) and have oil/water contacts, which suggests
lineament (Figure 11). Therefore, only the southern that cementation did not result in isolation of many oil
area of the Powder River Basin has a continuous verti- reservoirs.
cal succession of low-permeability rocks separating In summary, the Minnelusa is a highly complex
eastern from western areas of porous and permeable stratigraphic unit, dominantly containing stratigraphi-
sandstones in the upper Minnelusa. During middle cally confined sandstone reservoirs rather than iso-
Minnelusa time, the northern boundaries of the Lusk lated sandstone reservoirs. Although stratigraphically
embayment were north of the Powder River Basin, isolated sandstones may exist throughout the Min-
even during regressive phases, so an area of low- nelusa, for the most part only middle Minnelusa reser-
permeability rocks exists across the entire basin, which voirs in the southeastern Powder River Basin are both
effectively separates eastern areas of porous and per- stratigraphically isolated and productive, and have
meable sandstone from western areas. been identified as fluid-pressure compartments. Obvi-
ously, the development of pressure compartments in
Types and Characteristics of Stratigraphic the Minnelusa is dependent on more than strati-
Compartments graphic complexity, including enclosure of high-
porosity, high-permeability rocks by relatively
The intercalation of various clastic, carbonate, and
low-permeability rocks.
evaporite lithologies that typifies the Minnelusa, plus
numerous intraformational unconformities, have
Stratigraphic Compartments, Source Rocks,
resulted in a highly complex Minnelusa stratigraphy
and Oil Migration Routes
in which multiple reservoir-trap combinations are
expected (Mallory, 1967; Desmond et al., 1984; Fry- Multiple source rocks for the Minnelusa Formation
berger, 1984; Trotter, 1984; Desmond, 1985). However, oil fields have been proposed, and hydrocarbons with
although this complex stratigraphy has resulted in a varying genetic profiles have been identified. Until the
high degree of vertical partitioning, lateral partition- 1980s, long-distance migration of oil from the Phos-
ing of the sandstone reservoirs is less extensive. In the phoria Formation of western Wyoming into eastern
western and northern portions of the basin (areas of Wyoming was believed to be the only source of Min-
Tensleep and upper Minnelusa production; Figure nelusa oil (Sheldon, 1967; Stone, 1967). Subsequently,
10A, B), the large dune fields deposited extensive sand Tromp (1981) and others (Clayton and Ryder, 1984;
sheets that do not have an abundance of lateral seals. Momper and Williams, 1984; James, 1989) identified
The Opeche unconformity, which forms the upper the organic-rich black shales within the middle Min-
contact for the Minnelusa, often shows considerable nelusa as an indigenous source for the oils.
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 237

,,,,,,,,,,

DAKOTA
SOUTH
WYOMING NEBRASKA

,,,
yyy
yyyyyyy
,,,,,,,,
yyyy
,,,
B

,,,,,,,,,,,,,,
,,,,
,
,,,,, ,,
,,,,
,,,,,, ,,,,,,
,,,,, ,,
NORTH

,,,,,,,,,,
,,, ,,,
46 N R 64 W
B'
55 N W 49 N R 68 W 42 N R 62 W 37 N R 62 W SOUTH

yy
,, yyyy
,,,,
,,,
yyy ,,,,
yyyy
39 N R 61 W

,,,,,,,
yyyyyyy
51 N 67 W 12 S R E

,,,,,,,,,,,,,,

EGG FM
Minnekahta

GOOSE
mbr

,
,,,,,
,,,,
,,,,,,
,,,,, ,,
,,,,,
,,,
,,,,,
,
,,,,,
, ,,
,,,,
,
,,, , e
Opeche

nd A
yyy
,,,
yy,,,
,,yyy
,,,,
yyyy on
e

st
nd B
on

,,,,,,,
yyyyyyy
yyyyyy
,,,,,,y
,

PERMIAN
Shale

,,,,,,,,,,,,,,
,,,,,,,,
st

sa

ne
to
e
sa
nd C
on

ds
st

n
e
sa
rs
upper

ve
sa

,, yyy
,,,
,,,,,,,,
on
mbr

,,,,,,,,
,,,,, ,,,,,,,,
,,,,,,,,,,,,,,
,,,,,
, ,,,
,,,,,, ,
C

MINNELUSA FM
st eo
e
on
nd L
,,,,,,,,
,,,,,,,,,,,,,,
middle

PENNSYL-
sa
, , y,y, ,,,,,,,
,,,

VANIAN
mbr

,,,,,,,,
lower
mbr
EXPLANATION

SIPPIAN
MISSIS-
Sandstone Dolomite Anhydrite Shale Shale Unconformity
and and (black
limestone siltstone and gray) FEET METERS
(red and green) 200 500
100
25 MILES

SCALE 50 KILOMETERS

Figure 11. Northwest-southeast well-log cross section through the Minnelusa Formation showing erosional
truncation of upper Minnelusa sandstones to the northwest and facies terminations to the southeast. Note that,
with the exception of the Leo sands, sandstones are not enclosed by low-permeability rocks. Line of section
shown in Figure 10B (Clayton and Ryder, 1984, their figure 2).

The fact that many sandstones within the Min- After filling these structural traps, additional migrat-
nelusa are laterally continuous parallel to structural ing hydrocarbons would have spilled updip to the east
dip makes long-distance migration tenable. Within the until they encountered the western margin of low-
Powder River Basin, Minnelusa production occurs in permeability facies deposited within the Lusk embay-
three distinct geographic areas (Figure 10A, B). In the ment. These low-permeability rocks provided an effec-
western Powder River Basin, production is all struc- tive barrier that prevented the hydrocarbons from
turally controlled and is from the upper Minnelusa. In migrating further east (i.e., no upper Minnelusa pro-
the northeastern part of the basin, the fields produce duction in the southeastern Powder River Basin).
mostly from the upper Minnelusa, and all occur in a Instead, the low-permeability facies deflected the
fairway whose southern boundary coincides with the migrating hydrocarbons to the northeast, where they
northern “regressive” phase boundary of the Lusk accumulated in various stratigraphic and combined
embayment. In the southeastern Powder River Basin, structural-stratigraphic traps in what is now the north-
the fields typically produce from the middle Min- eastern Powder River Basin (Figure 10A). Stratigraph-
nelusa (Leo sands) and geographically coincide with ically isolated reservoirs in the upper Minnelusa
the best development of the organic-rich black shales. generally were excluded from the long-distance
Stratigraphically and geographically, this distribution migration route, and therefore did not fill. Today,
of oil fields accommodates a model that entails long- many of these stratigraphically isolated reservoirs are
distance migration for upper Minnelusa oils (western either wet or tight. Productive reservoirs that appear
and northeastern areas of production) and a local, to be stratigraphically isolated (e.g., a dune sandstone
indigenous source for middle Minnelusa production surrounded by dolomite) could have been filled as a
(southeastern area). result of fractures providing migration pathways.
Within the context of this model, hydrocarbons Truly fluid-enclosed reservoirs may also exist in reser-
migrating eastward from western Wyoming would voirs diagenetically isolated after migration and accu-
have first filled structural accumulations in the west- mulation by either cementation or tar-mat formation
ern Powder River Basin, such as Reno and Reno East. below the hydrocarbon-water contact. These types of

,,
,,
,,,,


,
,,,,



,
,,,,,,,
,,,,,,,



238

,
,,


,
,,,,
Martinsen


 

,,,,,,

 
,,,, ,,
,,,,,,,,,
,,,
,,,,,,,



,,
,,,,,,
,,,
Opeche


 ,,
,,,,,,,,,
, ,,,

Upper Minnelusa
,
,,,,,
,,,,
,,,,,,,,,
,,,,
,,,,,,,




 ,
,,,
,,,,,,,,,
,,,,,


 ,,,,,,,,,
,,,,,,,




,, 
,,,y,
, , y,y,
,,,,,,,,,
,

,
y
y
,
y

Middle Minnelusa
,,,,,

,
 ,
,,,,,,,,,




,
 
,
 ,,
,,,,,,,,, Migration path

,, Sandstone Opeche shale Oil accumulation


,,
,,

Tightly cemented Tight carbonates Organic-rich


sandstone black shale Unconformity

Figure 12. Schematic representation of the various types of stratigraphic traps in the Minnelusa
Formation.

reservoirs could correspond to pressure compart- within the middle Minnelusa, extend north-to-south
ments, and could be overpressured by the cracking of across the entire Powder River Basin) (Figure 10B).
oil to gas within the reservoir. According to Mac- It is reasonable to assume that the intraformational
Gowan et al. (1993), this probably only occurred in the black shale in the middle Minnelusa is the most likely
deepest parts of the Powder River Basin. hydrocarbon source for the fields in the southeastern
Middle Minnelusa reservoirs are unlikely to have Powder River Basin. Oils expelled under high pres-
been filled by long-distance migration of hydrocar- sure from these black shales were injected into the
bons for a number of reasons. First, hydrocarbons most proximal porous and permeable reservoir beds,
migrate updip along the tops of carrier beds and not which include both the stratigraphically confined and
throughout the entire thickness of the carrier beds the stratigraphically isolated dune sandstones of the
(Schowalter, 1979). Unless there are low-permeability middle Minnelusa. Most of the data suggesting reser-
beds within the Minnelusa that are continuous from voir compartmentalization within the Minnelusa per-
western to eastern Wyoming, any hydrocarbons tain to middle Minnelusa reservoirs (e.g., Buck Creek
migrating from western Wyoming most likely would Field) (Martinsen, 1992).
have been within the upper Tensleep/upper Min- Thus, although the ratio of stratigraphically isolated
nelusa interval, and not within the middle Minnelusa. to confined reservoirs is much lower in the Minnelusa
Furthermore, even if hydrocarbons did migrate from than in most of the Cretaceous rocks, stratigraphically
the west within the middle Minnelusa interval, they isolated reservoirs do occur. However, only where
would have been prevented from entering eastern these reservoirs are closely associated with potential
middle Minnelusa sandstone reservoirs by the low- source rocks (middle Minnelusa member of the south-
permeability facies of the Lusk embayment (which, eastern Powder River Basin) has evidence of fluid
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 239

compartmentalization been found. A diagrammatic Formation (Permian and Pennsylvanian), Powder


summary of types of stratigraphically defined oil River Basin, Wyoming, in Hydrocarbon Source
fields in the Minnelusa is presented in Figure 12. Rocks of the Greater Rocky Mountain Region:
Rocky Mountain Association of Geologists,
p. 231–253.
CONCLUSIONS Coughlin, J.P., and J.R. Steidtmann, 1984, Depositional
environment and diagenesis of the Teapot Sand-
Both the Cretaceous Muddy Sandstone and the stone, southern Powder River Basin, Wyoming:
Permo-Pennsylvanian Minnelusa Formation display a Mountain Geologist, v. 21, p. 91–103.
high degree of stratigraphic complexity. While this Desmond, R.J., Jr., 1985, Stratigraphic and depositional
complexity resulted almost exclusively in the forma- environment of the middle member of the Minnelusa
tion of sandstone reservoirs enclosed by low-perme- Formation, Central Powder River Basin, Wyoming:
ability rocks in the Muddy Sandstone, it created M.S. thesis, University of Wyoming, 115 p.
mostly stratigraphically confined sandstone reservoirs Desmond, R.J., J.R. Steidtmann, and D.F. Cardinal,
in the Minnelusa, only locally resulting in stratigraph- 1984, Stratigraphy and depositional environments
ically isolated reservoirs. Furthermore, whereas the of the Middle Member of the Minnelusa Formation,
majority of Muddy Sandstone hydrocarbon reservoirs Central Powder River Basin, Wyoming, in J.
comprise fluid-pressure compartments, only a Goolsby and D. Morton, eds., The Permian and
few compartments have been identified within the Pennsylvanian Geology of Wyoming: Wyoming
Minnelusa, much fewer than the number of strati- Geological Association 35th Annual Field Confer-
graphically isolated sandstones that probably exist. ence Guidebook: Casper, Wyoming Geological
However, the conditions under which fluid-pressure Association, p. 213–240.
compartments within each of these two formations Dolson, J.C., D.S. Muller, M.J. Evetts, and J.A. Stein,
occur are similar; these compartments occur only 1991, Regional paleotopographic trends and pro-
where stratigraphic, structural, or diagenetic variations duction, Muddy Sandstone (Lower Cretaceous),
have resulted in not overly large volumes of porous central and northern Rocky Mountains: AAPG
and permeable rock encased in low-permeability rock, Bulletin, v. 75, p. 409–435.
and where reservoirs are closely associated with iden- Downey, M.A., 1984, Evaluating seals for hydrocarbon
tified source rocks. The majority of reservoirs filled by accumulations: AAPG Bulletin, v. 68, p. 1752–1763.
long-distance migration of hydrocarbons have geome- Foster, D.I., 1958, Summary of the stratigraphy of
tries that are conducive to confining fluid flow, not the Minnelusa Formation, Powder River Basin,
isolating fluid systems, and generally comprise con- Wyoming, in J. Strickland, ed., Powder River Basin:
ventional traps; only rarely can a reservoir be charac- Casper, Wyoming Geological Association, p. 39–44.
terized as a fluid-pressure compartment. Maintenance Fryberger, S.G., 1984, The Permian Upper Minnelusa
of anomalous pressure over significant periods of geo- Formation, Wyoming: Ancient example of an off-
logic time is possible within fluid-pressure compart- shore-prograding eolian sand sea with geomorphic
ments, whereas conventional traps are more typically facies, and system-boundary traps for petroleum, in
characterized by normal pressure. J. Goolsby and D. Morton, eds., The Permian and
Pennsylvanian Geology of Wyoming: Wyoming
Geological Association Guidebook 35th Annual
ACKNOWLEDGMENTS Field Conference: Casper, Wyoming Geological
I would like to thank Robert Berg, John Dolson, Association, p. 241–272.
John Kaldi, W. Richard Moore, Paul Slack, James R. George, G.R., 1984, Cyclic sedimentation and deposi-
Steidtmann, and John Warme for their help in review- tional environments of the Upper Minnelusa For-
ing and revising this manuscript. Kathy Kirkaldie pro- mation, Central Campbell County, Wyoming, in J.
vided useful editorial help, and Allory Deiss drafted Goolsby and D. Morton, eds., The Permian and
the illustrations. This study was funded in part under Pennsylvanian Geology of Wyoming: Wyoming
GRI Contract 5089-260-1894. Geological Association Guidebook 35th Annual
Field Conference: Casper, Wyoming Geological
Association, p. 75–96.
REFERENCES CITED Gill, J.R., and W.A. Cobban, 1973, Stratigraphy and
geologic history of the Montana Group and equiva-
Asquith, D.O., 1970, Depositional topography and lent rocks, Montana, Wyoming, and North Dakota:
major marine environments, Late Cretaceous, U.S. Geological Survey Professional Paper 776, 37 p.
Wyoming: AAPG Bulletin, v. 54, p. 1184–1224. Gustason, E.R., 1988, Depositional and tectonic history
Berg, R.R., 1975, Capillary pressure in stratigraphic of the Lower Cretaceous Muddy Sandstone, Lazy B
traps: AAPG Bulletin, v. 59, p. 939–956. field, Powder River Basin, Wyoming, in R.
Bethke, C.M., W.J. Harrison, C. Upson, and S.P. Diedrich, M. Dyka, and W. Miller, eds., Eastern
Altaner, 1988, Supercomputer analysis of sedimen- Powder River Basin—Black Hills: Wyoming Geo-
tary basins: Science, v. 239, p. 261–267. logical Association 39th Field Conference Guide-
Clayton, J.L., and R.T. Ryder, 1984, Organic geo- book: Casper, Wyoming Geological Association,
chemistry of black shales and oils in the Minnelusa p. 129–146.
240 Martinsen

Gustason, E.R., T.A. Ryer, and S.K. Odland, 1988a, Wyoming, in Remote Sensing for Exploration Geol-
Stratigraphy and depositional environments of the ogy: Proceedings of the Fourth Thematic Confer-
Muddy Sandstone, northwestern Black Hills, ence, San Francisco, p. 9–19.
Wyoming: Wyoming Geological Association Earth Maughn, E.K., 1990, Summary of the ancestral Rocky
Science Bulletin, v. 20, p. 49–60. Mountains epeirogeny in Wyoming and adjacent
Gustason, E.R., D.A. Wheeler, and T.A. Ryer, 1988b, areas: U.S. Geological Society Open-File Report
Structural control on paleovalley development, 90-447, 8 p.
Muddy Sandstone, Powder River Basin, Wyoming Maucione, D., V. Serebryakov, P. Valasek, Y. Wang,
(abs.): AAPG Bulletin, v. 72, p. 871. and S. Smithson, 1995, A sonic log study of abnor-
Heasler, H.P., and R.C. Surdam, 1992, Pressure com- mally pressured zones in the Powder River Basin
partments in the Mesaverde Formation of the Green of Wyoming, in P. Ortoleva, ed., Basin compart-
River and Washakie basins, as determined from ments and seals: AAPG Memoir 61, p. 333–348.
drill-stem test data, in C. Mullen, ed., Rediscover McBane, J.D., 1984, Buck Creek field, Niobrara
the Rockies: Casper, Wyoming Geological Associa- County, Wyoming, in J. Goolsby and D. Morton,
tion Guidebook, p. 207–220. eds., The Permian and Pennsylvanian geology of
Heasler, H.P., R.C. Surdam, and J.H. George, 1995, Wyoming: Wyoming Geological Association
Pressure compartments in the Powder River Basin, Guidebook 35th Annual Field Conference: Casper,
Wyoming, as determined from drill-stem test data, Wyoming Geological Association, p. 111–118.
in P. Ortoleva, ed., Basin compartments and seals: Merewether, E.A., W.A. Cobban, and E.T. Cavanaugh,
AAPG Memoir 61, p. 235–262. 1979, Frontier Formation and equivalent rocks in
Hobson, J.P., M.L. Fowler, and E.A. Beaumont, 1982, eastern Wyoming: The Mountain Geologist, v. 16,
Depositional and statistical exploration models, p. 67–102.
Upper Cretaceous offshore sandstone complex, Momper, J.A., and J.A. Williams, 1979, Geochemical
Sussex Member, House Creek field, Wyoming: exploration in the Powder River basin: Oil and Gas
AAPG Bulletin, v. 66, p. 689–707. Journal, p. 129–134.
Iverson, W.P., R.S. Martinsen, and R.C. Surdam, 1995, Moore, W.R., 1983, The nature of the Minnelusa–
Pressure seal permeability and two-phase flow, in Opeche contact in the Halverson field area, Powder
P. Ortoleva, ed., Basin compartments and seals: River Basin, Wyoming: The Mountain Geologist,
AAPG Memoir 61, p. 313–319. v. 20, p. 113–120.
James, S.W., 1989, Diagenetic history and reservoir Powley, D.E., 1982, Pressures, normal and abnormal:
characteristics of a deep Minnelusa reservoir, Hawk AAPG Advanced Exploration Schools Unpublished
Point field, Powder River Basin, Wyoming, in E.B. Lecture Notes, 48 p.
Coalson, ed., Rocky Mountain Reservoirs—1989: Powley, D.E., 1990, Pressures and hydrogeology in
Denver, Rocky Mountain Association of Geologists, petroleum basins: Earth Science Reviews, v. 29,
p. 81–96. p. 215–226.
Jiao, Z.S., 1992, Thermal maturation/diagenetic aspects Schowalter, T.T., 1979, Mechanics of secondary hydro-
of the abnormal pressure in Cretaceous shales and carbon migration and entrapment: AAPG Bulletin,
sandstones, Powder River Basin, Wyoming: Ph.D. v. 63, p. 723–760.
dissertation, University of Wyoming, Laramie, 242 p. Shinn, E.A., 1983, Tidal flat environment, in P. Scholle,
MacGowan, D.B., Z.S. Jiao, R.C. Surdam, and F.P. D. Bebout, and C. Moore, eds., Carbonate deposi-
Miknis, 1993, Normally-pressured vs. abnormally- tional environments: AAPG Memoir 33, p. 172–210.
pressured compartments in sandstones in the Pow- Sheldon, R.P., 1967, Long-distance migration of oil in
der River Basin, Wyoming: a comparative study of Wyoming: Mountain Geologist, v. 4, p. 53–65.
the Muddy Sandstone and the Minnelusa Forma- Sheppy, R.J., 1986, Slattery field, Powder River Basin,
tion: Chicago, Gas Research Institute report. Wyoming: a multidisciplinary interpretation of a
Mallory, W.W., 1967, Pennsylvanian and associated complex Minnelusa (Permian) field: Wyoming Geo-
Rocks in Wyoming: U.S.G.S. Professional Paper, logical Association Symposium, Rocky Mountain
554–G, p. G1–G31. Oil and Gas Fields: Casper, Wyoming Geological
Martinsen, R.S., 1992, First quarter report, multidisci- Association, p. 245–256.
plinary analysis of pressure chambers in the Pow- Slack, P.B., 1981, Paleotectonics and hydrocarbon
der River Basin, Wyoming and Montana: Chicago, accumulation, Powder River Basin, Wyoming:
Gas Research Institute Contract no. 5089-260-1894, AAPG Bulletin, v. 65, p. 730–743.
unpublished report. Smith, D.A., 1966, Theoretical considerations of seal-
Martinsen, R.S., 1995, Stratigraphic compartmentation ing and non-sealing faults: AAPG Bulletin, v. 50,
of reservoir sandstones: examples from the Muddy p. 363–374.
Sandstone, Powder River Basin, Wyoming, in P. Stone, D.S., 1967, Theory of Paleozoic oil and gas accu-
Ortoleva, ed., Basin compartments and seals: AAPG mulation in Big Horn Basin, Wyoming: AAPG
Memoir 61, p. 273–296. Bulletin, v. 51, p. 2056–2114.
Martinsen, R.S., and R.W. Marrs, 1985, Comparison of Tillman, R.W., and R.S. Martinsen, 1984, The Shannon
major lineament trends to sedimentary rock thick- shelf-ridge sandstone complex, Salt Creek anticline,
nesses and facies distributions, Powder River Basin, Powder River Basin, Wyoming, in R. Tillman and
Stratigraphic Controls on the Development and Distribution of Fluid-Pressure Compartments 241

C. Siemers, eds., Siliciclastic shelf sedimentation: Trotter, J.F., 1984, The Minnelusa revisited, in
SEPM Special Publication 34, p. 85–142. J. Goolsby and D. Morton, eds., The Permian and
Tillman, R.W., and R.S. Martinsen, 1987, Sedimento- Pennsylvanian geology of Wyoming: Wyoming
logic model and production characteristics of Geological Association Guidebook 35th Annual
Hartzog Draw field, Wyoming, a Shannon shelf- Field Conference: Casper, Wyoming Geological
ridge sandstone, in R. Tillman and K. Weber, eds., Association, p. 127–152.
Reservoir sedimentology: SEPM Special Publica- Tyler, D.L., and A.C. Modroo, 1986, High Road field,
tion 40, p. 15–112. Campbell County, Wyoming: Wyoming Geological
Tromp, P.L., 1981, Stratigraphy and depositional envi- Association Symposium, Rocky Mountain Oil and
ronment of the “Leo Sands” of the Minnelusa For- Gas Fields, p. 233–243.
mation, Wyoming and South Dakota: M.S. thesis, Wyoming Geological Association, 1981, Oil and gas
University of Wyoming, Laramie, 69 p. fields symposium, Powder River Basin: Casper,
Tromp, P.L., D.F. Cardinal, and J.R. Steidtmann, 1981, Wyoming Geological Association, 472 p.
Stratigraphy and depositional environments of the Wyoming Geological Association, 1991, Mineral
“Leo Sands” in the Minnelusa Formation, Wyoming resources of Wyoming: Casper, Wyoming Geologi-
and South Dakota, in Energy Resources of cal Association, p. xii.
Wyoming: Casper, Wyoming Geological Associa- Wyoming Oil & Gas Commission, 1991: Casper,
tion, p. 11–21. Wyoming Oil and Gas, 52 p.
Jiao, Z.S., and R.C. Surdam, 1997, Characteristics of
anomalously pressured Cretaceous shales in the
Laramide Basins of Wyoming, in R.C. Surdam,
ed., Seals, traps, and the petroleum system:
Chapter 14 AAPG Memoir 67, p. 243–253.

Characteristics of Anomalously Pressured


Cretaceous Shales in the Laramide Basins
of Wyoming
Z.S. Jiao
R.C. Surdam
Institute for Energy Research, University of Wyoming
Laramie, Wyoming, U.S.A.

ABSTRACT
Most of the anomalously pressured Cretaceous shales in the Laramide
basins of Wyoming (LBW) are overpressured and form basinwide, dynamic
pressure compartments. The driving mechanism of the overpressuring is the
generation and storage of liquid hydrocarbons that subsequently react to
gas. This results in the conversion of the fluid-flow system from a single-
phase to a multiphase regime in which capillarity controls the relative
permeability, resulting in elevated displacement pressures within the shales.
The boundary between the normally pressured (i.e., pressure gradient
ranging from 0.433 to 0.444 psi/ft, single-phase regime) and anomalously
pressured (i.e., pressure gradient 0.433 psi/ft or 0.444 psi/ft, multiphase
regime) Cretaceous shales is marked by significant differences in the geo-
chemical and geophysical properties of the shales. The top boundary of the
anomalously pressured zone is characterized by a marked decrease in sonic
velocity and significant changes in the production index (PI), clay diagenesis
(illite/smectite), vitrinite reflectance (Ro), nuclear magnetic resonance spec-
tra (NMR), aromaticity, and displacement pressure of the Cretaceous shales.
In this paper, we document and attempt to characterize the fundamentally
different geophysical and geochemical properties of the anomalously pres-
sured Cretaceous shales below the pressure boundary in the LBW.

ANOMALOUSLY LOW hydrocarbon generation, migration, and accumulation,


and the development of reservoir characteristics and
SONIC VELOCITIES anomalous pressure. Surdam et al. (1995, 1997) have
demonstrated that Cretaceous shales below a depth of
Because of their widespread occurrence and unique ~8000 ± 2000 ft (2440 ± 610 m) are typically anomalously
chemical and physical nature, the Cretaceous shales in pressured in the central portions of these basins. In the
the Laramide basins of Wyoming (LBW) (Figure 1) have basin centers, the top of the anomalously pressured
played a definitive role in basin fluid/rock evolution, zone is transitional [1000–2000 ft (305–610 m) thick] and

243
244 Jiao and Surdam

Figure 1. Index map of the


WYOMING Laramide basins of
45
Wyoming (LBW).

Powder
Bighorn River
44 Basin
Basin

Wind River
Latitude

43 Basin

Green
42 River Washakie
Basin Basin

50 miles

41
111 110 109 108 107 106 105
Longitude

typically occurs within the Upper Cretaceous shales A marked decrease in sonic velocity, or increase in
(i.e., Steele, Cody, or Lewis), persisting down to the sonic transit time, has conventionally been attributed
lowermost organic-rich shale in the Cretaceous. Below to overpressure, but some of the shales in the LBW
the anomalously pressured zone in the Powder River exhibiting the reversal are actually underpressured. In
and Wind River basins, the shales are, for the most part, addition, shales with anomalously low sonic velocities
normally pressured. are usually characterized by undercompaction and
As indicated by the typical sonic logs from seven abnormally high porosity, but this has not been
basins shown in Figures 2–5, the Cretaceous shales observed in the Cretaceous shales in the LBW.
(i.e., the Skull Creek, Mowry, Frontier, Niobrara, and Although one could infer from the decompacted sonic
Steele) in the LBW display anomalously low sonic logs that all the Cretaceous shales in the LBW are
velocities. Three plots were constructed for each figure undercompacted (Figures 2–5), these shales are actu-
and, from left to right, represent the following: ally characterized by normal compaction, normal to
low porosity, and low permeability <8000 ±2000 ft
• Sonic velocity filtered for lithology. The filtering is (2440 ±610 m) (where basinwide anomalously pres-
accomplished by using gamma-ray logs, which are sured compartments have been investigated).
very useful in separating shaly lithologies from For instance, the measured porosity of the Creta-
sandy ones in the LBW, and results in only fine- ceous shales in the Powder River Basin is 11.7% at 5000
grained, clastic lithologies being represented. ft (1520 m) depth, and rapidly decreases to 2.2% at
• The sonic velocity profile after a decompaction 10,000 ft (3050 m) depth. The measured permeability is
correction is applied. The decompaction correc- 0.012 md at 5000 ft (1520 m) depth and 0.004 md at
tion assumes that sonic velocity increases expo- 10,000 ft (3050 m) depth (Table 1). The measured
nentially with burial and is constrained by the porosity of the Cretaceous shales in the Washakie
highest observed sonic velocity value in the sonic Basin is 5.5% at 5000 ft (1520 m) depth, and decreases
logs (i.e., typically ~5000 m/s). to 2.1% at 11,000 ft (3350 m) depth. The measured per-
• The anomalous sonic velocity profile. This profile meability is 0.013 md at 5000 ft (1520 m) depth and
is derived by taking the difference between the 0.008 md at 11,000 ft (3350 m) depth (Table 1). These
observed sonic velocity and the calculated sonic low-permeability shales provides potential seals for
velocity (trend representing normal compaction). hydrocarbon/pressure compartments in the Creta-
While the decompacted sonic velocity profile is ceous shale and sandstone sections in the LBW.
obtained by using curve-fitting techniques, this Because undercompaction and abnormally high
profile is obtained by difference. For more detail, porosity are not observed in the Cretaceous shales
see Surdam et al. (1997). with anomalously low velocities in the LBW, and
Characteristics of Anomalously Pressured Cretaceous Shales in the Laramide Basins of Wyoming 245

Sonic Velocity After Decompaction Correction Anomalous Velocity


Powder River Basin
0 0 0

500 500 500

1000 1000 1000

1500 1500 1500

2000 2000 2000


Depth (m)

Depth (m)

Depth (m)
2500 2500 2500

3000 3000 3000

3500 3500 3500

4000 4000 4000

4500 4500 4500

5000 5000 5000


7000 5000 3000 4500 3500 2500 1500 1000 500 0 -500 -1000
Velocity (m/s) Velocity (m/s) Velocity (m/s)

Figure 2. Typical sonic velocity vs. depth plots for the Powder River Basin.

because the shales can be either overpressured or compressional wave velocity than does any
underpressured, other factors must be affecting observed porosity change. Thus, variations in fluid
sonic velocity in these shales. Surdam et al. (1997), composition, particularly at gas saturations >15%
among others, suggest that fluid type, effective rock (Timur, 1987), are likely to be, at least in part, a cause
stress, lithology, matrix compaction, salinity, and of the observed sonic velocity reversals observed in
bed thickness can also affect sonic velocity. Surdam the LBW (Figures 2–5).
et al. (1997) also suggest that in the LBW, the gas These observations, exploration and drilling expe-
content of fine-grained rocks has a greater impact on rience in the Laramide basins, and pore pressure

Sonic Velocity After Decompaction Correction Anomalous Velocity


Bighorn Basin
0 0 0

500 500 500

1000 1000 1000

1500 1500 1500

2000 2000 2000


Depth (m)

Depth (m)

Depth (m)

2500 2500 2500

3000 3000 3000

3500 3500 3500

4000 4000 4000

4500 4500 4500

5000 5000 5000


7000 5000 3000 4500 3500 2500 1500 1000 500 0 -500 -1000
Velocity (m/s) Velocity (m/s) Velocity (m/s)

Figure 3. Typical sonic velocity vs. depth plots for the Bighorn Basin.
246 Jiao and Surdam

Sonic Velocity After Decompaction Correction Anomalous Velocity


Wind River Basin
0 0 0

500 500 500

1000 1000 1000

1500 1500 1500

2000 2000 2000


Depth (m)

Depth (m)

Depth (m)
2500 2500 2500

3000 3000 3000

3500 3500 3500

4000 4000 4000

4500 4500 4500

5000 5000 5000


7000 5000 3000 4500 3500 2500 1500 1000 500 0 -500 -1000
Velocity (m/s) Velocity (m/s) Velocity (m/s)

Figure 4. Typical sonic velocity vs. depth plots for the Wind River Basin.

measurements such as repeat formation tests (RFTs) and rocks. In the Powder River, Bighorn, Wind River, and
drill-stem tests (DSTs) (Heasler et al., 1995; Surdam et Washakie basins of Wyoming (Figures 2–5), the rever-
al., 1995) indicate that the observed sonic velocity (tran- sals correlate well with the boundary between normally
sit time) reversals mark the boundary between normally pressured and overpressured Cretaceous shales. The
pressured (i.e., pressure gradient ~0.433 psi/ft), water- sonic velocity trends and patterns are similar for all the
saturated rocks and underlying, anomalously pressured anomalously pressured Cretaceous shales in the LBW,
(i.e., pressure gradient > or <0.433 psi/ft), gas-saturated whether they are overpressured or underpressured.

Sonic Velocity After Decompaction Correction Anomalous Velocity


Washakie Basin
0 0 0

500 500 500

1000 1000 1000

1500 1500 1500

2000 2000 2000


Depth (m)

Depth (m)

Depth (m)

2500 2500 2500

3000 3000 3000

3500 3500 3500

4000 4000 4000

4500 4500 4500

5000 5000 5000


7000 5000 3000 4500 3500 2500 1500 1000 500 0 -500 -1000
Velocity (m/s) Velocity (m/s) Velocity (m/s)

Figure 5. Typical sonic velocity vs. depth plots for the Washakie Basin.
Characteristics of Anomalously Pressured Cretaceous Shales in the Laramide Basins of Wyoming 247

Table 1. Displacement Pressure (Pd) and Sealing Capacity (H) of the Cretaceous Shales, Powder River and
Washakie Basins.

Depth Porosity Permeability Pd (Gas/Water) Sealing Capacity (H)


(ft) φ (%) K (md) (psi) (ft) (gas column)
Powder River Basin
5500 11.7 0.012 400 600
7900 5.2 0.059 1300 2000
8000 5.6 0.013 1500 2300
8800 2.9 0.012 1800 2700
9900 2.7 n/a 2800 4300
10,000 2.2 0.004 3000 4600
13,000 2.2 0.004 4000 6100

Washakie Basin
5000 5.5 0.013 1600 2400
7300 4.6 0.009 1900 2900
10,000 1.3 0.007 2200 3300
10,500 5.2 0.035 2000 3000
11,000 2.1 0.008 2200 3300
14,000 3.0 0.009 2000 3000

It is interesting to note that the anomalously low As shown by a sample suite from the Powder River
velocities observed in these basins could have been Basin of Wyoming (Figure 6), the major diagenetic
recorded at the time of either maximum burial or over- events that occurred in the LBW were the transforma-
pressuring. It is possible that the pressure imprint (i.e., tion of smectite to illite and of kaolinite to chlorite, and
the anomalously low sonic velocity) at the time of the ordering of the mixed-layer I/S clays. Figure 6
maximum burial could have been maintained even if shows the x-ray diffraction patterns of the clay-size frac-
the rock volume was subsequently uplifted and tion from the Mowry Shale between ~3600 ft (1100 m)
eroded, and the rocks became underpressured as a and 14,000 ft (2440 m) present-day depth in the Pow-
result of these processes. der River Basin. Figure 7 shows the percentage of illite
in the mixed-layer illite/smectite clays from the Creta-
Clay Diagenesis ceous shales in the LBW. Generally, illitization of the
smectite becomes significant at ~8000 ft (2440 m) pres-
The diagenesis of mixed-layer illite/smectite clays
ent-day depth and is largely complete by 9500 ft (2900
during progressive burial is widely recognized as an
m) present-day depth (Figure 7). This interval coin-
important empirical diagenetic geothermometer
cides well with the transition between normally pres-
(Burst, 1969; Hower et al., 1976; Boles and Franks,
sured, water-saturated Cretaceous shales and
1979; Hower, 1981; Pytte and Reynolds, 1989). The dia-
anomalously pressured, hydrocarbon-saturated Creta-
genetic trend in illite/smectite (I/S) clay is known to ceous shales in the LBW. The high percentage of illite
be temperature dependent and may be related to present in the shallow section of the Washakie Basin
regional hydrocarbon generation (Bruce, 1984; Hagen was caused by extensive uplift and erosion (Figure 7).
and Surdam, 1984). Several authors (Hower, 1981;
Pytte and Reynolds, 1989) indicate that the main com- Organic Geochemistry
positional and structural changes in the I/S burial dia-
genetic sequence are: (1) an increase in illite layers; (2) Organic geochemical analyses performed on cut-
an increase in interlayer potassium; (3) an increase in tings from the Cretaceous shales in the LBW include
the amount of aluminum substituted for silicon in the vitrinite reflectance (R o ), anhydrous pyrolysis
tetrahedral layer; and (4) release of Mg2+, Fe2+, Ca2+, [reported as production index (PI); Tissot and
Si 4+, Na +, and water. The released water can make Welte, 1984], and solid-state 13C nuclear magnetic
≤35% of the volume of the smectite crystallite (Petty resonance (reported as NMR spectra and carbon
and Hower, 1972). aromaticity).
In the LBW, these compositional and structural Vitrinite reflectance is a measure of organic matter
changes had a significant impact on the porosity and thermal maturity. The liquid oil generation window is
permeability of the Cretaceous shales and on the generally thought to occur from an Ro of ~0.5%–0.7% to
development of basinwide pressure compartments. 1.0%–1.3% (Tissot and Welte, 1984). Oil is thermally
Released cations likely also affected the porosity and cracked to wet gas at an Ro of 1.3%–2.0%. In the LBW,
permeability of adjacent sandstones due to the precip- R o increases slowly from 0.47% at ~1000 ft (305 m)
itation of quartz, chlorite, kaolinite, and late carbonate depth to 0.6% at 9000 ft (2740 m) depth, where the
cement. anomalously pressured Cretaceous shale section is
248 Jiao and Surdam

Figure 6. X-ray
diffractograms of ethylene-
glycolate treated samples
(<0.5-µm fraction) of the
Mowry Shale, Powder River
Basin. With increasing
burial depth, smectite alters
to illite in the mixed-layer
illite/smectite clays as
indicated by the 5.2°2θ
peak shifting to 6.6°2θ.
The structure of the
mixed-layer illite/smectite
clays changes from random
to ordered at ~3000 m, as
indicated by the advent of
the 6.6°2θ peak. The content
of chlorite also increases
with burial.

present. Vitrinite reflectance increases rapidly below 14,000 ft (4270 m) depth (Figure 9). The PI values also
this depth, from 0.6% at 9000 ft (2740 m) to 1.7% at appear to indicate that shales <9000 ft (2740 m) pres-
15,000 ft (4570 m) (Figure 8). Thus, there is likely a ent-day depth in the LBW have an accelerated geot-
different rapid thermal maturation regime for Creta- hermal gradient and a different thermal maturation
ceous shales above, than for those below, 9000 ft regime than shales above this depth.
(2740 m) depth. Organic material from the Cretaceous shales in the
Production indices are calculated as S 1 /S 1 + S 2 , Laramide basins was also analyzed by solid-state 13C
where S1 = amount of hydrocarbon already present nuclear magnetic resonance (NMR) by using cross-
before pyrolysis, and S 2 is the amount of hydrocar- polarization with magic angle spinning (MAS) and
bon generated during pyrolysis. Like Ro, the PI of the high-power decoupling. Samples with low total
Cretaceous shales in the LBW increases slowly from organic carbon (TOC) values were analyzed without
0.03 at 1000 ft (305 m) depth to 0.1 at 9000 ft (2740 m) performing kerogen-isolation procedures, as these
depth (Figure 9). Below this depth, PI increases procedures are time-consuming and often alter the
rapidly from 0.1 at 9000 ft (2740 m) depth to 0.4 at composition and structure of the kerogen (Vandergrift
Characteristics of Anomalously Pressured Cretaceous Shales in the Laramide Basins of Wyoming 249

0 0

Powder River
2000 2000 Bighorn
Wind River
Washakie
4000 4000

6000 6000

Depth (ft)
Depth (ft)

8000 8000

10000 10000

12000 12000
Powder River
Bighorn
14000 14000
Wind River
Washakie
16000 16000
0 20 40 60 80 100 0.0 0.4 0.8 1.2 1.6 2.0
Illite in the I/S (%) R o (%)
Figure 7. Plot of percent illite in the mixed-layer
illite/smectite clays vs. present-day depth for the Figure 8. Plot of vitrinite reflectance (Ro) vs. present-
Cretaceous shales in the LBW. Color version of this day depth for the Cretaceous shales in the LBW.
figure appears on p. 213. Color version of this figure appears on p. 213.

et al., 1980). Instead, the samples were prewashed in the 13 C nuclei in the various structural and func-
an acid solution to remove oxides, hydroxides, and tional groups in kerogen and changes in kerogen
carbonates, all of which can contain paramagnetic structure associated with thermal maturation (Sur-
nuclei that can reduce the signal-to-noise ratio. dam and Crossey, 1985).
Additionally, use of a large-volume sample spinner Figure 10A–D shows the 13C NMR spectra of Creta-
facilitated the analysis of low TOC samples without ceous shales from various depths in the LBW. The 13C
performing kerogen isolation. NMR spectra have dominantly broad bands due to
For a typical NMR spectrum, the x-axis scale (in aliphatic and aromatic carbon structures. The first
parts per million) is the chemical shift, a measure of major upfield peak is the resonance signal from the
the magnetic field strength required to cause a para- carbon in aliphatic functional groups. The major
magnetic nucleus (here, 13C) to produce a resonance downfield peak is carbon in aromatic functional
signal. A paramagnetic nucleus that is shielded by groups; signals from carbon in the oxygenated func-
electrons produces a signal that is shifted upfield tional groups (phenol and carbonyl, respectively) are
(right) on the chemical-shift scale; a paramagnetic observed on the downfield side of this peak and are
nucleus that is deshielded by electrons produces a sig- present only in very small quantities in these shales,
nal that is shifted downfield (left). In typical kerogen even at very shallow depths (Figure 10A–D).
NMR spectra, the broad band between 0 and 60 ppm With progressive burial or thermal maturation, a
is associated with aliphatic carbons in branched and decrease occurs in the peak area of the aliphatic func-
straight chains and with naphthanic structures. The tional group and an increase occurs in the peak area
broad band between 100 and 200 ppm is associated and sharpness of the aromatic functional group. This
with aromatic and carbonyl carbon structures. Thus, trend reflects changes in the kerogen structure due to
NMR spectra can be used to study the abundance of thermal maturation. Carbon aromaticity (Figure 11)
250 Jiao and Surdam

PRODUCTION INDEX SEALING CAPACITY AND


Laramide Basin, WY CAPILLARY SEALS
0 Shales have highly ductile clay matrices, which
Powder River cause them to have some of the lowest permeability
Bighorn values (10 –22 to 10 –18 md) among the various rock
2000 types under normal compaction (Best and Katsube,
Wind River 1995). Because of their relative impermeability, shales
Washakie are the most common seals present for hydrocar-
bon/pressure compartments; in fact, North (1985)
4000 indicates that shales act as seals for sandstone reser-
voirs in more than 60% of the known giant oil fields.
Certain petrophysical and flow properties of a rock
determine the height of hydrocarbon column that can
6000 be supported by a hydrocarbon/pressure seal, or the
Depth (ft)

sealing capacity. Necessarily, then, the sealing capac-


ity is a function of pore radius and fluid characteris-
8000 tics. Smith’s (1966) equation is used in this study for
calculating the sealing capacity of Cretaceous shales:

10000
H=
( PdB – PdR )
0.433( ρ w – ρ h )
(1)

12000 where H = maximum hydrocarbon column in feet that


a seal can support, PdB = subsurface displacement pres-
sure is psi of the seal rock, PdR = subsurface displace-
14000 ment in psi of the reservoir rock, ρ w = subsurface
density in g/cm3 of water, ρh = subsurface density of
the hydrocarbon, and 0.433 = units conversion factor.
In our investigation, 30 bulk shale samples from the
16000 Cretaceous section in the LBW were subjected to high-
0.0 0.2 0.4 0.6 pressure mercury injection tests (Figure 12). The injec-
tion pressure curve (plateau portion) was extrapolated
PI, fraction to 0% mercury saturation (Figure 12) and, using
Schowalter’s (1979) nomograms, this pressure was
Figure 9. Plot of the production index (PI) vs. converted to produce the displacement pressure (Pd ),
present-day depth for the Cretaceous shales in the minimum pressure to start mercury moving
the LBW. Color version of this figure appears on through a sample, for the subsurface gas/water and
p. 212. oil/water systems. Results are shown in Table 1. For
the gas/water system, the sealing capacity of the Cre-
ranges from below this depth, increasing to only 83% taceous transgressive shales is 600 ft (183 m) (gas col-
at 14,000 ft (4270 m) depth. By 9000 ft (2740 m) depth, umn) at a present-day depth of 5500 ft (1670 m) and
the aliphatic carbon peak is almost gone, and kerogen rapidly increases to 2150 ft (655 m) at 8000 ft (2440 m)
has virtually exhausted its capacity to generate liquid depth and 6150 ft (1875 m) at 13,000 ft (3960 m) depth.
hydrocarbons and has little capacity to generate Displacement pressure also shows a significant
gaseous hydrocarbons. This geochemical evidence increase with burial. At a present-day depth of 13,000 ft
indicates that the Cretaceous shales in the LBW were (3960 m), transgressive shales can withstand a differ-
at one time exposed to much higher temperatures, ential pressure of 4000 psi.
which is consistent with the Ro and PI data. Sneider et al. (1991) proposed a classification of
The organic geochemistry values of these Creta- hydrocarbon seal type based on sealing capacity
ceous shales are characterized by fairly low gradients (Table 2). According to this classification, the sealing
with depths to ~8000 ±1000 ft (2440 ±610 m) present- capacity of the Cretaceous shales in the LBW increases
day depth, where steeping gradients occur. This is also exponentially with increasing depth of burial. Note
the depth at which there is a rapid steeping of the pres- that the increase in sealing capacity coincides with
sure/depth gradient in the Cretaceous sandstones in progressive clay diagenesis as the structure of mixed-
the LBW (Heasler et al., 1995; Surdam et al., 1995). layer I/S clay changes from random to ordered at
Thus, the changes in the geochemistry and the onset of ~8000 ft (2440 m) to 10,000 ft (3050 m) depth (Figure 6);
the anomalous pressure appear to be genetically in other words, the sealing capacity of the Cretaceous
linked. shales changes from B to A type (Table 1).
Characteristics of Anomalously Pressured Cretaceous Shales in the Laramide Basins of Wyoming 251

Figure 10. 13C nuclear


Powder Washakie magnetic resonance
(NMR) spectra of Mowry
River Basin Shale samples from
Basin depths of 3000 ft (910 m)
to 18,000 ft (5490 m) in
4900 ft the LBW. Note that the
3085 ft aliphatic carbon peak (the
7441 ft right-hand major peak)
4010 ft diminishes or disappears at
9159 ft <10,000 ft (93,050 m) depth.
5720 ft
10663 ft
8377 ft
13659 ft
9230 ft
18420 ft

9695 ft
300 200 100 0 -100
11425 ft ppm

Bighorn
300 200 100 0 -100
Basin
ppm
Wind River 3015 ft
Basin 4100 ft
4530 ft
10350 ft 10500 ft
12440 ft 11010 ft

14280 ft 14060 ft

300 200 100 0 -100 300 200 100 0 -100


ppm ppm

Calculations by Iverson et al. (1995) indicate that on dominantly under water drive (i.e., a single-phase
the basis of single-phase fluid flow, even with an fluid-flow system). Below this depth, the shales are
assumed permeability of only 1 nd, a pressure differ- characterized by a multiphase fluid-flow system
ential will dissipate in 1 m.y. for a large reservoir, or (oil/water or gas/oil/water), are dominantly under gas
0.01 m.y. for a small reservoir. However, results from drive, and are typically overpressured (i.e., a multi-
thermal maturation modeling indicate that overpres- phase fluid-flow system). With the addition of hydro-
sured compartments in the LBW have existed for at carbons to the system, these low-permeability
least 40 m.y. (Jiao, 1992; Surdam et al., 1995). Cretaceous shales become orders of magnitude more
Investigation of the fluid-flow regimes of the Creta- effective as overpressured media than when they were
ceous shales in the LBW has helped to account for the in a single-phase (water) fluid-flow system. In fact,
maintenance of this pressure differential through geo- only in a multiphase system could displacement pres-
logic time. For instance, the vertical trends of Ro, PI, sures measured in the laboratory be of the same order
clay diagenesis, and NMR spectra indicate that above of magnitude as those observed in the Cretaceous sys-
9000 ft (2740 m) depth, few hydrocarbons remain in the tem in the Powder River Basin. The relative permeabil-
shales, and that below this depth, significant quantities ity of the shales has more control over the system than
of hydrocarbons have been generated and retained by their absolute permeability; in other words, the rela-
the shales. Above ~9000 ft (2740 m) depth in the LBW, tive permeability is the effective permeability charac-
the Cretaceous shales are normally pressured and terizing the fluid-flow system in the Cretaceous shales
252 Jiao and Surdam

Carbon Aromaticity Laramide Basin, WY


Laramide Basin, WY
100000
0
Powder River 50000
Bighorn
2000 Wind River
Washakie 9000
4000
4000

Hg Pressure, psi
6000
Depth (ft)

800

8000 300
Skull Creek, 7908 ft
Skull Creek, 7909 ft
10000 70 Mowry, 5430 ft
Mowry, 7885 ft
Mowry, 9906 ft
12000 20 Mowry, 10008 ft
Mowry, 13388 ft
Fuson, 11419 ft
6
14000 Lewis, 5076 ft
Lewis, 10440 ft
Lewis, 14392 ft
16000 1
0.6 0.8 1.0 100.0 80.0 60.0 40.0 20.0
Aromaticity, fraction Hg Saturation, %
Figure 11. Plot of carbon aromaticity vs. present-day Figure 12. Plot of mercury saturation vs. injection
depth for the Cretaceous shales in the LBW. pressure for transgressive shales from the LBW.

in these basins. Thus, in a multiphase system, capillary pressure compartmentalization within the Cretaceous
pressure becomes very important in the formation of shales in the LBW. In addition, these changes are asso-
the pressure seals and compartment boundaries. ciated with a rise in the geothermal gradient within
such pressure compartments. It is likely that the mech-
anism of compartmentalization is capillarity resulting
CONCLUSIONS from the establishment of a multiphase fluid-flow sys-
Inorganic and organic geochemical evidence from tem. In a single-phase fluid-flow system (water), the
the Cretaceous shales in the Laramide basins of Cretaceous shales act as low-permeability fluid-flow
Wyoming (LBW) indicates that fundamental changes barriers, but in a multiphase fluid-flow system
occur in the rock/fluid system between the present- (oil/gas/water), the low-permeability fluid-flow barri-
day burial depths of 8000 ft (2440 m) and 10,000 ft ers convert to capillary seals, which can hold the
(3050 m). Changes are observed in (1) the organic observed pressure differential for a geologically signif-
geochemistry of the source rock related to thermal icant period of time in the LBW.
maturation and oil generation, expulsion, and con-
version to gas (i.e., Ro, PI, NMR spectra, and carbon ACKNOWLEDGMENTS
aromaticity); (2) the level of clay diagenesis (e.g., illi-
tization and chloritization); and (3) the formation This study was funded through the Gas Research
pressure (i.e., DST and RFT measurements). These Institute under Contract 5091-221-2146. The original
changes occur at the same depth interval as regional document was reviewed by Kathy Kirkaldie (Institute
Characteristics of Anomalously Pressured Cretaceous Shales in the Laramide Basins of Wyoming 253

Table 2. A Classification of Hydrocarbon Seal Type by Sealing Capacity.*

Type Sealing Capacity (H)


A H>300 m (>1000 ft)
B H>150 m but <300 m (>500 ft but <1000 ft)
C H>30 m but <150 m (>100 ft but <500 ft)
D H>15 m but <30 m (>50 ft but <100 ft)
E H<15 m (<50 ft)
F Waste zone rocks
*From Sneider et al., 1991.

for Energy Research). Allory Deiss (IER) generated or shales and sandstones, Powder River Basin,
improved on all figures in this document. Wyoming: Ph.D. dissertation, University of
Wyoming, 242 p.
North, F.K., 1985, Petroleum geology: Boston, Allen
REFERENCES CITED and Unwin, 607 p.
Best, M.E., and T.J. Katsube, 1995, Shale permeability Petty E.A., Jr., and J. Hower, 1972, Late stage dehydra-
and its significance in hydrocarbon exploration: The tion in deeply buried pelitic sediments: AAPG
Leading Edge, p. 165–170. Bulletin, v. 56, p. 2013–2021.
Boles, J.R., and S.G. Franks, 1979, Clay diagenesis in Pytte, A.M., and R.C. Reynolds, Jr., 1989, The thermal
Wilcox sandstones of southwest Texas, implications transformation of smectite to illite, in N. Naesar and
of smectite diagenesis on sandstone cementation: T. McCullock, eds., Thermal history of sedimentary
Journal of Sedimentary Petrology, v. 49, p. 55–70. basins: New York, Springer-Verlag, p. 133–140.
Bruce, C.H., 1984, Smectite dehydration—its relation Schowalter, T.T., 1979, Mechanics of secondary hydro-
to structural development and hydrocarbon accu- carbon migration and entrapment: AAPG Bulletin,
mulation in the northern Gulf of New Mexico Basin: v. 63, p. 723–776.
AAPG Bulletin, v. 68, pp. 673–683. Smith, D.A., 1966, Theoretical consideration of sealing
Burst, J.R., Jr., 1969, Diagenesis of Gulf Coast clay sed- and nonsealing faults: AAPG Bulletin, v. 50,
iments and its possible relationships to petroleum p. 363–374.
migration: AAPG Bulletin, v. 53, p. 487–502. Sneider, R.M., K. Stolper, and J.S. Sneider, 1991, Petro-
Hagen, E.S., and R.C. Surdam, 1984, Maturation his- physical properties of seals: AAPG Bulletin, v. 75,
tory and thermal evolution of Cretaceous source p. 673–674.
rocks of the Big Horn Basin, Wyoming and Mon- Surdam, R.C., and L. Crossey, 1985, Mechanisms of
tana, in J. Woodard, F. Meisser, and J. Clayton, organic-inorganic interactions in sandstone/shale
eds., Hydrocarbon source rocks of the Greater sequences; relationship of organic matter and min-
Rocky Mountain Region: Denver, Colorado, Rocky eral diagenesis: SEPM Short Course Notes 17,
Mountain Association of Petroleum Geologists, p. 177–232.
p. 321–338. Surdam, R.C., Z.S. Jiao, and J. Liu, 1995, Anomalous
Heasler, H.P., R.C. Surdam, and J.H. George, 1995, pressure regime in the Washakie Basin: Gas
Pressure compartments in the Powder River Basin, Research Institute, Topical Report GRI-95/0390,
Wyoming and Montana, as determined from drill- 24 p.
stem test data, in P. Ortoleva, ed., Basin compart- Surdam, R.C., Z.S. Jiao, and H.P. Heasler, 1997, Anom-
ments and seals: AAPG Memoir 61, p. 235–262. alously pressured gas compartments in Cretaceous
Hower, J., 1981, Shale diagenesis, in F.J. Longstaff, ed., rocks of the Laramide basins of Wyoming: a new
Short course in clay diagenesis and the resource class of hydrocarbon acumulation, in R.C. Surdam,
geologists: Toronto Mineralogical Association of ed., Seals, traps, and the petroleum system: AAPG
Canada, p. 60–80. Memoir 67, p. 199–222.
Hower, J., E.F. Eslinger, M.E. Hower, and E.A. Perry, Timur, A., 1987, Acoustic logging, in H. Bradley,
1976, Mechanism of burial and metamorphism of ed., Petroleum engineering handbook: Dallas,
argillaceous sediments, 1, Mineralogical and chemi- Texas, Dallas Society of Petroleum Engineers,
cal evidence: Geological Society of America Bul- p. 51-1–51-52.
letin, v. 87, p. 725–737. Tissot, B.P., and D.H. Welte, 1984, Petroleum forma-
Iverson, W.P., R.S. Martinsen, and R.C. Surdam, 1995, tion and occurrence; 2nd edition: Berlin, Springer-
Pressure seal permeability and two-phase flow, in Verlag, 699 p.
P. Ortoleva, ed., Basin compartments and seals: Vandergrift, G.F., R.E. Winans, and E.P. Horowitz,
AAPG Memoir 61, p. 313–319. 1980, Quantitative study of the carboxylic acids in
Jiao, Z.S., 1992, Thermal maturation/diagenetic the Green River oil shale bitumen: Fuel, v. 59,
aspects of the abnormal pressure in Cretaceous p. 627–633.
Engelder, T., and J.T. Leftwich, Jr., 1997, A pore-pressure
limit in overpressured South Texas oil and gas
fields, in R.C. Surdam, ed., Seals, traps, and the
petroleum system: AAPG Memoir 67, p. 255–267.

Chapter 15

A Pore-Pressure Limit in Overpressured


South Texas Oil and Gas Fields
Terry Engelder
The Pennsylvania State University
University Park, Pennsylvania, U.S.A.

John T. Leftwich, Jr.


Old Dominion University,
Norfolk, Virginia, U.S.A.

ABSTRACT
One way to simplify the characterization of pore pressure, Pp, in deep,
overpressured basins is to divide oil and gas fields into stratified zones,
based on average pressure-depth trends that are approximately linear. With
this approximation, each zone is assigned a constant pressure-depth gradi-
ent. In the shallow portion of South Texas oil and gas fields (i.e., zone ONE),
Pp has a hydrostatic gradient, whereas in the uppermost overpressured por-
tions of these fields (i.e., zones TWO and THREE), Pp is characterized by gra-
dients that exceed the lithostatic trend of 1 psi/ft (22.6 MPa/km) (Leftwich
and Engelder, 1995). At greater depth (i.e., zone FOUR), Pp increases along a
gradient of about 0.9 psi/ft (20.3 MPa/km). The transition between zones
THREE and FOUR defines the depth at which Pp reaches a limit that is
85%–90% of the lithostatic (i.e., vertical) stress. Because a Pp gradient of ~0.9
psi/ft (20.3 MPa/km) is maintained throughout pressure zone FOUR in sev-
eral South Texas fields, the Pp limit is a regional phenomenon. Two condi-
tions leading to a Pp limit involve a cyclic leakage of pore fluid through zone
FOUR. In both cases, leakage is governed by a balance between Pp and the
minimum horizontal total stress, Sh. One condition favors leakage of pore
fluid through zone FOUR upon the opening of existing joints or the propa-
gation of new joints by natural hydraulic fracturing. The other condition
favors leakage along faults following refracturing during slip events. The
difference between these conditions is that leakage through joints can regu-
late Pp at a constant value through repeated cycles, whereas leakage by fault
slip leads to an ever-increasing Pp as Sh increases through repeated cycles.

255
256 Engelder and Leftwich

INTRODUCTION Frio-Vicksburg, the Rio Grande Embayment was


filled by the Norias delta system, which deposited
Starting with Dickinson’s (1953) compilation of thick sections of sand throughout the area.
pore pressure, Pp , in the Louisiana Gulf Coast region, it
became apparent that Pp in sedimentary basins rarely, Pressure vs. Depth Profiles
if ever, exceeds total vertical stress, Sv, arising from the
weight of overburden. Vertical stress is also called the The Pp data in this paper were taken from fields in
lithostatic stress; when Pp = Sv, the pore pressure is at the Frio-Vicksburg trend. Leftwich (1993) compiled
lithostatic pressure. An interesting detail of Dickin- pore pressure–depth (Pp–z) data largely from original
son’s (1953) compilation is that Pp is generally less than shut-in bottom-hole pressure (BHP) measurements
lithostatic pressure even at the maximum depth of recorded in numerous wells within each field. These
sampling. Because mechanisms for generating abnor- data were recorded by the various operators during
mal Pp are capable of generating pressures equal to the completion and production phases of each well,
lithostatic pressure (Sahay and Fertl, 1989), the geolo- and are available from Petroleum Information Service
gist is left to presume that some mechanism in the of Houston, Texas. In some fields, repeat formation
Louisiana Gulf Coast region allows the draining of tester (RFT) and multitester (WLT) pressure data sup-
pore fluid before pressures reach lithostatic levels. plemented the BHP data.
Leakage will occur if the pressure of the hydrocarbon The Pp–z data were used to define Pp zones within
exceeds the capillary entry pressure (i.e., the sealing South Texas oil and gas fields. In general, high-quality
capacity) of either a cap rock or a fault zone seal (Berg, Pp–z data from each well in a given field follow the
1975; Schowalter, 1979; Vavra et al., 1992). A second same trends because the fields have relatively small
explanation for draining pore fluid at Pp < Sv is that [≈500 psi (3 MPa)] differences in original pressures
high Pp causes the propagation of natural hydraulic due to structural relief or faulting across a field. Some
fractures (Hubbert and Willis, 1957; Secor, 1965; larger interfield faults in South Texas act as seals sepa-
Engelder and Lacazette, 1990). Hydraulic fracturing rating larger Pp differences [>500 psi (3 MPa)]. This is
would add to the porosity and bulk permeability of in contrast to fields elsewhere in the world where
the rock and, hence, allow Pp to relax by draining pore much larger pressure differences across faults are
fluid into the new fracture porosity. A third possibility detected within fields (Weedman et al., 1992; Grauls
is that fault refracturing during slip under the influ- and Baleix, 1994; Hart et al., 1995). Leftwich (1993)
ence of low effective normal stress allows pore fluid to used only original preproduction Pp data and was
drain at Pp < S v. Refracturing opens a local zone of careful to exclude low Pp data taken after a particular
high permeability along the fault, allowing pore fluid zone had experienced some production drawdown.
to leak up the fault (Sibson, 1992). In the absence of large structural relief, we inter-
In this paper, we use Pp data compiled by Leftwich preted Pp data that fell below a given Pp–z trend by
and Engelder (1995) from a number of South Texas oil more than 500 psi (3 MPa) as indicative of production
and gas fields to further refine our understanding of drawdown. Pressure data that were determined to be
the upper bound for Pp found in the interconnected defective or drawn down were left off the final plots,
porosity of overpressured rocks in sedimentary but the number of excluded data is indicated in cap-
basins. We shall accomplish this by developing a tions (e.g., Figure 2). Otherwise, 500 psi (3 MPa) dif-
sense for the variation of Pp with depth within several ferences across fields are due to the structural relief
South Texas fields. We then discuss mechanisms that common in many South Texas fields. Detailed pres-
may operate to limit Pp at levels less than lithostatic sure records are not available for any of the BHP
pressure. data; therefore, evaluation of the quality for each test
is not possible.
GEOLOGY AND PORE PRESSURE IN
Pore Pressure Zones Within South Texas Fields
SOUTH TEXAS OIL AND GAS FIELDS
Leftwich and Engelder (1995) argue that oil and gas
Our analysis draws on pore pressure data from fields of the northern Gulf of Mexico may be divided
seven fields located in the Tertiary oil- and gas-pro- into stratified layers or zones. In analyzing Pp–z data
ducing trends of South Texas, geologically a part of from each field, Leftwich and Engelder (1995) identi-
the Rio Grande Embayment (Figure 1). During the fied clusters of data to which lines could be fit. There-
Eocene and Oligocene, onshore South Texas was fore, the Pp–z data were divided into three or four
characterized by rapid deltaic sedimentation and a clusters characterized by a constant Pp gradient line.
complex system of coastward-dipping syndeposi- These linear trends are established by interpolating
tional growth faults that become progressively pressure data between sand bodies and through inter-
younger to the east and toward the Gulf (Galloway layered shale. At the scale of individual sand bodies
et al., 1982). The thickest sequences of the and interlayered shale beds, the pressure-depth profile
Eocene–Oligocene deltaic sedimentation are progra- is probably not linear, but it is the general trend that is
dational wedges of sand deposited during the Late important for our analysis. The linear trend of Pp data
Paleocene–Early Eocene (Wilcox) and the mid- to late from individual zones in a single well can have a cor-
Oligocene (Frio-Vicksburg). During deposition of the relation coefficient as high as 0.99 (Leftwich and
A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields 257

Ann Mag-type field Figure 1. Map of Gulf


Coast oil and gas fields
Alazan-type field for which pressure zone
FOUR gradients have been
compiled (Leftwich and
Engelder, 1995). Ann
Texas Salt Dome Mag-type and Alazan-type
Dominated fields are distinguished in
Structures this map.

d
d ren
en eT
Tr nd en
nd
e
rg Tr oc
Mi
Tre

bu io
s Fr4
ck
x
lco

Vi
Wi

1. ANN MAG
1 6 2. EL PAISTLE
5 2 3. MONTE CHRISTO
7 4. REDFISH BAY
Roll-over
Anticline 5. RITA
100 km 3
Structures 6. SARITA-SARITA EAST
7. S.E. RITA

Engelder, 1995). Pressure zones are reported for other shallow portion of Gulf Coast sediments, has a normal
basins of the world where Pp data extrapolate in a lin- hydrostatic gradient of approximately 0.45 psi/ft (≈10
ear fashion between sand bodies (Dainelli and Vig- MPa/km) (Figure 2). Pressure zone TWO has a gradi-
nolo, 1993). However, these zones are not related to ent on the order of 1.00 psi/ft (≈22.6 MPa/km) and is
pressure compartments as proposed by Powley (1990). located between the top of abnormal pressure and a
Pressure zone boundaries in the oil and gas fields of deeper top to undercompaction. The presence of zone
South Texas are further based on the depth to under- TWO is characteristic of the Ann Mag-type field. Zone
compaction, a poorly understood phenomenon that is THREE, located in abnormally pressured rocks below
a characteristic of shales commonly associated with the top of undercompaction, shows a gradient in
overpressure basins (Leftwich and Engelder, 1995). excess of 2.00 psi/ft (>45.2 MPa/km). Without excep-
The depth to the top boundary of the undercompacted tion, the Pp gradient is much steeper below the top of
zone is identified using a combination of conductivity, undercompaction, even though abnormal Pp may
sonic, and density logs (Hottman and Johnson, 1965). occur above the top of undercompaction.
The top boundary of undercompaction is defined as If the top of undercompaction and the top of abnor-
an abrupt increase in electrical conductivity, an mal pressure correspond, then the field is missing
increase in sonic travel time, and a drop in density, all pressure zone TWO, and formation pressure increases
of which are associated with a retention of pore water precipitously at the top of abnormal pressure. This
in shales (Fertl, 1976; Magara, 1978). occurs in Alazan-type fields like the Sarita-Sarita East
Using the distribution of zones, Leftwich and and Red Fish Bay fields (Figures 3, 4).
Engelder (1995) argue that there are two types of fields
in the Tertiary portion of South Texas. These are dis-
tinguished based on whether the top of undercom- PORE-PRESSURE LIMITS IN
paction and the top of abnormal Pp correspond. PRESSURE ZONE FOUR
Tertiary fields have pressure-depth profiles that are
characterized by Pp gradients showing either three We suggest that the transition dividing pressure
(i.e., the Alazan-type field where the depth to the top zones THREE and FOUR (point λ in Figures 2–5) rep-
of undercompaction is the same as the depth to the resents a natural limit to pore pressure in South Texas
top of abnormal pressure) or four (i.e., the Ann Mag- fields. Pore pressure at the transition is normalized by
type field where the depth to the top of undercom- dividing the magnitude of Pp by Sv . The ratio of fluid
paction is deeper than the depth to the top of abnormal pressure to overburden stress is known as λ in the geo-
pressure) linear segments. These linear segments con- mechanics literature (Hubbert and Rubey, 1959; Suppe
stitute the basis on which pressure zones are identified and Wittke, 1977). When λ refers to the transition point
on South Texas fields. Pressure zone ONE, found in the between pressure zones THREE and FOUR, we call λ
258 Engelder and Leftwich

PRESSURE (PSI) Figure 2. Geopressure profile


for the Monte Christo field,
Hidalgo County, Texas.
0 2000 4000 6000 8000 10000 12000 14000 Pressure zones ONE through
FOUR are located on the
0 geopressure profile. The
Monte Christo field is a poor
example of the correlation
MONTE CHRISTO between the top of undercom-
2000 paction and the boundary
between pressure zones TWO
and THREE. Leftwich and
Engelder (1995) show data for
several fields where this cor-
4000 relation is much better (100
BHP measurements from 87
wells with 11 data points are
0.46 psi/ft
excluded).
Pressure
6000 Zone
ONE
DEPTH (FT)

8000
---------Top of Abnormal Pressure 8300'
Pressure 1.23 psi/ft
Zone
TWO
10000 -----Top of Undercompaction 9850'
Pressure
Zone 3.71 psi/ft λ
THREE

12000 Sv
Pressure
Zone 0.71 psi/ft
FOUR
14000

Pore Pressure Limit λ = 0.91

16000

the geostatic ratio. In many cases an exact figure for FOUR has a gradient of 0.70 psi/ft (15.8 MPa/km) in
lithostatic stress is unknown, so we assume a litho- the Monte Christo field. Sarita-Sarita East is an
static gradient of 1 psi/ft (22.6 MPa/km). Making this Alazan-type field with three pressure zones: pressure
assumption means that λ specifying the Pp limit in zone FOUR is well constrained and has a pressure gra-
Figures 2–6 is slightly low (≈1%–3%), largely because dient of 0.91 psi/ft (20.6 MPa/km) (Figure 3). In both
the low density of sediments in the upper 3 km the Monte Christo and Sarita fields, Pp increases
of basins generates a lithostatic gradient of slightly rapidly with depth through pressure zone THREE
<1 psi/ft. until reaching an abrupt transition at the boundary
Interpretation of the point λ as the Pp limit relies above zone FOUR.
heavily on data from a deeper zone (FOUR), where the We suggest that a sharp transition in Pp is present
pressure segment is poorly constrained for many at the point λ even when pressure zone FOUR is
South Texas fields because there are so few Pp data poorly defined, as is the case for the Red Fish Bay field
from that portion of the fields. Monte Christo is an (Figure 4). Here, pressure zone FOUR is defined by
Ann Mag-type field with four zones (Figure 2). Zone three data points following a gradient of 1.13 psi/ft
FOUR is constrained by a cluster of data at about (25.6 MPa/km), a gradient that will eventually inter-
11,000 ft (3.4 km) and then three data points down to sect the Sv. Several other fields also have a sharp transi-
16,000 ft (4.9 km). These Pp data suggest that zone tion between pressure zones THREE and FOUR in the
A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields 259

PRESSURE (PSI) Figure 3. Geopressure profile


for the Sarita-Sarita East
0 2000 4000 6000 8000 10000 12000 14000 Field, Kenedy County, Texas.
This is an Alazan-type field
in which the top of under-
0 compaction and the top of
abnormal pressure are found
SARITA - SARITA EAST at the same depth.

2000

4000
0.43 psi/ft

6000
Pressure
Zone
DEPTH (FT)

ONE
8000

-------Top of Abnormal Pressure 9200'


----------Top of Undercompaction 9350'
10000

Pressure
2.56 psi/ft
Zone λ
THREE
12000 Sv

Pressure
Zone 0.91 psi/ft
14000 FOUR

Pore Pressure Limit λ = 0.90

16000

South Texas region. In some cases, the transition is with Ann Mag-type fields, we predict that zone TWO
defined indirectly. For example, in our compilation extends to the top of undercompaction. Having
of Pp data from the El Paistle field of Kenedy County, drawn pressure zone TWO on Figure 5, we predict a
Texas, we were able to obtain numerous high-quality minimum gradient for zone THREE of 2.90 psi/ft
pressure data down to about 9200 ft (2.8 km) and nine (65.5 MPa/km) and, at the transition between zones
data points below 11,800 ft (3.6 km) (Figure 5). The THREE and FOUR, λ = 0.79.
data below 11,800 ft (3.6 km) follow a Pp gradient typ- In summary, we were able to identify a Pp limit at
ical of pressure zone FOUR. We have only one data the point λ in seven fields of South Texas (Figure 6).
point for a 2600-ft (0.8-km) interval through a critical The geostatic ratio varies between 0.79 in the El Pais-
portion of the geopressure profile (pressure zones tle field to 0.91 in the Monte Christo field, with an
TWO and THREE). The top of abnormal Pp is indi- average of 0.86. The gradients for pressure zone
cated by three data points near a depth of 8700 ft FOUR show a variation from 0.71 psi/ft (16.0
(2.6 km). These three data points below the top of MPa/km) in the Monte Christo field to 1.13 psi/ft
abnormal pressure indicate that pressure zone TWO (25.6 MPa/km) in the Red Fish Bay field. On average,
has a gradient of 1.18 psi/ft (26.7 MPa/km). Electric the Pp gradient in pressure zone FOUR, the deepest
logs indicate that the top of undercompaction is at a zone in South Texas, is 0.86 psi/ft (19.9 MPa/km) for
depth of 10,900 ft (3.3 km). Based on our experience Tertiary shales in the Gulf of Mexico.
260 Engelder and Leftwich

PRESSURE (PSI) Figure 4. Geopressure profile


for the Red Fish Bay field,
0 2000 4000 6000 8000 10000 12000 Nueces County, Texas. This is
an Alazan-type field with the
0 top of undercompaction above
the top of abnormal pressure.
RED FISH BAY
2000

4000
0.46 psi/ft

6000
DEPTH (FT)

8000 ----Top of Undercompaction 8183'

-------Top of Abnormal Pressure 9100'

10000
3.06 psi/ft λ

12000 Sv
Pressure 1.13 psi/ft
Zone
FOUR
14000

Pore Pressure Limit λ = 0.87

16000

DISCUSSION A composite plot of all Pp data from pressure zone


FOUR of South Texas shows the same trend as the 0.86
In the South Texas fields, Pp increases rapidly below psi/ft (19.9 MPa/km) average calculated above (Fig-
the top of abnormal pressure to reach the Pp limit within ure 7). Our data set is consistent with Dickinson’s
~3000 ft (1 km). The Pp gradient of pressure zone THREE (1953) compilation of Louisiana Gulf Coast fluid-pres-
immediately above the transition to zone FOUR is at sure data. Both compilations leave little doubt that
least 2 psi/ft (45.2 MPa/km), and in some fields it may fluid pressure does not exceed the lithostatic stress, Sv,
exceed 4 psi/ft (90.4 MPa/km). The transition to zone in that area of the Gulf Coast. In fact, Dickinson’s data
FOUR is abrupt in all cases and occurs at a geostatic ratio rarely fall above the 0.95 psi/ft (20.3 MPa/km) pres-
between 0.79 and 0.91, but the depth of the point λ varies sure gradient line. The combination of Dickinson’s
among the South Texas fields between 10,500 ft (3.2 km) (1953) data with our South Texas data suggests that
and 12,300 ft (3.8 km) (Figure 6). Because geostatic ratio the limit for fluid pressures in both the Texas and
is similar from field to field in South Texas, we presume Louisiana Gulf Coast areas falls near the 0.9 psi/ft
that this similarity is significant and has a physical pressure gradient line.
explanation that is common among the fields. Our inter- In searching for a mechanism to explain a natural
pretation is that the average geostatic ratio (λ) of ~0.9 is a leakoff pressure from South Texas fields, other obser-
measure of the natural leakoff pressure from pressure vations are important. Below the point λ, the Pp gradi-
zone FOUR, deep within fields of South Texas. ent in zone FOUR in all seven South Texas gas fields
A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields 261

PRESSURE (PSI) Figure 5. Predicted geopressure


profiles from the El Paistle
0 2000 4000 6000 8000 10000 12000 14000 field, Kenedy County, Texas.
The predicted shape of the
0 geopressure profile is based
on the large vertical separation
between the top of abnormal
EL PAISTLE pressure and the top of
undercompaction. Such
2000 separation suggests that the
El Paistle field is an Ann
Mag-type field.

4000
0.43 psi/ft

6000
DEPTH (FT)

8000
Top of Abnormal Pressure 8700'
Predicted
Pressure 1.18 psi/ft
10000
Zone
TWO ----Top of Undercompaction 10900'

12000 Predicted
Pressure 2.90 psi/ft λ Sv
Zone
THREE
0.80 psi/ft
14000

Pore Pressure Limit λ = 0.79


16000

comes close to 0.9 psi/ft when the gradient is defined Leakage Through Pore Throats
by more than five data points. This means that what-
ever pathway is available for leakage, it does not In the fields of South Texas, pressure zone FOUR
maintain the hydraulic continuity necessary to estab- is found within a thick shale section of the Vicksburg
lish a hydrostatic gradient (≈0.46 psi/ft) below the Formation. If leakage along the same gradient
zone THREE–FOUR transition. Consequently, the (0.9 psi/ft) over several thousand feet in several
rocks in pressure zone FOUR do not have the fields is controlled by capillary pressure, then each
hydraulic conductivity of a porous sandstone or other of the fields must consist of rocks with roughly the
attributes of a pressure compartment (Powley, 1990). same pore-throat architecture. This seems to be the
Rather, it appears that a dynamic system is at work case in South Texas because zone FOUR falls within
where fluid pressure is maintained at the level of the the Vicksburg Formation. However, the thick shale
Pp limit. Leakage is slow and probably governed by the section of the Vicksburg often begins at the top of
dynamic balance between Pp and some mechanism to pressure zone THREE (Leftwich and Engelder, 1995).
relieve the pressure. Mechanisms that might allow Furthermore, the transition between zones THREE
high pressure along a gradient of ~0.9 psi/ft include and FOUR is not marked by a noticeable change in
capillary leakage through pore throats, selective leak- lithology or any other characteristic that would sug-
age along through-cutting faults, and leakage through gest a unique zone FOUR pore-throat architecture.
a network of joints. Because a steeper pressure gradient would favor
262 Engelder and Leftwich

λ = P /S
p v
Figure 6. Plot of pore-
pressure limit and the
0.5 0.6 0.7 0.8 0.9 1.0
gradient of pore pressure
10000
in pressure zone FOUR for
seven oil and gas fields in
P =S South Texas. The vertical
p h
10500 ANN MAG
line is a prediction for Sh in
South Texas, assuming that
leakage by joint opening
limit (ft)

MONTE CHRISTO
RED FISH BAY
takes place.
11000
RITA
p
depth to P

11500
SE RITA

12000 SARITA

EL PAISTLE

12500
0.5 0.6 0.7 0.8 0.9 1 1.1 1.2
pore pressure gradient (psi/ft)

λ AT PORE PRESSURE LIMIT


GRADIENT OF ZONE FOUR

leakage against capillary pressure, it is zone THREE ratio of effective stress, k of Matthews and Kelly
that may drain by leakage through pore throats. If a (1967), is controlled by friction
pressure gradient of 2.0–4.0 psi/ft is required for
leakage against capillary pressure, leakage through  1 
–2

( )
Sh – Pp
pore throats seems unlikely as the mechanism lead- k= =  µ2 + 1 2 + µ (1)
ing to the pore-pressure limit and the 0.9 psi/ft gra- Sv – Pp  
dient for pressure zone FOUR.
where µ = the coefficient of static friction. At hydro-
Selective Leakage Along Faults static pore pressure with µ = 0.85, Sh = 0.59Sv, consis-
Shale sections of the Vicksburg have occasional tent with leakoff data from normally pressured
sand layers that provide higher permeability chan- portions of the Gulf of Mexico (Zoback and Healy,
nels to faults. These are the layers from which most 1984). In sedimentary basins, rock friction acts as a
of the zone FOUR Pp data come. The 0.9 psi/ft Pp gra- governor, limiting differential stress, σd = Sv–Sh and
dient of zone FOUR indicates that these through-cut- thereby controlling S h. However, Pp has an equally
important role in setting the level of Sh. The idea is
ting faults do not maintain open communication
that in actively subsiding basins dominated by nor-
from one sand layer to the next. If the faults leak
mal faulting, the vertical stress, Sv, increases by sedi-
episodically, they act as a valve that opens and closes
mentary loading until σ d becomes large enough to
(Sibson, 1992). A fault may open to fluid flow when
initiate frictional slip on listric normal faults. Slip will
it refractures during slip, but then heal to keep the
act to laterally compress sediments of the basin,
overpressured section from draining completely. thereby increasing Sh and decreasing σd. The value of
The fault-valve process offers some hope of identify- σd at which frictional slip occurs is controlled by the
ing a mechanism for regulating Pp at some value less frictional strength of the faults (i.e., µ) as expressed
than lithostatic pressure. Regulation of Pp by leakage by the equation
of pore fluid upward along faults is governed by at
least two important considerations: Pp at the time of
fault slip and the interaction between Pp and Sh. (
σ d = 2 µ σ n – Pp ) (2)
Zoback and Healy (1984) propose that the shallow
Earth is in a constant state of brittle failure along where σn = the total normal stress across the fault zone
favorably oriented planes. In such a scenario, the and σn–Pp = the effective normal stress across the fault
A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields 263

Pressure or Stress (psi) Figure 7. A composite plot


of all pressure zone FOUR
Pp data from the seven
0 2000 4000 6000 8000 10000 12000 14000 16000 fields of South Texas as
0 listed in Figure 6.

2000 SOUTH TEXAS

4000

6000
Depth (ft)

8000

10000

12000 Pressure Sv
Zone
FOUR
14000

16000
P p = 0.8 S v
P p = 0.9 S v

zone. Equation 2 is a simplification assuming that drop during faulting (Scholz, 1990). For normal faulting,
faults have no cohesive strength. An increase in Pp this shear-stress drop takes place with an increase in Sh ,
causes a decrease in effective normal stress, and thus as indicated by a modest increase in post-slip stress to
slip occurs at a smaller σd. We can rewrite equation 2 Sh = 0.94Sv after slip during cycle a. Following fault slip,
in terms of total horizontal and vertical stress: the enhanced fault-zone permeability allows the fault to
partially drain sandstone and maybe interlayered shales
–2 of pressure zone FOUR. The effect of poroelastic con-
 1 

(
Sh =  µ 2 + 1 ) 2 + µ

(Sv – Pp ) + Pp (3)
traction (as discussed below) is shown as a simultaneous
drop in S h (Figure 8A). Unless the fault eventually
reseals, the pressure in the interconnected sandstone
layers will reequilibrate at a hydrostatic gradient.
Equation 3 predicts that S h becomes relatively The present abnormally high Pp of zone FOUR is
higher in overpressured parts of sedimentary basins, consistent with faults resealing and the sandstones
as is seen in leakoff pressure data from several basins recharging, as indicated by an increased Pp during the
of the world (Engelder and Fischer, 1994). latter part of cycle a. Immediately upon resealing in
Our premise is that leakage along faults is periodic post-slip time, the state of stress and Pp along the nor-
and takes place mainly after faults slip. The situation mal fault do not favor further slip. After the fault
during and after faulting is illustrated on a Pp-Sh-time reseals, either stress or Pp or both must recover for
plot for a typical South Texas field where, prior to slip at another slip cycle: Sv could increase with further basin
the beginning of cycle a, Pp = 0.90S v below the zone filling, Sh could decrease with downbasin slumping, or
THREE–FOUR transition (Figure 8A). At the moment of Pp can be recharged through such mechanisms as the
fault slip, Sh is higher than Pp according to equation 3. continued generation of hydrocarbons. Because three
For the example illustrated in Figure 8, µ = 0.85 and pre- variables operate independently, it would seem fortu-
slip Sh = 0.92Sv. Studies of fault mechanics suggest that itous that each should return to its previous values for
fault slip is arrested as a consequence of the shear-stress each repeated slip event.
264 Engelder and Leftwich

If sedimentation rate and tectonic deformation are flaws) are found by substituting KIc for KI and rear-
slow relative to the recharging of Pp, they would not ranging equation 5 to derive an equation that gives the
effect a change in Sh and Sv between slip cycles. In this tensile stress necessary to initiate growth of a flaw of
case, Pp would be the only independent variable con- any size and idealized shape:
trolling slip events and Sh will fluctuate with Pp as a
poroelastic response. If so, Sh increases to its post-slip K Ic
value with gradually increasing Pp , as shown in the σf = (6)
Y πc
latter part of cycle a (Figure 8A). Note that cycle b
starts at a higher Pp because Sh is at a higher level than
at the start of cycle a. Recharging Pp will periodically Here the crack-normal tensile stress σf is referred
to as the fracture stress for a given flaw (i.e., the tensile
reactivate slip, with each reactivation cycle at an incre-
stress necessary to extend a flaw).
mentally higher Sh. This process will not regulate Pp in
When fractures are driven by abnormal fluid pres-
zone FOUR at a gradient near 0.9 psi/ft because fric-
sures, the process is called natural hydraulic fracturing
tional slip always maintains Sh at a level above Pp. To
(Secor, 1965, 1969; Engelder and Lacazette, 1990). The
illustrate this point, we take the derivative of Sh with
stress available to initiate a natural hydraulic fracture
respect to Pp in equation 3 to derive is referred to as the driving stress (σ ∆; Pollard and
Segall, 1987) and is given by
–2
∆Sh  1 
∆Pp
(

)
= 1 –  µ2 + 1 2 + µ

(4) σ ∆ = Pp + σ 3 (7)

where Pp = the ambient pore pressure in the rock and


Equation 4 indicates that, in the absence of sedi- σ3 = the minimum principal remote total stress, with
mentation and tectonics, Sh increases at a rate less than compressive stress being negative. Note that in the
Pp and that the two converge at S v . The difference context of LEFM, natural hydraulic fracturing takes
between Pp and Sh is smaller at the beginning of cycle b place if Pp exceeds the minimum principal total stress
than at the beginning of cycle a. In summary, in order by an amount equal to the fracture stress (σf ) of the
to maintain a Pp gradient at 0.9 psi/ft by leakage rock as given by equation 6. For the formation of ver-
through fault slip, a combination of three independent tical fractures, σ 3 is equivalent to S h, the minimum
variables (Pp, Sv , and S h) must act in unison. Other- total horizontal stress, and a general initiation crite-
wise, the interaction between Pp and Sh will eventually rion for vertical natural hydraulic fracturing is
drive Pp to a lithostatic pressure. obtained by substituting σ f for σ ∆ and rearranging
equation 7 to arrive at
Leakage Through a Network of Joints
We may appeal to a second mechanism to regulate  K Ic 
Pp = Sh +   (8)
Pp at a limit less than lithostatic pressure. Like leakage  Y πc 
by fault slip, the Pp limit by leakage through joints is
governed by a failure criterion involving Pp necessary
Equation 8 illustrates that vertical natural hydraulic
to open joints and the interaction between Pp and Sh.
fractures can initiate only when the ambient pore pres-
The criterion for opening joints (i.e., cracks) is taken
sure equals the sum of the rock fracture stress and the
from linear elastic fracture mechanics (LEFM). The
minimum total horizontal stress acting to keep flaws
governing equation for joint propagation is based on
closed. If flaws are pervasive joints with a large length,
the crack-tip stress intensity factor (K), which charac-
terizes the crack-tip stress field. For cracks loaded 2c, the second term of equation 8 goes to zero, thus
solely by a far-field tensile stress (σt ) normal to the implying that joints open and leak when Pp = Sh .
crack, the crack-tip stress intensity factor is given by Like leakage by fault slip, our premise is that leak-
age through a joint network is periodic and takes place
mainly when Pp ≥ Sh . The situation during and after
K I = Yσ t πc (5) joint opening and propagation is illustrated on a Pp-
Sh-time plot for a typical South Texas field where Pp
where Y and c are parameters characterizing the shape prior to joint opening is 0.90Sv below the zone THREE–
and size of the crack; tensile stresses are positive; and FOUR transition (cycle w in Figure 8B). One difference
the I denotes pure tensile loading (i.e., mode I; Lawn between leakage by joint opening and by fault slip is
and Wilshaw, 1975; Broek, 1986). The LEFM criterion that when Pp = 0.90Sv, Sh is lower at joint opening than
for the initiation of crack propagation states that a at fault slip (0.90S v vs. 0.92S v). As in fault slip, the
crack will extend, thus increasing its size, when the enhanced permeability from joint opening up to the
stress intensity at the tip of the crack reaches some crit- zone THREE–FOUR transition may allow the section
ical value. This critical value of stress intensity is below to reequilibrate at a hydrostatic gradient before
denoted KIc, and is commonly referred to as fracture resealing during cycle w.
toughness. The parameters controlling the initial prop- Like leakage during fault slip, Sh changes by poro-
agation of macroscopic joints from smaller cracks (i.e., elastic behavior. With a drop in Pp, poroelastic relaxation
A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields 265

LEAKAGE BY FAULT SLIP Figure 8. (A) Theoretical


pressure-stress-time plot
slip & illustrating the change in
fracture Sv = lithostatic stress pore pressure, Pp, and least
horizontal stress, Sh,
pressure & stress

before and after frictional


slip on a fault-cutting
Sh pressure zone FOUR.
(B) Theoretical pressure-
stress-time plot illustrat-
ing Pp and Sh before and
recharge Sh = 0.9 Sv after opening of joints
leak
through pressure zone
Pp seal FOUR.
cycle a cycle b cycle c cycle d
A
time

LEAKAGE BY JOINT OPENING


Sv = lithostatic stress
pressure & stress

crack &
open

Sh = 0.9 Sv
Sh

Pp leak recharge

B cycle w cycle x cycle y cycle z


seal
time

allows a drop in S h (Biot, 1941; Kümpel, 1991). where ν is the drained Poisson ratio. Rewriting equa-
According to the theory, a shrinkage ∆V of the rock tion 11, we find that
with initial volume, V, is induced by an decrease in
pore pressure, ∆Pp, ∆Sh 1 – 2v
=α (12)
∆Pp 1– v
∆V
= αβ∆Pp (9)
V If α = 1 and ν = 0, a change in Pp will induce the
same change in Sh. Otherwise, changes in Sh are smaller
where ∆V is controlled by the Biot coefficient of than changes in Pp , as is the case for frictional slip and
effective stress, α, and the compressibility of the rock, primeval elasticity. Anderson et al. (1973) were among
β (Detournay et al., 1989). Assuming that volume the first to suggest that poroelastic behavior was an
strain is zero where the compressibility is β = important element in the stress–pore pressure inter-
(1/V)(∆V/∆Pc), equation 9 may be rewritten as play leading to variation in leakoff pressure.
The sign of the net change in Sh distinguishes the
∆Pc = α∆Pp fault-slip leakage from joint leakage (Figure 8). Signifi-
(10)
cantly, after the closing or sealing of joints, poroelastic
behavior will drive the Sh back to its starting position
where Pc is the confining pressure (Segall, 1992). For for another cycle of both joint opening and fault slip.
α < 1, a change in Pp causes a smaller change in Pc. An This happens because Pp increases faster than S h .
exact equation may be derived for ∆Sh/∆Pp for total Unlike fault slip, where slip causes Sh to operate inde-
horizontal stress under uniaxial strain conditions: pendently of Pp in returning to a post-slip value, dur-
ing joint opening, Sh operates in concert with Pp so that
v 1 – 2v only one independent parameter operates. The depen-
Sh = Sv + α Pp (11)
1– v 1– v dent variable, Sh, does not show a net change unless
266 Engelder and Leftwich

helped by distant tectonics. If the mechanism for regu- Biot, M.A., 1941, General theory of three-dimensional
lating Pp is leakage through open joints, then Pp serves consolidation: Journal of Applied Physics, v. 12,
as a measure of the least stress in South Texas, with Pp p. 155–164.
= 0.86Sv (Figure 6). Broek, D., 1986, Elementary engineering fracture
mechanics: Boston, Martinus Nijhoff, 516 p.
Further Remarks Cartwright, J.A., 1994, Episodic basinwide hydrofrac-
turing of overpressured Early Cenozoic mudrock
Until better in-situ data are obtained, the nature of sequences in the North Sea Basin: Marine and
the mechanism for leakage from and subsequent seal- Petroleum Geology, v. 11, p. 587–607.
ing within pressure zone FOUR will remain debatable. Dainelli, J., and A. Vignolo, 1993, Regional assessment
Little is known about whether or not the thick shale of sealing faults through pore-pressure analysis in
section of South Texas contains joints or microfrac- the northern Adriatic: First Break, v. 11, p. 287–294.
tures such as those seen in the Travis Peak Formation Detournay, E., A. Cheng, J-C. Roegiers, and J.D.
of East Texas (Laubach, 1988, 1989). Likewise, the McLennan, 1989, Poroelasticity considerations in in
shales of South Texas may contain smaller scale but situ stress determination by hydraulic fracturing:
more pervasive faults such as those seen in the chalks International Journal of Rock Mechanics and Min-
of the North Sea (Cartwright, 1994). The discovery of ing Science, v. 26, p. 507–513.
either of these two deformation mechanisms will have
Dickinson, G., 1953, Geological aspects of abnormal
a great bearing on our interpretation of leakoff from
reservoir pressures in Gulf Coast Louisiana: AAPG
and subsequent sealing within pressure zone FOUR.
Bulletin, v. 37, p. 410–432.
Engelder, T., and M.P. Fischer, 1994, Influence of poro-
CONCLUSIONS elastic behavior on the magnitude of minimum hor-
izontal stress, Sh , in overpressured parts of
The geostatic ratio in the deep Gulf Coast basin of sedimentary basins: Geology, v. 22, p. 949–952.
South Texas is reached at depths between 10,500 and Engelder, T., and A. Lacazette, 1990, Natural hydraulic
2300 ft (3.2–3.8 km), as indicated by Pp data from seven fracturing, in N. Barton and O. Stephansson, eds.,
gas fields. This maximum is characterized by a geosta- Rock joints: Rotterdam, A.A. Balkema, p. 35–44.
tic ratio that averages 0.86 among the seven fields. Our Fertl, W.H., 1976, Abnormal formation pressures: New
interpretation is that the point λ marks pressure York, Elsevier Scientific Publishing Co., 382 p.
required for the slow leakoff of fluids in the deep basin Galloway, W.E., D.K. Hobday, and K. Magara, 1982,
by either leakage after fault slip or leakage through nat- Frio Formation of the Texas Gulf Coast Basin—
ural hydraulic fractures, or both. This situation depositional systems, structural framework, and
requires a mechanism for self-sealing to prevent Pp hydrocarbon origin, migration, distribution, and
from draining to a hydrostatic gradient. The difference exploration potential: The University of Texas at
between leakage after fault slip and leakage by jointing Austin, Bureau of Economic Geology Report of
is that the Sh does not return to its pre-slip magnitude Investigations 122, 78 p.
upon recharging of the reservoir, whereas this is the Grauls, D.J., and J.M. Baleix, 1994, Role of overpres-
case for recharging following leakage by jointing. sures and in situ stresses in fault-controlled hydro-
Although leakage by fault slip cannot be ruled out as a carbon migration: a case study: Marine and
mechanism for maintaining a Pp gradient greater than Petroleum Geology, v. 11, p. 734–742.
hydrostatic in pressure zone FOUR, leakage by jointing Hart, B.S., P.B. Flemings, and A. Deshpande, 1995,
is a simpler model because Sh can vary with Pp through Porosity and pressure: role of compaction disequilib-
poroelastic relaxation and recharging. rium in the development of geopressures in a Gulf
Coast Pleistocene basin: Geology, v. 23, p. 45–48.
Hottman, C.E., and R.K. Johnson, 1965, Estimation of
ACKNOWLEDGMENTS formation pressures from log-derived shale proper-
ties: Journal of Petroleum Technology, p. 717–722.
This work was supported by Contract 5088-260-
1746 from the Gas Research Institute and Penn State’s Hubbert, M.K., and W.W. Rubey, 1959, Role of fluid
Seal Evaluation Consortium. We thank Mark Fischer, pressures in mechanics of overthrust faulting: I,
Bill Higgs, Bruce Hart, and David McConaughy for Mechanics of fluid-filled porous solids and its
constructive reviews of an early version of this paper. application to overthrust faulting: Geological Soci-
ety of America Bulletin, v. 70, p. 115–166.
Hubbert, M.K., and D.G. Willis, 1957, Mechanics of
REFERENCES CITED hydraulic fracturing: Journal of Petroleum Technol-
ogy, v. 9, p. 153–168.
Anderson, R.A., D.S. Ingram, and A.M. Zanier, 1973, Kümpel, H.-J., 1991, Poroelasticity: parameters
Determining fracture pressure gradients from well reviewed: Geophysical Journal International, v. 105,
logs: Journal of Petroleum Technology, v. 26, p. 783–799.
p. 1259–1268. Laubach, S.E., 1988, Subsurface fractures and their
Berg, R.R., 1975, Capillary pressure in stratigraphic relationship to stress history in East Texas Basin
traps: AAPG Bulletin, v. 59, p. 939–956. sandstone: Tectonophysics, v. 156, p. 37–49.
A Pore-Pressure Limit in Overpressured South Texas Oil and Gas Fields 267

Laubach, S.E., 1989, Paleostress directions from the faulting: New York, Cambridge University Press, 439 p.
preferred orientation of closed microfractures Schowalter, T.T., 1979, Mechanics of secondary hydro-
(fluid-inclusion planes) in sandstone, East Texas carbon migration and entrapment: AAPG Bulletin,
Basin, U.S.A.: Journal of Structural Geology, v. 11, v. 63, p. 723–760.
p. 603–612. Secor, D.T., 1965, Role of fluid pressure in jointing:
Lawn B.R., and T.R. Wilshaw, 1975, Fracture of brittle American Journal of Science, v. 263, p. 633–646.
solids: London, Cambridge University, 204 p. Secor, D.T., 1969, Mechanics of natural extension frac-
Leftwich, J.T., 1993, The development of zones of turing at depth in the Earth’s crust; in A.J. Baer and
undercompacted shale relative to abnormal subsur- D.K. Norris, eds., Proceedings, Conference on
face pressures in sedimentary basins: Ph.D. disser- Research in Tectonics (Kink Bands and Brittle
tation, Pennsylvania State University, University Deformation): Geological Survey of Canada Paper
Park, Pennsylvania, 442 p. 68-52, p. 3–48.
Leftwich, J.T., and T. Engelder, 1995, The characteris- Segall, P., 1992, Induced stresses due to fluid extrac-
tics of geopressure profiles in the Gulf of Mexico tion from axisymmetric reservoirs: Pure and
Basin, in P. Orteleva, ed., Pressure compartments: Applied Geophysics, v. 139, p. 535–560.
AAPG Memoir 61, p. 119–129. Sibson, R.H., 1992, Implications of fault-valve behav-
Magara, K., 1978, Compaction and fluid migration: ior for rupture nucleation and recurrence: Tectono-
practical petroleum geology: New York, Elsevier physics, v. 211, p. 283–293.
Scientific Publishing Co., 319 p. Suppe, J., and J. Wittke, 1977, Abnormal pore-fluid
Matthews, V.M., and J. Kelly, 1967, How to predict for- pressures in relation to stratigraphy and structure
mation pressure and fracture gradient: Oil and Gas in the active fold-and-thrust belt of northwestern
Journal, v. 65, p. 92–106. Taiwan: Petroleum Geology of Taiwan, v. 14,
Pollard, D.D., and P. Segall, 1987, Theoretical displace- p. 11–24.
ments and stresses near fractures in rock: with Vavra, C.L., J.G. Kaldi, and R.M. Sneider, 1992, Geo-
applications to faults, joints, veins, dikes, and solu- logical applications of capillary pressure: a review:
tion surfaces, in B. Atkinson, ed., Fracture mechan- AAPG Bulletin, v. 76, p. 840–850.
ics of rock: Orlando, Academic Press, p. 227–350. Weedman, S.D., A.L. Guber, and T. Engelder, 1992,
Powley, D.E., 1990, Pressures and hydrogeology in Pore pressure variation within the Tuscaloosa
petroleum basins: Earth-Science Reviews, v. 29, trend: Morganza and Moore-Sams Fields, Louisiana
p. 215–226. Gulf Coast: Journal of Geophysical Research, v. 97,
Sahay, B., and W.H. Fertl, 1989, Origin and evaluation p. 7193–7202.
of formation pressures: Dordrecht, Kluwer Acade- Zoback, M.D., and J.H. Healy, 1984, Friction, faulting,
mic Publishers, 292 p. and in situ stress: Annals of Geophysics, v. 2,
Scholz, C.H., 1990, The mechanics of earthquakes and p. 689–698.
Maucione, D.T., and R.C. Surdam, 1997, Seismic response char-
acteristics of a regional-scale pressure compartment bound-
ary, Alberta Basin, Canada, in R.C. Surdam, ed., Seals, traps,
and the petroleum system: AAPG Memoir 67, p. 269–281.
Chapter 16

Seismic Response Characteristics of a


Regional-Scale Pressure Compartment
Boundary, Alberta Basin, Canada
Debi T. Maucione
Exxon Exploration Company
Houston, Texas, U.S.A.

Ronald C. Surdam
Institute for Energy Research, University of Wyoming
Laramie, Wyoming, U.S.A.

ABSTRACT
Acoustic impedance contrast “dim spots” (previously described as the
“fuzzy effect” in Maucione, 1993) have been associated with anomalously
pressured hydrocarbon accumulations in the Powder River Basin of
Wyoming (Boyd et al., 1996a, b). It is important to know if the observations
made in the Powder River Basin apply to anomalously pressured hydrocar-
bon accumulations in other Rocky Mountain Laramide basins. In order to
test the applicability of these concepts in another basin, a ~58 km (36 mi)
long seismic profile from the Alberta Deep Basin of western Canada has
been constructed and analyzed. The profile crosses various pressure bound-
aries, but the reservoir geometry and seismic acquisition and processing
parameters remain nearly constant throughout the length of the profile.
These parameters may thus be eliminated as causes of seismic anomalies
that could falsely indicate anomalous pressures.
Anomalous pressures in the Alberta Deep Basin were originally identified
using well log and production information. However, using seismic data
acquired and processed to detect character response changes other than
those associated with structural features, one can easily observe deviation
from the expected increase in velocity with depth. Understanding the exact
character of anomalous seismic responses requires knowledge of some geo-
logic data, including rock properties (e.g., lithology and velocity) and depo-
sitional environments present in a similar geologic setting (i.e., another
Laramide basin). Additionally, integration of well log and production infor-
mation corroborates the seismic character response changes associated with
a regional-scale pressure boundary. However, it is important to note that
identification of areas of anomalous pressure using seismic data is not
dependent on well information.

269
270 Maucione and Surdam

In the present study, four seismic response characteristics of the seismic


data from the Alberta Deep Basin indicated that a regional-scale pressure
compartment boundary, separating areas of normal and anomalous pres-
sure, was present in the basin. Two of these characteristics can be
observed in a routine processing flow. The other two characteristics can
be observed only after an extremely detailed velocity analysis.

INTRODUCTION direction are of noticeably better quality, both in fre-


quency range and amplitude, than the data acquired
The application of seismic reflection data to identify in the strike direction.
and delineate pressure boundaries has been discussed Because the pressure compartment boundary was
for years (Pennebaker, 1968). Because it is a multivari- not determined to be associated with a reservoir inter-
ate problem, unique solutions are difficult to obtain. val but rather with a thick shale sequence, a discussion
For example, one could argue that a difference in seis- of resolution will not be given here. Using the three-
mic velocity response is caused by a change in lithol- eighths wavelength criteria (Yilmaz, 1987) with an aver-
ogy, results from fluid saturation characteristics, or age velocity of 3333 m/sec and a frequency of 32 Hz,
reflects a change in pressure regime. The answer is not the minimum detectable thickness is 39 m (128 ft).
one of these options exclusively, but rather is a combi- Because the shale section is at least three times thicker
nation of factors. Previous work has shown that than the minimum detectable thickness, resolution of
regional-scale pressure boundaries can be determined the stratigraphic interval is not a problem (see Mau-
from well-log data (Wang, 1992; Jiao et al., 1994; Mau- cione, 1996).
cione et al., 1994; Surdam and Jiao, 1994). Recent work The 2-D seismic data were processed to preserve
(Boyd et al., 1996a, b; Maucione, 1996) shows that the amplitude information and to determine detailed
same boundaries can be mapped by using surface seis- velocity functions. Line 3 contained industrial noise;
mic data. The following discussion will illustrate one therefore, a reduced frequency range was passed
successful example of pressure boundary description (12.5–17–45–60 Hz), compared with Lines 1 and 2
by two-dimensional (2-D) seismic reflection data. (8–12–65–85 Hz). A general description of the process-
The study area is located in a portion of the Western ing flow is illustrated in Figure 4.
Canada Basin referred to as the Northwest Plains of These lines were processed with nearly identical
the Alberta Basin and includes coverage into a region steps. The seismic data were obtained in such a way as
called the Deep Basin (Masters, 1984). A profile trend- to maintain identical field acquisition parameters. It is
ing approximately south-southwest was constructed encouraging to see consistent results from velocity
over a distance of ~58 km (36 mi) through an area that functions hand-selected every 125 m (410 ft) over a
includes parts of the Wapiti contract (Figure 1). It is distance of ~61 km (36 mi). The velocity functions at
immediately south and slightly east of the well-known the points where the seismic lines tie to each other
Elmworth Field (Masters, 1984). were checked to ensure that they were identical, in
order to prevent the formation of discontinuities when
the lines were tied together.
METHODOLOGY
Two 120-fold, regional Vibroseis™ seismic lines VARIATIONS IN SEISMIC CHARACTER
were combined with part of a third line to form a profile AS INDICATORS OF A PRESSURE
~58 km (36 mi) in length. Table 1 shows representative COMPARTMENT BOUNDARY
acquisition parameters for the three lines that form the
profile. The profile trends north-northeastward; it is at A number of conventional techniques can be used
an oblique angle to the basin’s structural axis (Figure in nonstandard ways to obtain detailed velocity and
2). It also parallels nearly half of a gas/pressure profile stratigraphic information from seismic data. Changes
(Figure 3) constructed from well logs by Jiao and Sur- in velocity can signal changes in lithology, porosity,
dam (1996). density, pressure regime, fluid saturation, and temper-
The northernmost regional seismic line, Line 1, is ature (Sheriff, 1992). The 2-D seismic data such as that
most nearly oriented in the direction of the basin’s used in this study can provide a continuous sampling
structural strike. The middle seismic line, Line 2, is ori- [common depth point spacing of 12.5 m (41 ft)] and
ented in the basin’s dip direction. The southernmost yield a more accurate picture of lateral variation than a
regional line, Line 3, is again oriented nearly in the prediction made by interpolating between wells. The
basin’s strike direction. The data acquired in the dip detailed velocity function that was created during data
Seismic Response Characteristics of a Regional-Scale Pressure Compartment Boundary 271

T Figure 1. Location map of


73 the study area with the
T CONTRACTS OF THE seismic profile indicated by
72
ELMWORTH ELMWORTH/WAPITI the thick line. Arrow points
CONTRACT AREA
Reserve based T
Grande 71
AREA along general line of seis-
Prairie mic profile. Modified from
T
70
Smith (1984).
N
T
69

WAPITI CONTRACT AREA T


Reserve based 68

T
67

T
British Columbia

66
Alberta

T
65
GOLD CREEK SOUTH
CONTRACT AREA T
Deliverability KARR
64
CONTRACT
AREA T
Deliverability 63
R 13 R 12 R 11 R 10 R9 R8 R7 R6 R5 R4 R3 R2 R1 W6

processing was hand-selected every 125 m (410 ft) hor- (3500–3900 m/sec) in the next 800 msec. This increase
izontally and at ~30-msec intervals vertically through is 40% of what it was in the first 800 msec. If this
the section of interest. When picking velocities from change were caused by a processing artifact, it would
semblance coherency plots, it is typically possible to occur at the same time on each line.
distinguish velocities at 30-msec intervals (see Mau-
cione, 1996, appendix D).
Table 1. Data Acquisition Parameters, Line 1.
Velocity Variations Company Enertec Geophysical
Stacking Velocity Services
Crew 176
During seismic processing, a change in the typical Date August 1993
stacking velocity analysis semblance plot can be Location T65-68N, R07W6
observed on individual gathers. In the region where Source Vibroseis, 4, Hemi 48, 30 m
the Lower Cretaceous is predominantly water satu- Sweeps 6/station, 8 sec, 12–90 Hz
rated and normally pressured, the velocity semblance Shooting Direction South-North
plot has more variation than in the anomalously pres- Spread Split: –3025 to –50 m,
sured, gas-saturated area (Figure 5). In Figure 5, the 50–3025 m
velocities in the Cretaceous section, ~1000–1500 msec, Gap 2 stations (50 m)
vary from 3400–4000 m/sec, but the velocities in Fig- Shot Point Interval 25 m
ure 6 are nearly constant, 3500–3600 m/sec, over the Seismometer Group
same time interval. Interval 25 m
A normal move-out (NMO) velocity was used to Seismometer Groups 9/group, inline, 3.125 m
correct the quasi/hyperbolic primary reflection to Seismometers Mark L15/L10, 14 Hz,
time alignment with zero source-receiver offset. For a 60% damping
horizontally stratified medium, the velocity required Amplifiers DFS-Vx2, 48 dB pre-amp
for NMO correction (also called stacking velocity), is gain
equal to the root mean squared (RMS) velocity (Yil- Filters low = out, high = 128 Hz
maz, 1987). Filter Ramps none
As seen in Figures 5 and 6, the typical velocity 60 Hz Notch out
range for rocks in the Cretaceous section is 3300–3600 Sample Rate 2 msec
m/sec. This corresponds to the blue-green portion of Record Length 4 sec
the velocity color scale in Figure 7. A marked color Line Length ~15.5 mi
gradient change occurs in the velocity function at Number of Traces 244,680
~800 msec. Whereas it took only 800 msec for the Maximum Fold 120
velocities to increase through half of the total spec- Disk File Size 1.06 Gb
trum (2500–3500 m/sec), the increase is much less
272 Maucione and Surdam

Figure 2. Schematic
diagram of the Deep Basin
location in relation to the
Disturbed Belt in Alberta,

a
British Columbi
Canada. From Smith (1984).

Alberta

Fort St. John

ELMWORTH/WAPITI
AREA
Grande Prairie

Edmonton
D
D

E
E
IS

P
T

BA
U
R

SI
B

\
E

N
D
BE

Calgary
LT

0 50 100 mi

0 100 km

However, as shown in Figures 7–9, the change does instance, the boundary’s TWTT increases as one
not always occur at the same time. Figure 7 is a portion moves south (toward the basin center ) along the pro-
of the velocity function for Line 1 plotted as stacking file. Adjusting lines to individual datums is one reason
velocity vs. two-way traveltime (TWTT). The color for such differences; Line 1 datum is 875 m (2870 ft), Line
scale represents velocities ranging from 2500 m/sec in 2 datum is 1076 m (3530 ft), and Line 3 datum is 1090 m
purple to 4500 m/sec in red. Figures 8 and 9 are the (3580 ft). More significantly, the boundary's TWTT
velocity functions determined for lines 2 and 3. The increase is also consistent with the fact that these strati-
horizontal and vertical scales and the velocity color graphic units are more deeply buried as one moves
gradient scale are the same as in Figure 7. The bound- southward toward the basin axis (Line 1 to Line 3).
ary between normally pressured (magenta to blue-
green) and anomalously pressured (green to yellow) Interval Velocity
sections can be seen in all these figures. Although the Interval velocity provides a useful correlation with
general character of the velocity functions pictured in sonic logs when the velocity is made a function of
Figures 7 to 9 is similar, there are some differences. For depth rather than TWTT. Using a smoothed gradient
Seismic Response Characteristics of a Regional-Scale Pressure Compartment Boundary 273

S N Figure 3. An anomalous
pressure/gas profile
2000 constructed from sonic logs
(Jiao and Surdam, 1996).
The left half of the profile
parallels the seismic profile
4000
in this study.
Depth, ft

1
6000

8000 2
1 Cardium
3 2. Cadotte
10000
3. Falher
5 15 25 35 45 55
Distance, mi

Increasing Anomalous Pressure/Gas Saturation

method provided by the processing software, Pro- pressured section (0–1150 msec). Since amplitude
MAX, the stacking velocity was used to calculate an information was preserved during processing (the
associated interval velocity. (The smoothed gradient only amplitude corrections applied were a surface
method will not create large velocity discontinuities amplitude correction and a spherical divergence cor-
between blocked intervals, as does the Dix equation, rection), the correlation of this amplitude response
which calculates the velocity between two parallel change with the location of the top of the regional-
reflectors, given the associated stacking velocities.) scale anomalous pressure boundary is interesting. Jiao
Figure 10 illustrates the same region and bound- and Surdam (1996; their figure 34) show that the top of
aries as in Figure 7, but it plots interval velocity vs. the anomalously pressured section is in the shale sec-
depth. This display clearly shows that the interval tion of the Upper Cretaceous between the Cardium
velocity (as determined by conversion of stacking and Cadotte formations. When the top predicted by
velocity vs. TWTT to interval velocity vs. depth) the amplitude change in seismic response is converted
increases steadily with depth until ~1500 m (4920 ft), to depth, one calculates 1780 m (5842 ft), as compared
the predicted pressure compartment top boundary. with ~1767 m (5900 ft) predicted from well logs (Mau-
Figure 11 is a similar interval velocity vs. depth dis- cione, 1996).
play for Line 3. Although the horizontal and vertical
scales are similar to those in Figure 10, the velocity
scale now ranges from 2500 to 5000 m/sec. Line 3 is INTEGRATION OF WELL
from a deeper portion of the basin, and it contains a INFORMATION
greater thickness of Tertiary cover. Both characteris-
tics contribute to the thicker section present in the In order to determine which stratigraphic interval is
steadily increasing velocity function zone. In Figure associated with the pressure compartment boundary,
11, the normally increasing section is present from 0 the thickness of geological units must be known. This
to 2000 m (0–6560 ft) depth. The area shown in Figure information was determined using well logs, which
11 is ~19.4 km (12 mi) from that in Figure 10, but are the best source available to this study. Ten wells
despite this distance, a similar change in velocity with log suites consisting of deep resistivity, gamma-
function occurs when crossing the top pressure com- ray, sonic, bulk density, and compensated neutron
partment boundary. porosity logs were examined. A stratigraphic correla-
tion was constructed. Using sonic and density logs
Stacked Section and the tops determined by the stratigraphic correla-
Comparison of a stacked seismic section that is dis- tion, synthetic seismograms were generated for the
played without automatic gain control (AGC) (Figure wells. Each well log was adjusted to its kelly bushing
12) with a seismic section that is displayed with AGC (KB) as datum, and visually matched to the seismic
(Figure 13) reveals an obvious lack of reflection ampli- data. The reflection coefficient series was convolved
tude (no yellow or red present) in the upper normally with a wavelet extracted from the final stacked section.
274 Maucione and Surdam

Patch and order master and slave traces


Filter, Spherical divergence correction
Surface consistent amplitude correction
Air blast attenuation (air wave and ground roll)
Trace edits

Load survey and navigation information


Establish geometry
CDP sort
Quality check CDP and shot gathers

First break mutes


Semblance velocity analysis to pick NMO velocities
3-D delay time refraction statics (incl. elev. statics)
Spiking and predictive decon, surface consistent
(shot and receiver)

Velocity analysis using constant velocity stacked gathers


Merge velocity functions
Smooth merged function; horiz.=50 CDPs, vert.=150 ms
Velocity analysis using IVA
Hand select function every 10 CDPs (125 m)
Residual statics correction using Maximum Power
Autostatics; two horizons

Offset bin sort


Common offset F-K DMO
IVA on DMO corrected CDPs
Final stack

Figure 4. Schematic diagram of the processing flow used for this data. The data sets from all
three lines followed this path. CDP = common depth point; DMO = dip move-out; IVA =
integrated velocity analysis.

This section did not have automatic gain control good correlation between the predictions given by
(AGC) applied, and the wavelet was extracted from interval velocity and those from sonic logs. Since inter-
the seismic section where the well tied to the seismic val velocity was determined from stacking velocity,
line. Using these well ties, the stratigraphic horizons and interval velocity can be associated with geologic
of interest were interpreted. This enables the identi- features, one can further associate stacking velocity
fication of intervals in terms of both time and depth. with geologic features in this study area.
Another method of generating synthetic seismo- It seems likely that there is a physical reason
grams was also used, but with less success (see Mau- beyond the changeover from mechanical compaction
cione, 1996). to chemical compaction to account for this velocity
Many people avoid associating changes in stacking gradient change. The first change in gradient, at ~800
velocity with geologic features. In this study, there is msec, can be identified in terms of stratigraphic units
Seismic Response Characteristics of a Regional-Scale Pressure Compartment Boundary 275

Figure 6. A velocity coherence plot similar to Figure


5, except from a region where the Lower Cretaceous
Figure 5. A velocity coherence plot illustrating sem- section (between arrows) is gas saturated and anom-
blance velocities in a region where the Lower alously pressured. The velocity variation in the sec-
Cretaceous section (between arrows, 1000–1500 tion is virtually eliminated, and the average velocity
msec) is predominantly water saturated and normal- for this section is significantly lower than the equiv-
ly pressured. Here the semblance velocities vary alent section in Figure 5.
between 3400 and 4000 m/sec.

by using the synthetic seismograms generated from have considerably faster stacking velocities than sand
the well-log data. It corresponds to a thick shale sec- and shale. The nearly solid red color “below” 1700
tion in the Upper Cretaceous Kaskapau Formation. As msec represents velocities of ≥4500 m/sec. Since a typ-
mentioned, this is a likely candidate for the pressure ical carbonate velocity is about 6000 m/sec, carbonates
compartment top boundary. would be seen in red using this velocity scale.
Another marked change in color gradient occurs at
~1600 msec. Integrated stratigraphic information indi-
cates that this change coincides with a major unconfor- DISCUSSION
mity that separates the Cretaceous section from
the underlying Jurassic and older rocks. The pre- This paper presents four distinct methods of detect-
Cretaceous section differs greatly in lithology from the ing a pressure compartment boundary through use of
post-Jurassic because it contains abundant carbonates. 2-D reflection seismic data: examination of (1) stacked
Carbonate rocks in this depth range that lack porosity sections, (2) velocity semblance coherency plots,
276 Maucione and Surdam

Figure 7. A portion of the


stacking velocity function
for Line 1. As seen in
Figures 5 and 6, velocities in
the Cretaceous section range
from 3300 to 3600 m/sec (the
blue-green portion of the
velocity color scale). A
marked color gradient
change occurs at 800 msec.
This change occurs in a
thick shale section of the
Upper Cretaceous Kaskapau
Formation. The top pressure
compartment boundary is
marked by the dotted line at
the blue-green color change
(~800 msec).

(3) stacking velocities, and (4) interval velocities. The velocity difference between sands and shales is greatly
first two methods do not require extensive additional reduced (Figure 6), as shown by a lack of impedance
work beyond a typical processing flow. Comparison contrast, or a “dim spot.” Such observation does not
of the final stacked section display, with and without require time-intensive processing steps, such as
AGC applied, is very simple; it requires only that the detailed velocity analysis or migration.
data be processed to preserve amplitude and dis- After integration of the well data, synthetic seismo-
played without the additional scaling factors that are grams, and seismic data, the location of this regional-
usually applied to aid interpretation. The reduction in scale pressure compartment boundary is confirmed
velocity below the pressure boundary is easily observ- and further refined. In the study area, the compart-
able on semblance plots during processing if one takes ment boundary corresponds to the Kaskapau shale
into consideration the particular characteristics that section in the Upper Cretaceous, where a marine
constitute a “normal” difference in sand and shale incursion is associated with a maximum flooding sur-
velocities and velocity responses in the study area. The face and argillaceous sediments (Leckie, 1989). This

Figure 8. Illustration of the


top pressure compartment
boundary for a portion of
the stacking velocity func-
tion for Line 2. The vertical,
horizontal, and velocity
scales have been kept con-
stant.
Seismic Response Characteristics of a Regional-Scale Pressure Compartment Boundary 277

Figure 9. Illustration of the


top pressure compartment
boundary for a portion of
the stacking velocity func-
tion for Line 3. The vertical,
horizontal, and velocity
scales have been kept
constant. Note that as one
moves southward in the
basin, Line 1 to Line 3,
the top pressure compart-
ment boundary occurs at
increasing travel time. This
is consistent with the fact
that as one approaches the
basin axis, the deeper
(increased travel time) the
stratigraphic unit is found.

event is likely synchronous with the early Turonian The maximum flooding surface in the Kaskapau
maximum transgression of the Cretaceous Interior shale is present at 1500 m (4920 ft) in Figure 10 and
Seaway evident on the eustatic curve (Haq et al., 1987). shown in Figure 11. Below this depth, the interval
This type of surface is a time line. While the surface velocity decreases slightly and then remains nearly
represents an event in time (Posamentier and Allen, constant (at this scale) down through the rest of the
1994), it connects different stratigraphic units as deter- Cretaceous section. The lower interval velocities (yellow-
mined by lithology (Figure 14). Thus, if the boundary green color) present at ~2600–2700 m (8530–8860 ft)
is actually a time surface, it will appear to cut across depth correspond to the depth range of some of the
traditional stratigraphic boundaries. This can explain lower Falher member reservoir units. This lower inter-
how a top compartment boundary can cut across val velocity area could indicate increased porosity, gas
stratigraphy when its location is predicted by log data, saturation, or an increase in the amount of anomalous
but appear horizontal when predicted by seismic data. pressure.

Figure 10. The same region


and boundaries as seen in
Figure 7, but with interval
velocity in terms of depth
instead of stacking velocity
in terms of time. Below the
top pressure compartment
boundary, shown by the
dotted line at 1500 m, the
interval velocity remains
constant or decreases
slightly. The area of
significantly lower interval
velocities (yellow-green)
corresponds to the depth of
the lower Falher units. This
lower-interval-velocity area
could indicate increased
porosity, gas saturation,
or increased anomalous
pressure.
278 Maucione and Surdam

Figure 11. Representation similar


to that in Figure 9, with the interval
velocity function from a portion of
Line 3. A similar interval velocity
response is seen below the top
pressure compartment boundary.

Further comparison of Figures 10 and 11 indicates less velocity variation. In addition, finer grained rocks
that there is more variation in the interval velocity that are buried deeper have less bulk storage capacity,
below the top boundary of anomalous pressure in Fig- and therefore give increased seismic velocities.
ure 10 than in Figure 11. In fact, nearly horizontal Picking a very detailed velocity function requires
stripes of common velocity alternating between 4500 extensive velocity analysis, as well as many iterations
and 4000 m/sec are present in Figure 10. In contrast, and the use of several methods (Figure 4) to obtain the
below the top boundary (~2000 m depth) in Figure 11, most detailed and consistent results. However, if the
the interval velocity remains nearly constant. This dif- result of this additional effort leads to the reliable
ference may be due to a less variable stratigraphic sec- identification of a feature such as the top of a pressure
tion (Maucione, 1996). The densities in this part of the compartment boundary, it is worthwhile. Recent stud-
Cretaceous section do not vary along the profile (shale ies have demonstrated that the largest hydrocarbon
≈ 2.35, sand ≈ 2.65). The calculated porosities (Mau- reserves are located within ~610 m (2000 ft) of the top
cione, 1996) average 10%–15% (includes all lithologies: of a pressure compartment boundary (Leach, 1994). If
sand, shale, and coal). Farther south in this study area, seismic data can be used to target this highly success-
from Line 1 to Line 3, the lithology contains finer ful region near this top boundary before incurring the
grained constituents and is less varied, and sand depo- additional expense of drilling a well, the economic
sition has been more restricted, which also may result in value of a prospect could be much improved.

Figure 12. A portion of


the final stacked section
along Line 1 displayed
without automatic gain
control (AGC) applica-
tion. Note that there is a
lack of reflection ampli-
tude (yellow and red
areas) from the surface
down to ~1200 msec.
The purple horizon is
the Cardium sandstone
and the green horizon
corresponds to the
Dunvegan Formation.
The change in reflec-
tion amplitude seen
here occurs between
these geologic units.
Seismic Response Characteristics of a Regional-Scale Pressure Compartment Boundary 279

Figure 13. The same dia-


gram as in Figure 12, except
displayed with automatic
gain control (AGC) applied.
When AGC is used, the lack
of reflection amplitude in
the shallower section is not
seen. Although AGC is usu-
ally applied to data before
interpretation, this is one
example where comparison
of the stacked section with
and without AGC applied
can clearly illustrate a fun-
damental seismic response
characteristic— lack of
reflection amplitude.

High

7
6
5
4
3
2
1

A. Relative
7 Sea Level
6 Low
5 Maximum
4 Flooding
Time
Depth

Surface
3
2

Landward Seaward

B.
7

5
Time

4
Maximum Flooding Surface

Landward Seaward

Figure 14. Illustration of the difference between a maximum flooding surface represented in terms
of (A) depth and (B) time. (After Posamentier and Allen, 1994). If a boundary or seal is associated
with a maximum flooding surface, it will appear as a time line on a seismic section, but will seem
to cut across lithological boundaries established on the basis of traditional borehole
information such as core descriptions.
280 Maucione and Surdam

CONCLUSIONS investigator, Natural gas resource characterization


study of the Mesaverde Group in the Greater Green
Recent studies of pressure compartments have River Basin, Wyoming: a strategic plan for the
shown that the greatest amount of hydrocarbon exploitation of tight gas sands: Gas Research Insti-
reserves are located within ~610 m (2000 ft) of the top tute Topical Report, p. 53–72.
of a pressure compartment boundary (Leach, 1994). Haq, B.U., J. Hardenbol, and P.R. Vail, 1987, Chronol-
Our study has illustrated four seismic response char- ogy of fluctuating sea levels since the Triassic:
acteristics that can be used to identify a pressure com- Science, v. 235, p. 1156–1167.
partment boundary in the Alberta Deep Basin. There Jiao, Z.S., and R.C. Surdam, 1996, Pressure regime in
are two necessary conditions for such identification. the vicinity of the Elmworth Area, Western Canada
First, the seismic data must be processed to preserve Basin, in R.C. Surdam, principal investigator, Nat-
amplitude information and examined without apply- ural gas resource characterization study of the
ing scaling factors such asAGC (automatic gain con- Mesaverde Group in the Greater Green River Basin,
trol). Second, the time and effort to determine an Wyoming: a strategic plan for the exploitation of
extremely detailed velocity function may be necessary tight gas sands: Gas Research Institute Topical
to allow the other three observations: velocity sem- Report, p. 29–40.
blance coherency plots, stacking velocity, and interval Jiao, Z.S., H.Q. Zhao, and R.C. Surdam, 1994, Pressure
velocity variations. Although four variations are compartmentalization in Cretaceous shales and
shown here, one method, interval velocity, is superior sandstones in the Washakie and Denver basins:
to the others. When the top boundary is determined Annual report on contract No. 5089-260-1894, mul-
using interval velocities calculated from seismic tidisciplinary analysis of pressure chambers in the
response, the top boundary can be consistently identi- Powder River Basin, Wyoming and Montana: a
fied. The average error in the seismic prediction is only new, innovative exploitation strategy for gas accu-
8% of that of the sonic log interval velocity (Maucione, mulations within pressure compartments: Chicago,
1996). Thus, seismic interval velocities are reliable Gas Research Institute, p. 39–82.
indicators of a pressure compartment boundary in the Leach, W.G., 1994, Distribution of hydrocarbons in
Alberta Deep Basin. If seismic data can be used to abnormal pressure in south Louisiana, U.S.A., in
delineate such a high-success region before drilling a W.H. Fertl, R.E. Chapman, and R.F. Hotz, eds.,
well, economic value is added to the prospect (Mau- Studies in abnormal pressures: developments in
cione, 1996). petroleum science, 38: New York, Elsevier,
p. 391–428.
Leckie, D.A., 1989, Upper Zuni Sequence: Upper Cre-
ACKNOWLEDGMENTS taceous to Lower Tertiary, in B.D. Ricketts, ed.,
Western Canada Sedimentary Basin, a case history:
Funding for this study was provided by the Insti- Calgary, Canadian Society of Petroleum Geologists,
tute for Energy Research, University of Wyoming, p. 269–281.
under Gas Research Institute Contract No. 5089-260- Masters, J.A., 1984, Lower Cretaceous oil and gas in
1894. Canadian Hunter Exploration Ltd. provided the Western Canada, in J.A. Masters, ed., Elmworth,
seismic and well data used in this study, as well as case study of a deep basin gas field: AAPG Memoir
access to extensive databases for the Alberta Basin and 38, p. 1–33.
personal expertise. Ward-X Data Services Ltd. pro- Maucione, D.T., 1993, Seismic reflection delineation of
vided field tape copies of the seismic data. Advance abnormally pressured zones in the Powell Field of
Geophysical Co. ProMAX 2-D was used to process the the Powder River Basin, Wyoming: Master’s thesis,
seismic data, and GeoQuest IES was used for interpre- University of Wyoming, Laramie, 143 p.
tation. Special thanks go to Sharon Kubichek for her Maucione, D.T., 1996, Geologic significance and poten-
help with seismic data processing. tial economic impact of seismic signatures in pres-
sure compartments: a study of the Cadotte and
REFERENCES CITED Falher reservoirs in the Alberta Deep Basin,
Canada: Ph.D. dissertation, University of
Boyd, N.K., S.W. Kubichek, H.P. Heasler, S.B. Smith- Wyoming, Laramie, Wyoming, 297 p.
son, and R.C. Surdam, 1996a, Seismic response of Maucione, D.T., V. Serebryakov, P. Valasek, Y. Wang,
abnormally pressured gas compartments, Powder and S.B. Smithson, 1994, A sonic log study of abnor-
River Basin, Wyoming, in R.C. Surdam, principal mally pressured zones in the Powder River Basin of
investigator, Natural gas resource characterization Wyoming, in P. Ortoleva, ed., Basin compartments
study of the Mesaverde Group in the Greater Green and seals: AAPG Memoir 61, p. 333–348.
River Basin, Wyoming: a strategic plan for the Pennebaker, L.S., Jr., 1968, Seismic data indicate depth,
exploitation of tight gas sands: Gas Research Insti- magnitude of abnormal pressures: World Oil,
tute Topical Report, p. 5–28. v. 166, no. 7, p. 73–77.
Boyd, N.K., S.W. Kubichek, and S.B. Smithson, 1996b, Posamentier, H., and G. Allen, 1994, Siliciclastic
Remapping the boundary between gas-saturated sequence stratigraphy—concepts and applications:
and water-saturated rocks in the Northeast Elm- AAPG Short Course 10 Notes, 89 p.
worth Field, Alberta Basin, in R.C. Surdam, principal Sheriff, R.E., 1992, Petrophysical and geophysical
Seismic Response Characteristics of a Regional-Scale Pressure Compartment Boundary 281

background basic petrophysics and geophysics, in annual report on contract No. 5089-260-1894, multi-
R. E. Sheriff, ed., Reservoir geophysics: Investiga- disciplinary analysis of pressure chambers in the
tions in Geophysics 7, p. 37–49. Powder River Basin, Wyoming and Montana: a
Smith, R.D., 1984, Gas reserves and production perfor- new, innovative exploitation strategy for gas accu-
mance of the Elmworth/Wapiti area of the Deep mulations within pressure compartments: Chicago,
Basin, in J.A. Masters, ed., Elmworth—case study Gas Research Institute, p. 5–38.
of a deep basin gas field: AAPG Memoir 38, Wang, Y., 1992, Velocity study of overpressured zones in
p. 153–172. the Powder River Basin of Wyoming using sonic logs:
Surdam, R.C., and Z.S. Jiao, 1994, Executive summary: Master’s thesis, University of Wyoming, Laramie, 92 p.
anomalously pressured gas accumulations in Creta- Yilmaz, O., 1987, Seismic data processing: Tulsa, Soci-
ceous Rocks of the Laramide Basins of Wyoming: ety of Exploration Geophysicists, 526 p.
Surdam, R.C., 1997, A new paradigm for gas explo-
ration in anomalously pressured “tight gas sands”
in the Rocky Mountain Laramide Basins, in R.C.
Surdam, ed., Seals, traps, and the petroleum sys-
tem: AAPG Memoir 67, p. 283–298.
Chapter 17

A New Paradigm for Gas Exploration in


Anomalously Pressured “Tight Gas Sands” in
the Rocky Mountain Laramide Basins
Ronald C. Surdam
Institute for Energy Research, University of Wyoming
Laramie, Wyoming, U.S.A.

ABSTRACT
A significant portion of the Cretaceous shales in the Rocky Mountain
Laramide Basins (RMLB) are overpressured on a basinwide scale. The
change of pressure regime from normally pressured to overpressured coin-
cides with marked changes in the geochemical and geophysical properties of
the Cretaceous rock/fluid system. Sandstone bodies within the overpres-
sured shale section are subdivided stratigraphically and diagenetically into
relatively small, isolated, gas-saturated, anomalously pressured compart-
ments. The driving mechanism of the pressure compartmentalization is the
generation and storage of liquid hydrocarbons that subsequently react to
gas, converting the fluid-flow system to a multiphase regime in which capil-
larity controls permeability.
A new exploration paradigm and an exploitation strategy have been creat-
ed that significantly reduce exploration risk in the RMLB. Two elements cru-
cial to the development of prospects in the deep, gas-saturated portions of
the RMLB are (1) the determination and, if possible, three-dimensional eval-
uation of the pressure boundary between normal and anomalous pressure
regimes and (2) the detection and delineation of porosity/permeability
“sweet spots” (i.e., areas of enhanced storage capacity and deliverability) in
potential reservoir targets below this boundary. Certainly there are other
critical aspects, but completion of these two tasks is essential to the success-
ful exploration for the unconventional gas resources present in anomalously
pressured rock/fluid systems in the RMLB.

INTRODUCTION anomalously pressured (pressure gradient ≠0.40–


0.50 psi/ft) rock/fluid systems (Surdam et al., 1995,
Only recently have techniques been developed to 1997). Numerous gas discoveries made in the West-
examine in detail the boundary between normally ern Canada Basin in the last decade demonstrate that
pressured (pressure gradient = 0.40–0.50 psi/ft) and the delineation of this pressure boundary is an

283
284 Surdam

gas accumulations, results can be disappointing and


costly. Thus, delineating zones of enhanced storage
capacity (i.e., porosity) and deliverability (i.e., perme-
ability), or “sweet spots,” is also an important part of
the new exploration paradigm.
In order to demonstrate the applicability of a new
exploration paradigm, this chapter will focus specifi-
cally on the Cretaceous sandstones of the Mesaverde
Group in the Washakie Basin, but will be generally
applicable to all basins of the RMLB, which have very
similar fluid-flow characteristics. It is not the goal of
this chapter to provide a detailed overview of the pres-
sure regimes in the RMLB, or even in the Washakie
Basin. Previous research on the Mesaverde Group has
been published in annual and topical technical reports
available through the Gas Research Institute (GRI)
(Surdam et al., 1993, 1995; Iverson, 1995; Jaworoski et
al., 1995).
The aim of this paper is to (1) summarize previous
work on the delineation the pressure boundary in the
Washakie Basin, (2) discuss a methodology for detec-
tion and delineation of hydrocarbon production sweet
spots beneath the basinwide pressure boundaries in
Laramide basins, and, most importantly, (3) discuss
the new exploration paradigm (Figure 3) for basin-
center, or deep-basin gas, in the RMLB.

GEOLOGICAL SETTING
The Washakie Basin is the easternmost subbasin of
Figure 1. Index map showing the locations of most the Greater Green River Basin (GGRB). The subbasin
of the Rocky Mountain Laramide Basins (RMLB). is a symmetrical structural and topographic basin
bounded on the east by the Sierra Madre, on the
north by the Wamsutter arch, on the west by the Rock
important aspect in searching for unconventional gas Springs uplift, and on the south by Cherokee Ridge
accumulations in basins characterized by anom- (Figure 5). It is 42 mi (68 km) north-south and 54 mi
alously pressured sections (Surdam et al., 1995). For (87 km) east-west; its area is roughly 2200 mi 2 (3550
example, Leach (1994), using an enormous database km2). The surface elevation in the basin ranges from
in the Louisiana Gulf Coast, demonstrated that the 6100 to 8700 ft and averages 7000 ft (2130 m). The dip
majority of oil and gas discoveries are within 2000 ft of the Cretaceous strata into the basin is approxi-
(610 m) of the top of the anomalous pressure (i.e., the mately 8° along the eastern flank and 15° along the
pressure boundary). Surdam et al. (1995) have shown western flank (Love, 1970). Precambrian rocks lie at
that 80% of the gas production from Cretaceous rocks depths of >32,000 ft (9750 m) at the center of the
in the Rocky Mountain Laramide Basins (RMLB) Washakie Basin. Approximately 32,000 ft (9750 m) of
(Figure 1) is from an interval extending from the Cambrian through Tertiary sedimentary rocks are
pressure boundary down to 2000 ft (610 m) below present in the deeper part of the Washakie Basin
this boundary (Figure 2). Surdam et al. (1995) suggest (Hale, 1961). The deepest part of the Upper Cretaceous
that establishing the shape of the top of the anom- section, in the basin’s center, is at ~14,000 ft (4270 m)
alously pressured rock can be critical in the detection depth (Roehler, 1969).
of commercial gas accumulations (Figure 3). Positive The present study is concerned with the Upper Cre-
relief on the pressure boundary surface commonly taceous rocks in the basin, which include, from oldest to
can be correlated to commercial gas accumulations youngest, the Mesaverde Group (especially the Almond
(Figure 4). Formation), the Lewis Shale, the Fox Hills Sandstone,
In addition, unlike conventional gas accumulations and the Lance Formation. The Lewis Shale, which repre-
(e.g., with gas cap, oil leg, and oil/water contact), sents a major marine transgression, is ≤2700 ft (820 m)
accumulations associated with anomalously pres- thick, and is composed mainly of shale. The shale of the
sured regimes (e.g., depletion drive) like those in the lowermost Lewis is black, carbonaceous, and biotur-
RMLB (Figure 1) are not necessarily constrained by bated; in places, shell debris is abundant. The Lewis sea
either structural closure or stratigraphic pinch-out. opened eastward into the main part of the North Ameri-
Consequently, when conventional exploration tech- can Interior Seaway (Winn et al., 1985). The maximum
nology is used in the search for these unconventional extent of the Lewis transgression was to areas of the
A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 285
Distance to Top of Anomalous Pressure, ft

Distance to Top of Anomalous Pressure, ft


Cretaceous Major Gas Reservoir, Bighorn Basin Cretaceous Major Gas Reservoir, Powder River Basin
-5000 -5000
-4000 -4000
-3000 -3000
-2000 -2000
-1000 -1000
0 0
1000 1000
2000 2000
3000 3000
Reservoir: 22 Reservoir: 58
4000 CGP: 776 bcf 4000
CGP: 1456 bcf
5000 5000
0 100 200 300 400 500 0 200 400 600 800
Cumulative Gas Production, bcf Cumulative Gas Production, bcf
Distance to Top of Anomalous Pressure, ft

Distance to Top of Anomalous Pressure, ft


Cretaceous Major Gas Reservoir, Washakie Basin Cretaceous Major Gas Reservoir, Wind River Basin
-5000 -5000
-4000 -4000
-3000 -3000
-2000 -2000
-1000 -1000
0 0
1000 1000
2000 2000
3000 3000
Reservoir: 26 Reservoir: 32
4000 CGP: 1642 bcf 4000
CGP: 863 bcf
5000 5000
0 200 400 600 800 0 200 400 600 800
Cumulative Gas Production, BCF Cumulative Gas Production, bcf

Figure 2. Plots of cumulative production vs. distance to the top of the anomalously pressured zone for major
gas reservoirs (>5 bcf) in the Washakie, Powder River, Bighorn, and Wind River basins. Plots indicate that
>80% of the gas production from Cretaceous rocks in these basins is from an interval extending from the pres-
sure compartment boundary down to 2000 ft (610 m) below the boundary.

Rock Springs uplift in the west and the Wind River PRESSURE BOUNDARY
uplift in the north. The Almond Formation is the
uppermost sandstone of the Mesaverde Group (Miller, The trend in acoustic velocity on sonic logs is a con-
1977). It was deposited in marginal-marine and ventional method used to detect anomalous pressure
marine environments during the final transgression of in shales (Powley, 1982, personal communication) and
the Cretaceous seaway into southwestern Wyoming. to delineate the pressure boundary. Typically, a
The Almond ranges in thickness from 300 to 800 ft marked decrease in velocity (i.e., increase in sonic-
(90–240 m) (Roehler, 1990), and is composed of sand- transit time) is interpreted as indicating lower effective
stone, siltstone, shale, and coal beds. rock stress and, hence, overpressuring (Figure 4). Pow-
Depositional patterns in the Upper Cretaceous ley (1982) cites many examples of pressure compart-
Mesaverde Group and the Lewis Shale show a contin- ments delineated on the basis of analyzing trends in
uous trend toward strandline irregularity. The upper transit time from sonic logs. Almost all of the sonic
part of the Lewis Shale and the Fox Hills Sandstone logs from the Washakie Basin have a marked decrease
represent the withdrawal of the Cretaceous seaway in velocity in the deeper part of the well logs similar to
from the area (Law et al., 1986). The overlying Lance that illustrated in Figure 4. Thus, the pressure bound-
Formation was deposited primarily in an alluvial- ary can be neatly mapped regionally using this
plain environment. In response to continued Laramide marked decrease in acoustic velocity.
deformation and the emergence of adjacent foreland In order to study the basinwide configuration of
uplifts, an internal drainage system developed on the overpressuring in the Washakie Basin, three west-east
Lance Formation. A similar trend in the eventual cross sections across the basin were constructed from
development of internal drainage during the Late Cre- 41 sonic logs filtered to include only fine-grained rocks
taceous has been noted in the Uinta Basin of Utah (defined as those intervals with gamma-ray log values
(Fouch et al., 1983). greater than 65° API). Figure 5 shows the locations of
286 Surdam

Figure 3. Schematic
UNCONVENTIONAL CONVENTIONAL
diagram illustrating the
two elements crucial to
hydrocarbon exploration
in gas-saturated, anom-
alously pressured rock:
(1) the pressure boundary
and (2) sweet spots. Gas
accumulations below the
pressure boundary are
al independent of structural
rm ure
no ss closure or stratigraphic
e
"Pressure Seal" pr pinch-out.

s
ou
al re
om su
an res
p "Sweet Spots"

these cross sections, which are denoted T13N, T15N, Velocity from Sonic Logs
and T17N for the townships they occupy. The three
cross sections are of almost equal length and run from
R102W to R91W.
Figure 6A–C provides a better understanding of the 2000
pressure regime in the Upper Cretaceous shales in the
Washakie Basin. These figures show a panel of sonic
logs with the velocities corrected for compaction using
the method outlined by Surdam et al. (1995; Figure 7). 4000
Surdam et al. (1995) outline in detail the detection and
delineation of the regional pressure boundary in this
basin. The following is a brief synopsis of the results
reported: 6000
Depth, m

• In all areas studied away from the Washakie


Basin margin, the onset of anomalous pressures
(the overpressure top) is marked by a transitional 8000
zone of slightly overpressured rocks, ~500–1500 ft
(~152–457 m) thick. It appears that overpressur-
ing within the Cretaceous shales developed
regionally as one basinwide compartment (Law, 10000
1984). In the Washakie Basin, the present-day
depth of the overpressure top ranges from 8000 to
10,000 ft (2440–3050 m).
• Cross section T13N (Figure 6A), in the southern 12000
part of the Washakie Basin, shows that the Upper
Cretaceous Almond Formation is buried to
>14,000 ft (4270 m) depth [equivalent elevation,
14000
–7000 ft (–2130 m)]. The top of overpressuring
cuts across structural and stratigraphic bound-
5000 3000
aries and is not horizontal, but has a wavy form. Velocity, m/s
Toward the west, the top of overpressuring occurs
at about 8000 ft (~2440 m) depth [equivalent ele- Figure 4. Plot of velocity vs. depth for the
vation, –1000 ft (–310 m)], and toward the east at McPherson Springs 14-2 well, Sec. 14, T13N, R94W,
9000 ft (~2745 m) depth [equivalent elevation, Washakie Basin.
A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 287

108°

109° 28
N
27

26

25

SIN
24

GREAT DIVIDE BA
42°

ROCK
23

RA LIFT
UP
WL
INS
22

SPRINGS
Rawlins
21

Wamsutter 20
WAMSUTTE
N

R
ASI

AR 19
CH
RB

Rock Springs
IVE

18
EN R

Green River
T17N
GRE

UPLIFT

16

T15N

N
SI
BA
KI

E
14
HA
AS T13N
W

R 111 W 110 108 106 WYOMING 101 99 97 95 IDGE 93 91 89 R 87 W


41° R
T 24 CHEROKEE 12
COLORADO

3 20
R 25 EN R 18 E
UTAH

R 25 E 11
SAND
WAS
H
10
BAS IN

9
0 50 MILES
8

107°
7

Mesaverde Group outcrops


6

T
5
N

Figure 5. Location map of the sonic log profiles collected in the Washakie Basin. The cross sections shown are
T13N, T15N, and T17N.

–2000 ft (–610 m)]. In the eastern portion, some other cross sections. The boundary between nor-
lenticular compartments are shown. mal and anomalous pressure appears horizontal,
• Cross section T15N (Figure 6B), in the central part and occurs at ~11,000 ft (3350 m) depth [equiva-
of the basin, shows the Almond Formation is lent elevation, –2000 ft (–610 m)]. Some lenticular
buried down to 15,000 ft (4570 m) depth [equiva- compartments are shown.
lent elevation, –8000 ft (–2440 m)]. The top of over- • Viewed south to north, the top of overpressuring
pressuring is almost horizontal and occurs at in the Washakie Basin appears to be bowl-shaped,
~10,000 ft (3050 m) depth [equivalent elevation, deeper in the center and shallower at the north
–3000 ft (–910 m)]. and south margins.
• Cross section T17N (Figure 6C), in the northern
part of the basin near the Wamsutter arch, shows In summary, the top of overpressuring in the
that the Almond Formation is buried more shal- Washakie Basin is an uneven surface. Above the over-
lowly there than in other areas of the basin, at pressured zones, there is a transitional zone that is
<13,000 ft (3960 m) present-day depth [equivalent thicker in some areas than in others. Through the use
elevation, –6000 ft (–1830 m)]. The lower part of of the sonic-log panels, the top of the overpressured
the basin appears to narrow more than along the zone can be clearly determined. However, the bottom
288 Surdam

A Velocity Profile, T13N, Washakie Basin


Figure 6. Gridded
After Decompaction Correction velocity profiles, after
4000 W E
decompaction, derived
9500 from a panel of sonic
2000 9350 logs through the
9200 Lower Tertiary and
Upper Cretaceous

Velocity, ft/s
9050

Almo
Elevation, ft

0
8900
stratigraphic section in
the southern Washakie
nd 8750
Basin, Wyoming. (A)
-2000 8600
Decompacted velocity
8450
profile for cross
8300 section T13N, R102W-
-4000
8150 91W. (B) Decompacted
8000 velocity profile for
-6000 cross section T15N,
R102W-91W. (C)
1.8E4 5.4E4 9.0E4 1.3E5 1.6E5 2.0E5 2.3E5 2.7E5 3.1E5 3.4E5 Decompacted velocity
Distance, ft profile for cross
section T17N, R102W-
B Velocity Profile, T15N, After Decompaction
91W. For localities of
cross sections, see
W Washakie Basin, Wyoming E Figures 1–5.
4000

9500
2000 9250
9000

Velocity, ft/s
8750
Elevation, ft

0
8500
AL

8250
-2000
MO

8000
ND

ND

7750
O

-4000
M

7500
AL

7250
-6000 7000

-8000
1.8E4 5.4E4 9.0E4 1.3E5 1.6E5 2.0E5 2.3E5 2.7E5 3.1E5 3.4E5
Distance, ft

C Velocity Profile, T17N, Washakie Basin


4000 W After Decompaction Correction E
9500
2000 9350
9200
Velocity, ft/s

9050
Elevation, ft

0
8900
Al
m

8750
on
d

-2000 8600
8450
8300
-4000
8150
8000
-6000

1.8E4 5.4E4 9.0E4 1.3E5 1.6E5 2.0E5 2.3E5 2.7E5 3.1E5 3.4E5
Distance, ft
A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 289

After Decompaction Correction Porosity (%)


0 10 20 30 40 50
0
0 Avg Porosity

2000

Depth (ft x 1000)


5
4000

10 Tight Gas Sand


6000 Sweet Spot
Depth, ft

8000 15
Figure 8. Plot of porosity vs. depth for Almond
sandstones from 2300 measurements in the
10000 Washakie Basin. Shaded area shows a sweet spot in
these tight gas sandstones.
PZone:
12000
11680-12934
Mesaverde interval. Figure 8 is an example of sweet spot sand-
stones in the Almond Formation of the Washakie
Basin. As Iverson (1995) points out, these storage/
14000 deliverability sweet spots translate directly into
enhanced hydrocarbon production.
To illustrate the importance of sweet spots to gas pro-
16000 duction in anomalously pressured rocks, consider the
1000 2000 3000 4000 5000 giant Hoadley gas field in the Western Canada Basin
(Figure 9). As Chiang (1984) pointed out, “[t]he discovery
Velocity, m/s of the Hoadley gas field not only adds a great amount of
gas reserve in Canada [~10% at 6–7 tcf] but strongly indi-
Figure 7. Plot of velocity vs. depth corrected for com- cates the overlooked potential of both tighter sand reser-
paction for the McPherson Springs 14-2 well, Sec. voir rock and sweet spots in it. Prior to discovery, the
14, T13N, R94W, Washakie Basin. Cross-hatching Hoadley glauconitic sand bar had been penetrated by
shows area where rocks are anomalously pressured hundreds of wells, none recognizing that this sand bar
and gas saturated. complex contains a giant reserve of recoverable gas.” In
brief, the keys to unlocking the hydrocarbon reserves at
Hoadley (6–7 tcf of gas and 350–400 million bbl of con-
of the overpressured zone has not been identified densate) were recognizing that (1) a regional pressure
regionally, as only a few wells have been drilled through boundary existed and (2) that sand ridges and conglom-
the basinwide overpressured shale compartment in the eratic beaches represented sweet spots below the pres-
Washakie Basin. In the paper, the discussion is directed sure boundary (Figure 10). Once explorationists
at the exploration for gas beneath the pressure bound- recognized these keys, the focus of their efforts shifted
ary, in the gas-saturated portion of the column. from conventional hydrocarbon accumulations in struc-
tural and stratigraphic traps in the normally pressured,
SWEET SPOTS water-saturated portion of the column to porosity/per-
meability sweet spots below the pressure boundary in
Sweet Spot Definition anomalously pressured, gas-saturated rock. The spectac-
ular results of this exploration shift are summarized in
As shown in Figure 3, the key to gas production Figure 10. The history of the giant Hoadley gas field dis-
below the regional pressure boundary is the detection covery is a beautiful example of the importance of sweet
and delineation of so-called “sweet spots.” In this spot delineation and detection when searching for gas
paper, sweet spots are defined roughly as those reser- beneath the regional pressure boundary.
voir rocks that are characterized by porosity and per-
meability values greater than the average values for
Sweet Spot Importance
tight sands at a specific depth interval. In other words,
sweet spot sandstones represent enhanced hydrocar- The importance of porosity/permeability sweet spots
bon storage and deliverability within a target reservoir to production in basins characterized by regionally
290 Surdam

Northwest Territories NORMALLY


Gas Well
PRESSURED
mbia
Canadian Sand Ridge
Alberta Shield 0 10 mi
British Colu

UNDER
PRESSURED

Saskatchewan
Peace River

E
IN
Arch

AR
R

M
Elmworth BA
IER

N
RR

O
BA
F o R o Dis

O
G
Edmonton
o t ck

LA
ll s
hi y

o Hoadley
M

OVER
un
tu

ta

PRESSURED
\
rb

ins
ed

Milk River
Be

Calgary Figure 10. Map showing facies of the Hoadley barri-


lt

ss er bar complex. Well distribution is related to the


0 100 200 ra pressure boundaries. Modified from Chiang (1984).
Ar g
t
ee
ch

Miles
Sw

larger volumes than is possible from the sweet spot


Figure 9. Location map showing the largest gas alone. An example of this production mechanism is
fields in the Alberta Basin of Canada. Modified shown in Figure 11. As a consequence, fields like the
from Chiang (1984). Standard Draw-Echo Springs not only have produced
more gas than could be stored in the sweet sand (i.e.,
an Almond sandstone), but also show little or no
significant anomalous pressures is undeniable. This is drawdown after 14 years of production (Figure 12;
because sweet spots typically provide a large continu- Iverson and Surdam, 1995). Iverson (1995) discusses in
ous horizontal fluid conduit into an otherwise inacces- detail the significance of reservoir sweet spots to pro-
sible large volume of low-permeability rock (e.g., tight duction in the Washakie Basin.
sandstone). Above the pressure boundary, the rocks
tend to be more porous and permeable due to less bur-
Sweet Spot Characteristics
ial and, consequently, less cementation. Also, rocks
above the boundary tend to be in a single-phase fluid- Sweet spots are characterized by important attri-
flow system; here, low-permeability rocks tend to be butes in addition to enhanced porosity and permeabil-
flow barriers and not seals, except very locally and in ity. One of these is the tendency of a particular
close proximity to conventional hydrocarbon accumu- depositional facies to have the maximum sweet spot
lations. In contrast, rocks below the pressure bound- potential in a formation. Typically, in every basin and
ary are in a multiphase fluid-flow system, and targeted reservoir interval, one or two specific deposi-
rock-fluid characteristics regionally are dominated by tional facies will comprise the porosity/permeability
capillarity. As a consequence, low-permeability rocks sweet spot. In the Hoadley field, the best reservoir
below the pressure boundary, including tight sands, facies are the eolian beach ridges and conglomeratic
become capillary seals. Simply stated, these tight beaches; in the Mesaverde Group in the Washakie
sands give up their fluids only when their displace- Basin, the group of marine bar sands collectively
ment or threshold pressures are exceeded. Thus, the known as the Upper Almond bar sands comprise the
importance of sweet spots becomes apparent; sweet sweet spot (Figure 13). Thus, one of the first steps in
spots allow the fluid flow in the reservoir volume, evaluating targets for gas production beneath the pres-
including both the sweet spot and adjacent tight sure boundary is the determination of which deposi-
sands, to be manipulated. For example, pressure dif- tional facies has highest potential for enhanced storage
ferentials between a sweet spot and the adjacent tight capacity and deliverability.
sands can be created such that displacement pressures For those basins with production histories, the
can be overcome, resulting in the drainage of much size of the sweet spot targets can be determined by
A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 291

Lewis Shale
PD = 2000 psi
Upper Almond Bar Sand
PD = 20 psi Pf = 5200 psi
(Pp = 3000 psi) φ = 12%
K = 0.2 md

Lower Almond
Laminated Sands
PD = 400 psi Pf = 5200 psi
(Pp = 4800 psi) φ = 3%
K = 0.002 md
In order for gas to move
from laminated sandstones
to "bar" sand ∆P > 400psi
Figure 11. Schematic diagram of the production mechanism at work in a sweet spot. The threshold pressure of
the sands must be exceeded to allow gas to migrate from the tight, laminated sandstones to the adjacent
porous, permeable sweet sands.

constructing a potentiometric surface map of the basin the pressure boundary is affected not only by the
(Figure 14). Our potentiometric surface constructions weight of the water column above the boundary but
utilize pressures from drill-stem tests (DSTs). All also by the weight of the rock and fluid column from
available nonzero data are used. Poor tests and data the pressure boundary down to the depth of the reser-
are indicative of low-permeability rocks and tests too voir. Under these conditions, there is no fluid continuity
short to accurately measure pressure. Good tests and between individual reservoirs; they are compartmental-
data are indicative of high-permeability rocks. Our ized and isolated (Figure 15).
potentiometric surface maps delineate which rocks are In contrast, if the reservoir sandstones show signifi-
permeable and which are not, and give good approxi- cant scatter from a single gradient below the pressure
mations of the size of permeable sweet spots within boundary, there is a much greater potential for fluid
the basin of interest (Figure 14). continuity (connectivity) between individual reser-
The degree to which the sweet spots are in fluid voirs or fields. The tendency of connected reservoirs
continuity is of particular importance in evaluating below the pressure boundary to follow a gas gradient
gas column heights and designing well stimulation with depth provides some evidence of this. In other
strategies. Fluid continuity commonly can be evalu- Laramide basins, we have noted that at and immedi-
ated by considering the initial production pressures of ately below the pressure boundary, individual reser-
all available reservoirs from a variety of fields vs. voirs follow a gas gradient before changing to a
depth (Figure 15). Typically, reservoirs above the pres- sandstone pressure gradient with greater depth (Fig-
sure boundary follow a hydrostatic gradient with ure 15). This suggests that, in some basins, there is
depth. Below the pressure boundary, reservoir sand- more reservoir connectivity than at the top of the
stones follow either a single, but different, gradient anomalously pressured rock.
with depth, or they exhibit significant scatter.
When the sandstones follow a single gradient, it is
Relationship of Diagenesis to Sweet Spots
a pressure-depth gradient parallel to the regional
lithostatic gradient but shifted to much lower pres- Yin and Surdam (1995) describe in detail the diage-
sures (Figure 15). In such a case, each reservoir below netic history of Almond Formation sandstones. The
292 Surdam

1000 cementation (Yin and Surdam, 1995). In contrast, many


of the other depositional facies in the Almond sand-
Gas Production (Bscf/year)

stones lose significant intergranular volume during the


burial due to compaction and grain deformation. Volu-
number of wells metrically, the most damaging cements in Mesaverde
100 Group sandstones are quartz and late dolomite. Another
important factor in porosity damage is the presence and
deformation of ductile lithic framework grains (espe-
cially in shoreface sandstones).
10 To find those Almond sandstones that are charac-
terized by enhanced porosity (i.e., sweet spots), explo-
Champlin #226 rationists should target sandstones where quartz and
late dolomite cementation have been inhibited and
lithic framework grains are sparse. The tidal channel
1 sandstones contain less lithic grains and late dolomite
decline at -0.0545/yr
abandon at 30 MMscf/yr cement than the shoreface sandstones. In addition,
EUR = 55 Bscf at 2068 abundant early calcite cement in the tidal channel
sandstones provided an opportunity for enhancement
0.1 of porosity through its subsequent dissolution. Thus,
1980 1984 1988 1992 the tidal channel sandstones are usually associated
YEAR with permeability sweet spots. However, some quartz-
rich sandstones are tightly cemented by quartz over-
Figure 12. Plot of gas production vs. time showing growths at great depth, where both porosity and
that production has declined much less than expect- permeability have been seriously damaged.
ed in the Standard Draw area. The five most important processes that inhibit
cementation during diagenesis are as follows:

Almond sandstones and, for that matter, all Mesaverde • Grain-rimming clay rims (inhibition of quartz
Group sandstones lose porosity with progressive burial cementation)
at a greater rate than is typical for most sandstones (Fig- • Dissolution of early carbonate cements (preserva-
ure 8). This loss of porosity is the result of both mechan- tion of intergranular volume)
ical compaction and cementation. Some Almond • Overpressuring (compaction retardation)
sandstones (i.e., tidal channel) lose very little intergranu- • Liquid hydrocarbon accumulation (cementation
lar volume (i.e., ~ minus — cement porosity) with pro- abatement)
gressive burial and suffer porosity loss primarily from • Fracturing (porosity/permeability enhancement)

,
,,
,, , ,
,,,
,,,
,,,
,,
,,
W E Figure 13. Diagrammatic
A A'
Lance Fm. Medicine Bow Fm.
cross section showing

,,
,,,
,,,
restored Upper Cretaceous
Fox Hills Ss.

,,,,,
,,,,,, ,,,,
,,
Lewis Sh. rocks across northern Utah
Almond Fm. and southern Wyoming.
Eri Rocks of continental origin
Mesaverde Fm. cso
nS up are shaded; alluvial-plain
s. ro

,,,,,,
,,
,, ,,,
,,,
,,,,
MOXA G Allen Ridge Fm.
Rock d e and marine-shoreline, shelf,
ARCH Springs ver .
slope, and basin sandstone
Adaville Fm. Fm.
sa . Fm
Blair Fm. e ts and siltstone units are
M M

,,,,,,
,,,
,, ,,,,,,,,,
,,,
,,
k
tac shown by dot pattern;
ys marine shale and limestone
Ha
Hilliard Sh. Baxter Sh. are unpatterned. The loca-

,,,,,,,,,,,
Steele Sh. tion of subaerially formed
Fr

unconformities is indicated
on

by the squiggly line.


tie
rF

FEET
Modified from Roehler

,
Fron Niob

,,,,,,,,,
m

1500 tier F rara F


.

WY m. m. (1990).
A' 0 25 MILES
A
UT CO 0
APPROXIMATE SCALES

Marine siltstone and shale Marine and transitional marine sandstone

Non-marine sandstone, siltstone and shale Non-marine to estuarine sandstone


A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 293

Figure 14. Contoured potentio-


Hydraulic metric surface of drill-stem test
data from the Mesaverde
Head (feet) Formation in the Washakie Basin.
10000 Modified from Heasler and
42.0° 9000 Surdam (1992).
8000
7000
6000
5000
Latitude

4000
41.5° 3000
2000
1000
0

41.0°

40.5°

109.0° 108.5° 108.0° 107.5° 107.0°


Longitude

Yin and Surdam (1995) have shown that in the Pressure, psi
Mesaverde Group sandstones (including the Almond 0 2000 4000 6000 8000 10000
sandstones) in the Washakie Basin, clay coats are volu- 0
metrically insignificant and dissolution of carbonate Reservoir Pressure
cements is only locally important. Moreover, as Surdam "L Displacement Pressure
et al. (1995) have demonstrated, overpressuring of the ith
0.4

os
Mesaverde Group is coincident with maximum burial ta
33

tic
and the oil-to-gas reaction. As a consequence, overpres- "G
psi

ra
/ft

suring of these rocks occurs relatively late in the burial di


en
history, after the sandstones have suffered extensive 5000
t,
1.
cementation. Thus, fracturing and/or early migration of 0
ps
liquid hydrocarbons appear to be the two processes that i/f
t
Depth, ft

might play a role in enhancing porosity/permeability in


the Mesaverde Group sandstones in the Washakie Basin. Top of Anomalous
Pressure
Relationship of Fracturing and Sh
Regional Lineaments to Sweet Spots a
Pr le
10000 Sa es
Iverson (1995) used oil and gas production data to nd su "D
st re isp
classify oil and gas fields within the Washakie Basin on "
G lac
e ra em
into four categories. Cumulative production of >$500 Pr di
es en en
million constitutes a largest field; $50–$500 million, a su t t
re
significant field; $5–$50 million, a small field; and <$5 G
ra
million, a smallest field. di
en
Figure 16 shows the spatial relationship between t
the regional lineaments and the hydrocarbon accumu- 15000
lations. There is an obvious relationship between the
distribution of hydrocarbon accumulations in fields Figure 15. Pressure profile for the major Cretaceous gas
with >$5 million cumulative production and the most reservoir pressures in the Washakie Basin. Above the
prominent regional lineaments. For example, most of pressure boundary, the reservoir pressure follows a
the accumulations in these fields are either cut or ter- hydrostatic gradient. Below this boundary, the reser-
minated by the lineaments (Figure 16). Conversely, voir pressure is parallel to the regional lithostatic gradi-
few of the smallest fields (<$5 million) are associated ent, but is offset to lower pressures than this gradient.
with mapped regional lineaments (Figure 16).
A good illustration of this relationship is the hydro- Springs field (Figure 17). One lineament forms the
carbon accumulation at the Standard Draw-Echo northernmost limit of production in this field, whereas
294 Surdam

Figure 16. Map showing the relationship between major lineaments and oil/gas fields in the Washakie Basin.
Color coding represents the dollar value of cumulative production of each hydrocarbon field. Note the rela-
tionship between major fields and regional lineaments. Lineaments mark the position of regionally signifi-
cant shear zones characterized by recurrent movement.

a second parallel subsurface lineament cuts through a Relationship of Source Rocks to Sweet Spots
saddle in the isopach map of the Almond bar sand-
stone. This saddle coincides with some of the most If the relationship between source rocks and reser-
productive wells in the field. The coincidence of thin- voir rocks is examined, the significance of fracturing
ning of the bar sand and position of the geophysical becomes even more evident. García-González et al.
lineament suggests that basement faults were reacti- (1993a, b; 1997) demonstrated that coal is the major
vated during the deposition of the bar sand. The fact source of liquid and gaseous hydrocarbons in the
that some hydrocarbon accumulations are bounded by Mesaverde Group sandstones in the Washakie Basin
the lineaments further supports the hypothesis that a by showing that (1) oil in the coal was generated
significant cause-and-effect relationship exists during the alteration of desmocollinite and liptinite
between the regional lineaments and the more signifi- macerals into waxy oil and inertinite solid residue;
cant hydrocarbon accumulations in the Washakie (2) the waxy oil was initially stored in porous struc-
Basin (e.g., sweet spots). tures and subsequently in vesicles as the coal
Fracturing of reservoir sandstones adjacent to matured under increasing temperature; (3) primary
these lineaments is prevalent (Dunn et al., 1995). migration of the oil occurred as the generation of a
Moreover, the smaller scale fracturing adjacent to the sufficient volume of waxy oil microfractured the vit-
regionally prominent lineaments enhances perme- rinite-semifusinite vesicles, interconnecting vesicles,
ability (Figure 18; Dunn et al., 1995). Most impor- and pores; and (4) the thermal cracking of waxy oil
tantly, Jaworowski et al. (1995) have demonstrated to gas generated a sufficient volume of gas to frac-
that the regional lineaments and the associated ture the vesiculated coal as pore pressure increased
smaller scale fractures in adjacent sandstones typi- and allowed expulsion and migration of hydrocar-
cally have the same spatial orientation. There is little bons out of the coal.
doubt that fracturing has improved deliverability in Figure 19 summarizes the maturation scenario for
sweet spot sandstones, and has probably signifi- coals in the GGRB and elsewhere. The only way waxy
cantly improved the connectivity of sweet spot sand- oil could have been expelled from the coal was either
stones and nearby tighter sandstones. by reacting to gas, as outlined above, or by tectonically
A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 295

R95W R93W R91W Figure 17. Map showing


thickness of the Almond
RED DESERT BASIN bar sandstone in Standard
Sentinel Draw-Echo Springs. Open
Ridge circles indicate wells with
10 Bcf production. Two
regional lineaments from
T Figures 1–16 are shown.
22 The top lineament forms
the northern terminus. The

20
N
bottom lineament inter-
Siberia

40
sects the bar sand on a sad-
Ridge ge dle on the isopach; this
i es Chan
lineament correlates with
the better wells in the field.

10

t
Pinchou
Note that no correlation
T exists between the bar sand
30

0 thickness and the most


c

20
productive wells. Modified
Fa

N from Christiansen, 1995.


0
10

T
20

18
N

30
WASHAKIE BASIN

Gas fields
10' contour interval
0 3 6 miles
]
> 10 Bcf

fracturing the coal during the generation of the oil. (Surdam et al., 1993) a new exploration paradigm has
Only fracturing would have resulted in the migration been constructed that significantly reduces explo-
of oil into adjacent sandstones before diagenesis was ration risk for for hydrocarbons in anomalously pres-
complete and the sandstones became cemented. sured, gas-saturated rocks beneath the pressure
Regional lineaments characterized by recurrent move- boundary in the Rocky Mountain Laramide Basins
ment would have been ideal locations for the early (RMLB) (Figure 3).
migration of liquid hydrocarbons out of coals into This paradigm, which focuses on the Almond For-
adjacent sandstones, which would have resulted in mation in the Washakie Basin, can be used as a basis
cementation abatement. Thus, early liquid hydrocar- for the construction of a new and more innovative
bon migration and fracturing (i.e., establishing exploitation strategy for gas exploration in anom-
source/reservoir connection) were key elements in the alously pressured “tight gas sands” in the RMLB. The
formation of reservoir sweet spots in the Mesaverde application of this paradigm depends on the comple-
Group sandstones of the Washakie Basin. tion of two crucial tasks: (1) determining and evaluat-
ing, in three dimensions, the boundary between
normal and anomalous pressure and (2) detecting and
NEW EXPLORATION PARADIGM delineating porosity/permeability sweet spots.
It is imperative to delineate the position of the
On the basis of the present discussion of reservoir boundary between overlying, normally pressured
sweet spots and previous discussion of pressure rock and the underlying, anomalously pressured rock,
regimes, regional lineaments, fractures, and stratigraphy because 80% of the cumulative gas production in each
296 Surdam

Immature Coal

Generation of oil
with storage
COAL or migration

With
Faulting

increasing thermal exposure


BURIAL
WET
GAS
Figure 18. Partially cemented (quartz and kaolinite)
fracture in the Upper Almond bar sand, northern Expulsion of
end of the Echo Springs field, Wyoming. "wet gas"
(oil ⇒ gas)

DRY
basin of the RMLB comes from the stratigraphic GAS
interval between the pressure boundary and 2000 ft
(610 m) below the boundary (Figure 2). The detection Expulsion of
of any three-dimensional relief on the pressure "dry gas"
boundary is also very important, particularly ele- (kerogen ⇒ gas)
vated portions of the surface. For example, Davis
(1984) showed that in the Western Canada Basin,
major anomalously pressured gas accumulations are
associated with relief on the surface of the pressure Figure 19. Schematic diagram of a typical maturation
boundary. If the pressure boundary surface cuts scenario for coal. Vertical sequence of blocks on the
across stratigraphic units, it can serve nicely as a seal left indicates maturation scenario in the absence of
or trap updip (Figure 3). Establishing the position of fracturing; the block on the right indicates matura-
the pressure boundary has also been shown to be tion scenario after fracturing has occurred.
important in the Louisiana Gulf Coast, which, unlike
the RMLB, has experienced little or no uplift and ero-
sion, but does exhibit anomalous pressures (Leach, (including both the sweet spot and the adjacent tight
1994). Thus, delineating the pressure boundary is sands) to be manipulated during production.
crucial to any exploration strategy in basins charac- The following methodology is used to translate
terized by an anomalously pressured stratigraphic this new exploration paradigm into an exploration
interval. strategy:
Another critical aspect of exploration for gas
below the regional pressure boundary in the RMLB • Determination of the position of the pressure
is the detection and delineation of reservoir sweet boundary (i.e., the boundary between normal and
spots, as they typically provide significant continu- anomalous pressure regimes);
ous horizontal fluid conduits into otherwise inacces- • Evaluation of the three-dimensional (3-D)
sible, large volumes of low-permeability rock (e.g., aspects of the pressure boundary surface, with
tight gas sands). The depositional settings of most of special emphasis on areas characterized by posi-
the basins of the RMLB were highly variable, so a tive relief;
wide variety of factors control the formation and • Determination of which depositional facies have
position of sweet spots in the anomalously pressured the greatest potential for enhanced storage capac-
section; very rarely is only a single factor responsible ity and deliverability below the pressure bound-
for the development of a sweet spot (Surdam et al., ary (i.e., which are sweet spots);
1995). Because of this, sandstone reservoir systems in • Documentation of the potential determinative ele-
the RMLB are characterized by multiphase fluid- ments that control sweet spot development in the
flow behavior and are dominated by capillarity and targeted lithofacies (e.g., fractures, early migra-
relative permeability; these highly variable sands tion of liquid hydrocarbons, chlorite rims, over-
give up their fluids only when their displacement pressuring, and dissolution of early carbonate
(threshold) pressures are exceeded. Thus, the impor- cement);
tance of reservoir sweet spots becomes apparent; • Detection and delineation of sweet spots using
sweet spots allow the pressure regime and, hence, 2-D and 3-D models of electric log response and
the fluid-flow system in the total reservoir volume seismic data.
A New Paradigm for Gas Exploration in Anomalously Pressured “Tight Gas Sands” 297

ACKNOWLEDGMENTS Wyoming, in Symposium of Late Cretaceous Rocky,


Wyoming and adjacent area: Wyoming Geological
This study was funded through the Gas Research Association 16th Annual Field Conference Guide-
Institute under Contract No. 5091-221-2146. The origi- book, p. 129–137.
nal document was reviewed by Kathy Kirkaldie. Heasler, H.P., and R.C. Surdam, 1992, Pressure com-
Graphic assistance was provided by Allory Deiss. partments in the Mesaverde Formation of the Green
River and Washakie basins, as determined from
drillstem test data, in C. Mullen, ed., Rediscover the
REFERENCES CITED Rockies: Casper, Wyoming, Wyoming Geological
Association 43rd Annual Field Conference Guide-
Chiang, K.K., 1984, The giant Hoadley gas field, South book, p. 207–220.
Central Alberta, in J. Masters, ed., Elmworth—case Iverson, W.P., 1995, Detection and delineation of
study of a deep basin gas field: AAPG Memoir 38, porosity “sweet” zones in Mesaverde Group tight
p. 297–313. gas sands, in Gas reservoir sweet spot detection and
Christiansen, G.E., 1995, Factors influencing differen- delineation in Rocky Mountain Laramide Basins:
tial natural gas production from the upper Creta- Chicago, Gas Research Institute, Topical Report No.
ceous Upper Almond Formation, Wamsutter arch GRI-95/0443, Contract No. 5091-221-2146, p. 31–46.
area, Sweetwater and Carbon counties, Wyoming: Iverson, W.P., and R. Surdam, 1995, Tight gas sand
Master’s thesis, University of Wyoming, Laramie, production from the Almond Formation, Washakie
Wyoming, 157 p. Basin, Wyoming, Paper SPE 29559: Proceedings of
Davis, T.B., 1984, Subsurface pressure profiles in gas- the SPE Rocky Mountain Regional Meeting, Den-
saturated basins, in J.A. Masters, ed., Elmworth— ver, CO, March, 20–22, p. 163–176.
case study of a deep basin gas field: AAPG Memoir Jaworowski, C., R.C. Surdam, G. Christiansen,
38, p. 189–203. M. Grout, H.P. Heasler, W.P. Iverson, R.S. Martin-
Dunn, T.L., W.P. Iverson, B. Aguado, J. Humphreys, sen, M. Olson, and E. Verbeek, 1995, Natural frac-
and R.C. Surdam, 1995, Improvements to reservoir tures and lineaments of the east-central Greater
evaluation and characterization, Almond Forma- Green River Basin: Chicago, Gas Research Institute,
tion, Green River Basin, Wyoming, in An engineer- Annual Report No. GRI-95/0306, Contract No.
ing and geologic evaluation of a horizontal gas well 5091-221-2146, 81 p.
completion in the Almond sandstone-Echo Springs Law, B.E., 1984, Relationships of source-rock, thermal
field, Greater Green River Basin, Wyoming: GRI maturity, and overpressuring to gas generation and
Topical Report GRI-95/0066, p. 31–70. occurrence in low-permeability Upper Cretaceous
Fouch, T.D., T.F. Lawton, D.J. Nichols, W.B. Cashion, and Lower Tertiary rocks, Greater Green River
and W.A. Cobban, 1983, Patterns and timing of syn- Basin, Wyoming, Colorado, and Utah, in J. Wood-
orogenic sedimentation in Upper Cretaceous rocks ward, F.F. Meissner, and J.L. Clayton, eds., Hydro-
of central and northeast Utah, in M. Reynolds and carbon source rocks of the Greater Rocky Mountain
E. Dolly, eds., Mesozoic paleogeography of the Region: Rocky Mountain Association of Geologists,
west-central United States: Rocky Mountain Section p. 469–490.
of SEPM, Rocky Mountain Paleogeography Sympo- Law, B.E., R.M. Pollastro, and C.W. Keighin, 1986,
sium 2, p. 305–336. Geologic characterization of low-permeability gas
García-González, M., D.B. MacGowan, and R.C. Sur- reservoirs in selected wells, Greater Green River
dam, 1993a, Mechanisms of petroleum generation Basin, Wyoming, Colorado, and Utah, in C. Spencer
from coal, as evidenced from petrographic and and R. Mast, eds., Geology of tight gas reservoirs:
geochemical studies: examples from Almond AAPG Studies in Geology, v. 29, p. 253–269.
Formation coals in the Greater Green River Basin, in Leach, W.G., 1994, Distribution of hydrocarbons in
S. Andrew and B. Strook, eds., Wyoming geology– abnormal pressure in south Louisiana, U.S.A., in
past, present and future: Casper, Wyoming, W.H. Fertl, R.E. Chapman, and R.F. Hotz, eds.,
Wyoming Geological Association Jubilee Anniver- Studies in abnormal pressures: developments in
sary Guidebook, p. 311–324. petroleum science, 38: New York, Elsevier,
Garcia-González, M., D.B. MacGowan, and R.C. Sur- p. 391–428.
dam, 1993b, Coal as a source rock of petroleum and Love, J.D., 1970, Cenozoic geology of the Granite
gas—a comparison between natural and artificial Mountains area, Central Wyoming: USGS Profes-
maturation of the Almond Formation coals, Greater sional Paper 495-C, 154 p.
Green River Basin in Wyoming, in D. Howell, ed., Miller, F.X., 1977, Biostratigraphic correlation of the
The future of energy gases: U.S. Geological Survey Mesaverde Group in southwestern Wyoming and
Professional Paper 1570, p. 405–437. northern Colorado, in Exploration frontiers of the
García-González, M., R.C. Surdam, and M.L. Lee, 1997, central and southern Rockies: Rocky Mountain
Generation and expulsion of petroleum and gas from Association of Geologists, p. 117–137.
Almond Formation coal, Greater Green River Basin, Roehler, H.W., 1969, Stratigraphy and oil-shale
Wyoming: AAPG Bulletin, v. 81, no. 1, p. 62–81. deposits of Eocene rocks in the Washakie Basin,
Hale, L.A., 1961, Late Cretaceous (Montanan) stratig- Wyoming: Wyoming Geological Association 21st
raphy, eastern Washakie Basin, Carbon County, Annual Field Conference Guidebook, p. 197–206.
298 Surdam

Roehler, H.W., 1990, Stratigraphy of the Mesaverde Surdam, R.C., Z.S. Jiao, and H.P. Heasler, 1997, Anom-
Group in the central and eastern Greater Green alously pressured gas accumulations in Cretaceous
River Basin, Wyoming, Colorado, and Utah: U.S. rocks of the Laramide Basins of Wyoming: A new
Geological Survey Professional Paper 1508, 52 p. class of hydrocarbon accumulation, in R.C. Surdam,
Surdam, R.C., P. Yin, M. Garcia-González, D.B. Mac- ed., Seals, traps, and the petroleum system: AAPG
Gowan, H.P. Heasler, Z.S. Jiao, C. Jaworowski, R.S. Memoir 67, p. 199–222.
Martinsen, W.P. Iverson, J. Liu, D. Britton, and Winn, R.D., Jr., M.G. Bishop, and P.S. Gardner, 1985,
G. Christiansen, 1993, Natural gas resource charac- Lewis Shale, south-central Wyoming: shelf, delta
terization study of the Mesaverde Group in the front, and turbidite sedimentation: Wyoming Geo-
Greater Green River Basin: Chicago, Gas Research logical Association 36th Annual Field Conference
Institute, Annual Report No. GRI-93/0423, Contract Guidebook, p. 113–130.
No. 5091-221-2146, 452 p. Yin, P., and R.C. Surdam, 1995, Effects of diagenesis on
Surdam, R.C., Z.S. Jiao, and J. Liu, 1995, Anomalous petrophysical properties of reservoir rocks, in
pressure regime in the Washakie Basin, Wyoming: D. Jones, ed., Resources of southwestern Wyoming:
Chicago, Gas Research Institute, Topical Report No. Cheyenne, Wyoming, Wyoming Geological Associa-
GRI-95/0390, Contract No. 5091-221-2146, 24 p. tion 1995 Field Conference Guidebook, p. 247–254.
Reid, H.W., 1997, Evaluating seal facies permeability
and fluid content from drill-stem test data, in R.C.
Surdam, ed., Seals, traps, and the petroleum sys-
tem: AAPG Memoir 67, p. 299–312.

Appendix

Evaluating Seal Facies Permeability and


Fluid Content from Drill-Stem Test Data
H.W. Reid
Hugh W. Reid & Associates, Ltd.
Calgary, Alberta, Canada

ABSTRACT
Many facies assumed to act as seals contain stringers of sand and silt that
are potential low-grade reservoir units. The drill-stem tests (DSTs) from
these formations generally show them to be “tight,” but many of the wells in
these “barrier” facies ultimately become commercial producers after comple-
tion. Calculation of the permeability and fluid content of these facies from
DSTs has not been a common practice because the facies often do not pro-
duce oil or gas when tested, making them unattractive to operators, and
because the analysis of DSTs from tight formations can be problematic.
However, knowing the permeability of seal facies helps operators determine
which of these “barriers” are the leakiest and, hence, are the best potential
exploration targets. In this study, the shape of the shut-in curve on pressure
charts and other subtle indications are used to more accurately assess the
reservoir quality of these neglected formations. This paper will attempt to
demonstrate that a good approximation of the leakage potential of these
facies can be made using published empirical correlations if the permeabili-
ty, as from a DST, is known.

INTRODUCTION Estimation of permeabilities of reservoir-quality


rocks from DSTs are performed routinely, particularly
In many pinch-outs, lateral seals are not perfectly if hydrocarbons were produced during a test. How-
solid shale; stringers of low permeability sand and silt ever, the permeabilities of tighter barrier or seal facies
exist and are often oil stained (Figure 1A). If a well has of traps are rarely computed from DSTs. This is
“missed the sand” and penetrated the barrier, it is not because (1) operators are generally not interested in
likely to be cored. But any small stringer with a show further investigation of unsuccessful tests from dry
is often subjected to drill-stem test (DST) pressure holes and (2) there are many perceived difficulties in
buildup analysis before abandoning, just to be sure it performing DSTs in tight formations where the pres-
does not contain potential pay (Figure 1B); the leakiest sure curves are not sufficiently stabilized for standard
stringer in a barrier controls the trap holding capacity. Horner-type analyses. Several methods of analyzing
For this reason, many DSTs of “tight” barrier facies tight DSTs are currently used by DST service compa-
exist. These data are important because they can be nies and usually involve the use of proprietary soft-
used to determine permeability, fluid content, and ware, but good approximation of permeability/
leakage potential within a barrier facies. thickness (kh) can also be made using simpler “type

299
300 Reid

Kh
f low
in g ers o ilt exist,
Str and
s
ined
sand n oil sta
of t e

K = 0.01 md K = 0.10 md
Pd = 16 psi Pd = 61/2 psi
3 ft.
K = 0.03 md
K = 0.05 md
Pd = 51/2 psi
Pd = 1.3 psi
K = 0.07 md
Pd = 6 psi

DST 1
shale Rec. Gas
1 ft.
f t . shale Cut Mud
1

DST 2
Rec. Oil
Flecked Mud

Thin silts, sands or


conglomerates connect
individual channels

B
Figure 1. (A) Diagram of a typical sandstone pinch-out into a barrier facies;
inset shows interbedded sands in a lateral seal. Pd (displacement pressure) is
not directly proportional to K (permeability). It is also related to pore size dis-
tribution (Burdine, 1950). (B) Diagram showing missed potential pay in a bar-
rier facies. After Hill et al. (1961).

curves” designed specifically for DSTs (Crawford et produced on test. (The hydrocarbon presence may be
al., 1977). inferred from the configuration of the shut-in pres-
Useful information is often extracted from even sure curve and other subtle indications.) In fact,
the tightest DSTs. For example, the presence of resid- many times, a minor quantity of hydrocarbon is
ual nonhydrocarbons in the silty barrier facies, produced from the tight sandstones; thus, the perme-
updip from reservoir pinch-outs, may be detected ability calculated from the DST is actually the perme-
from “mud recovery DSTs,” even if no oil or gas was ability effective to the hydrocarbon phase. Since we
Liquid Tests Typical Gas Tests Typical Gas Rates
Recoveries or air blows
Shut-in pressures Minor mud recoveries
Type A show negligible rise Type A Air blows only
Rec. < 100' mud no gas to surface
Strong
Seal

Extremely Tight Zones Flow curves are Negligible kh


kh ≤ 0.1 md ft virtually horizontal kh ≤ 0.1 md ft

Very slow rise in Minor mud recoveries


Type B shut-in pressures Type B G.T.S. TSTM or
Rec. < 500' liquid <50 Mscf/D
mud, oil or water Fair
to Weak
Seal
Very Low kh Imperceptible rise Very Low kh
kh = 0.1 to 10 md ft in flow pressure kh = 0.1 to 10 md ft

Fairly rapid rise Measurable gas rates


Type C in shut in pressures Type C >500 Mscf/D
but static value Rec. 500' Some mud or condensate
not reached to 1000' liquid recovered

Low to Moderate kh Flow curves show Moderate to good kh


appreciable rise Normal
kh = 1 to 100 md ft kh = 10 to 100 md ft
Reservoir
Shut-in pressures Quality Gas rates over
Type D stabilize almost Type D 2 Mscf/D
immediately Rec. > 1000' liquid Usually only minor liquid
coproduced unless
very wet gas

High to Excellent kh Flow curve rises High kh


kh ≥ 100 md ft rapidly and almost
reaches shut in
pressure value

Figure 2. Model pressure charts showing an example of a liquid and a gas test for each of the following: strong seal rocks, fair to weak seal rocks,
and normal reservoir-quality rocks. Mscf/D= thousand standard cubic feet per day.
Evaluating Seal Facies Permeability and Fluid Content from Drill-Stem Test Data
301
302 Reid

values, in turn, the approximate trap holding capacity


Location #6-23-46-7 W4
DST #2
may be estimated for the nonleaky barriers. These data
INSIDE REC. #8849 have important implications; if a leak-prone seal is
Depth 582.8 known to contain residual oil, this may warrant mov-
ing updip to find the reservoir.
In addition, deep formation damage often leads to a
distortion of the shape of the DST shut-in pressure
curve, so that the DST looks tight, but in reality is not.
In fact, in some so-called tight tests in “barrier” facies,
the wells tested ultimately became commercial pro-
ducers after completion. Preexisting maps of the
“trap” have had to be changed somewhat, such that
the reservoir was enlarged and the “trapping” facies
shrank! Such a situation is not always unrewarding for
the geologist, who does not have a seal, but may have
kh = 0.4 md ft just discovered potential missed pay to redrill. Thus, a
Over 5.5 ft sand ∴K = 0.07 md DST may reveal that the barrier facies is not a seal at all
but a potential low-grade reservoir.
Figure 3. Field example of a weak seal updip from
the Viking/Kinsella gas field, southern Alberta.
RELIABILITY OF PERMEABILITY
MEASUREMENTS DERIVED FROM
DRILL-STEM TESTS
are interested in the potential for hydrocarbon leak-
age, not water leakage, the tight hydrocarbon tests In order to gain a qualitative estimate of permeabil-
provide the most direct tool for this assessment. Once ity/thickness, kh, and hence indirectly the leakage
the permeability effective to the hydrocarbon phase is potential of the facies, and seal capacity, Pd, from DSTs,
known, a range of possible displacement pressures can model pressure charts are shown (Figure 2). An exam-
be estimated from empirical correlations. From these ple of a liquid test and a gas test is shown for each of

A F
B F.S.I.P
I.S.I.P
1910 psi 1765 psi E

F.F.P
350 psi D P1 P2 P3 P4 P5 P6 P7 P8 P9
C
base line
initial shut-in final flow final shutin
initial ∆t1
flow ∆t2
∆t3
∆t4
∆t5
∆t6
∆t7
∆t8
∆t9

Figure 4. Analysis of a DST using the Horner Method; the chart is measured
in detailed pressure/time increments.
Evaluating Seal Facies Permeability and Fluid Content from Drill-Stem Test Data 303

2000 a
1910 Poorly developed SI curves
not suitable for standard
1800 Horner Analyses

1600 b
1536

1400 Straight line portion


Pws

“M” = 1910 - 1536


1200 = 374 psi/cycle

1000
Figure 6. Example of a DST of a tight seal showing
800 how the shut-in curve is not sufficiently built up to
reach the “straight-line” portion needed to compute
kh using the Horner method. SI = shut-in.
600
20 10 8 6 4 2 1
T = ∆t
∆t
seal (type B test), contains gas, and may presently be
Figure 5. Horner Plot constructed from a DST. leaking gas (Figure 3). These results are confirmed by
service company computations of kh = 0.4 md ft. and
>5.5 ft of sand, indicating k = 0.07 md.

The Horner Method


the following: rocks that constitute strong seals (type A
tests), rocks that constitute fair to weak seals (type B The Horner Method can be used to compute a
tests), and rocks of normal reservoir quality (type C quantitative kh value from a DST. The Horner Method
and D tests) (Figure 2). A field example (Figure 3) proceeds as follows:
from a barrier facies updip of the Viking/Kinsella gas
field in southern Alberta, Canada, is also shown; • A chart is read in detailed pressure/time incre-
(Cretaceous Viking Sand) pattern recognition indi- ments (Figure 4).
cates that the field example constitutes a fair to weak • A Horner Plot is constructed (Figure 5).

Core Figure 7. Saturation of area in


kabsolute DST drainage radius.
1.0
0.9
kr, relative permeability

0.8
DST ke = kro x ka
0.7
kro krw
0.6
0.5
0.4
keffective =
0.3
kabsolute x krelative
0.2
0.1 Swi

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Sw, water saturation, fraction
Saturation of area
in DSTdrainage radius
304 Reid

1.0 CHART FROM FIRST TEST ON PENETRATION


(GAS FLOW 45 Mscf/D)
Reduction of K
by an order
of magnitude 0.1 INVASION
with time
kh appears moderate
Relative Permeability

but evidence of damage


& DAMAGE

0.01
BECOMING

CHART FROM SECOND TEST 31 DAYS LATER


0.001 WORSE
(SAME INTERVAL; GAS FLOW T.S.T.M.)

WITH kh appears very low


0.0001 no evidence of damage

TIME

0.00001
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Water Saturation Figure 9. Drill-stem test charts from a Triassic sand
35% Sw after 58% Sw in northeast British Columbia; first test completed
penetration when tested on penetration, second test taken 31 days later.
by drill bit after 1 week

with time. In the chart from the first test on penetra-


Figure 8. Relative permeability change in a low- tion (gas flow 45 Mscf/D) of a Triassic sand in north-
permeability gas zone tested one week after drilling. east British Columbia (well location: 94-G-1-89-D), kh
appears to be moderate, but there is evidence of dam-
age (Figure 9). In the chart from the second test 31
• If a straight line develops, the slope is determined
days later in the same interval (gas flow too small to
and kh (if liquid) is computed as
measure), kh appears very low, and there is no evi-
dence of damage (Figure 9). Thus, a DST is simply a
162.6(q)( µ )(β ) “snapshot in time” of a changing kh, so one should
kh =
m test upon penetration to obtain an accurate kh.

where q = flow rate in s.t.b./d (stock tank barrels per DETERMINING FLUID CONTENT AND
day), µ = fluid viscosity in cp, β = the formation vol-
ume factor in r.b./s.t.b (reservoir barrels/stock tank “SUBTLE” OIL IN SEALS
barrels) and m = slope of the Horner Plot psi/cycle. A Fluid Content
slightly altered formula is used for gas DSTs. This
method is really not suitable for DSTs of “tight” seals Even though drill-stem tests (DSTs) of tight barri-
because the shut-in curve is not sufficiently built up to ers usually only produce mud, we can frequently
reach the “straight-line” portion necessary for this determine if hydrocarbons are present from the
technique (Figure 6). However, DST data can be ana- shape of the shut-in curves in the pressure chart. So
lyzed using Type Curve Analysis (Crawford et al., far we have seen that permeability affects chart
1977). shape, but thus far in all cases normal shut-in curves
There are certain advantages and disadvantages to have an ever-decreasing slope, regardless of the
calculating permeability measurements from DST data effects of permeability (Figure 10). The pressure
rather than from core data. DST values are frequently change (dP/dT) is directly proportional to the flow
less optimistic and more realistic than core-derived rate (q): dP/dT ∝ q (where single-phase fluid is pro-
numbers because a DST-derived permeability is an in- duced). Thus, as the well “slows down” on closing
situ value and is a permeability effective to the fluid the valve, the shut-in pressure slope decelerates.
produced at the saturation that existed around the This ever-decreasing slope (dP/dT) (Figure 11) exists
well bore during the test (i.e., at reservoir saturation) because downhole, after the valve is closed the well
(Figure 7). However, if deep invasion occurs, such as “slows down,” but still flows into the space below
in this example of a low-permeability gas zone tested the packer and closed valve (Figure 12). In most
one week after drilling (Figure 8), the DST-derived shut-in periods, there is ever-decreasing flow into a
permeability may be pessimistic. It is clear from this closed volume.
field example from a Triassic sand in northeastern However, some shut-in curves are not ever-
British Columbia (Figure 9) that kh values can change decreasing slopes. Sometimes a constant slope (type
Evaluating Seal Facies Permeability and Fluid Content from Drill-Stem Test Data 305

The shut-in curves Figure 10. Pressure charts


display ever- showing that usually shut-in
decreasing slope curves have an ever-decreasing
slope.

Permeability Type

Excellent

FSI 2nd ISI


1st flow
flow

Good

FSI 2nd ISI


1st flow
flow

Relatively low

FSI 2nd ISI


1st flow
flow

Very poor

FSI 2nd ISI


1st flow
flow

A curve) will result from severe shallow damage or “Subtle” Oil in Seals
deep formation damage (Figure 13). An S-shaped
curve results either when two phases are present in An S curve indicates that oil is potentially present
the formation or in the case of gas zone cleanup in the reservoir because of the pore geometry and
(Figure 13). Thus, even if only mud is produced by relative permeability problems encountered during
the DSTs, the shape of the shut-in pressure curve testing. The early acceleration and subsequent decel-
can often reveal whether hydrocarbons are present eration of the S shape result from gas blocking pore
in the rock. throat spaces and then going back to solution, which
,,

yy
z|{{
306 Reid

,,

zyy
|{{
TOOL SCHEMATIC
Ever-decreasing slope


|{{
,,
zyy
Shut-in valve closed

,,

zyy
|{{
Figure 11. Field example shut-in curve with an ever-
decreasing slope. Main valve open

Bypass valve closed


distorts the shut-in curve. More specifically, an S


|{{
,,
zyy
curve develops as follows (Figure 14). Before the tool
is opened, mud pressure in the annulus exceeds the
saturation pressure of the oil, so all of the gas
remains dissolved in the oil. After the tool is opened,
the reservoir is exposed to the atmosphere, so gas
breaks out suddenly in the oil, in the pores, and in

,,

zyy
|{{
the pore throats. In severe cases, if gas breaks out in
narrow pore throats, it breaks the connecting oil fila- Sump
In most shut-in periods
ment, blocking the oil flow (“gas block”) and causing Volume
there is ever-decreasing
relative permeability problems. Gas flows preferen- (Rat hole
flow into a closed volume
tially, but often at a negligible rate. Finally, during plus internal
the shut-in periods, as the pressure rises over the tool volume)
Pressure value
bubble point, most gas is redissolved or is in a very
compressed state, and the shut-in curve resumes its
usual deceleration. The S-curve DSTs are common in
tight barrier facies because the S shape is more pro-
nounced in lower kh rock. Figure 15 shows a range of
S curves; the S shape becomes more pronounced as
Figure 12. Tool schematic showing how in most
permeability decreases.
shut-in periods, there is ever-decreasing flow into a
closed volume.
ONE MAN’S SEAL MAY BE ANOTHER
MAN’S RESERVOIR! (CASE HISTORIES) The cumulative production of this well over 17
years was 9985 s.t.b. oil, 2850 s.t.b. water, and 41.5
Case 1: Economics (Northern Alberta) Mscf of gas. While this well is not economically viable
in Alberta, it might be commercially productive as a
There are two cases in which a barrier updip may stripper well in the eastern United States (depending
actually contain a producible reservoir (Almon and on the oil location and the price of oil). These findings
Reid, 1990). The first case is when low-grade silt/ are typical of many such tests of poor-quality stringers
sand developments in barriers updip often can pro- in barriers that can be considered reservoir rocks. If we
duce marginal oil wells that are economic if the take into account that the average U.S. oil well rate is
price of oil rises. Figure 16 is a typical, very poor only 14 bbl/d and the average Canadian oil well rate
looking DST from northern Alberta with develop- is 40 bbl/d (Society of Petroleum Engineers, 1988), the
ment of only one part of the S curve. Despite disap- importance of such stringers is evident.
pointing recovery, the subtle S shape indicated that
oil could be present in this barrier. In order to test Case 2: Cinderella Story (North Dakota)
this hypothesis, the operator stimulated the well in
The second case in which a barrier acts as a reser-
the following manner:
voir rock occurs when a high kh zone is mistakenly
perceived to be tight, but actually contains commer-
• Treatment: 500 gal acid; fractured 15,000 lb sand cially viable reservoirs. The DST from this zone (Fig-
gel & XW ure 17) is similar to that of Case 1, and has a poorly
• Initial production: 33 bbl/d (barrels of oil per day) developed S curve. The Horner Plot also indicates
Evaluating Seal Facies Permeability and Fluid Content from Drill-Stem Test Data 307

Figure 13. Shut-in


curves that do not have
ever-decreasing slopes.
Causes The type A curve has a con-
Type A stant slope and results from
Severe 'shallow' severe shallow damage or
Constant slopes
damage or deep formation damage.
'deep' damage The type B curve has an S-
shaped slope and results
from the presence of two
phases in the reservoir or in
Ever decreasing the case of gas zone
slope cleanup.

Causes
Type B
Two phases
'S' shaped slopes
present or Gas
zone cleanup
(usually oil)

Ever-increasing
slope

that the formation is tight and not of reservoir qual- initial production of 408 bbl/d (85% oil) was clearly
ity (Figure 18). However, sample examination shows economic. Based on this and similar examples, it is
that the unit has big pores and very small pore advisable when dealing with a tight S curve on a
throats, conditions that are ideal for gas block (Figure DST to look at the pore geometry before assuming
19). In addition, the Pitman diagram (Figure 20) shows the reservoir has been missed, and to consider
isolated large pores connected by tiny micropores. using a gas cushion above saturation pressure on a
While many pore throats are only 0.02 µ and represent retest to prevent gas block in the oil in the pore
micropore throats in the matrix, some pore throats are space immediately surrounding the well bore.
0.25–0.5 µ and represent big, intergranular throats
(Figure 21). The ratio of microgranular to intragranu-
lar pores is 1:1 in the unit matrix (Figure 20). CAUTIONARY NOTES
Since the DST indicates the unit is very tight, but
we know it has high permeability, some other factor Not all DST charts displaying an S curve are
must account for the overly pessimistic DST values. indicative of oil. Some can be caused by residual,
It is likely that this discrepancy has been caused by nonproducible oil in the tested unit (Figure 24). The
gas block; work by Exxon shows that mobile gas sat- saturation of these units is such that water would be
uration in a case like this is 10%–20% (R.M. McKin- produced if they were completed (in 10% of the
ley, 1990, personal communication). In fact, the cases, e.g., transition zones). However, these tests
relative permeability diagram (Figure 22) indicates indicate that operators are not far from oil; they
that the permeability and deliverability values fore- should move downdip from a waste zone and updip
cast by standard DST computations are two to five from a transition zone in order to locate the oil reser-
times less than the actual values due to gas block voir. Other tight S-curve DST charts are actually
alone (physical well-bore damage can increase this caused by the presence of gas, not oil, in the reser-
discrepancy as well). voir (Figure 25). Gas-bearing zones can be differenti-
The difficulty at this point becomes how to stop ated from oil-bearing zones by observing gas rates.
gas breakout when retesting the well. In order to If a gas zone is deeply damaged and is “cleaning
accomplish this in the example, the flowing pres- up,” the rates are increasing; as the well is shut in,
sure was kept above bubble point by utilizing nitro- the pressure change will accelerate also, since dP/dT
gen as a cushion. The DST chart of the retest (Figure ∝ q, yielding the accelerating part of the S shape
23) indicates high kh, and the recovery was 713 ft shut-in curve.
(217 m) of highly gas-cut oil and 360 ft (110 m) of We can learn about the trapping and the leakage
gas-cut saltwater in the unit. The Horner Plot shows potential of the barrier indirectly from a DST. For
that kh is 45 md ft. The well was completed, and the example, if we have a DST of a leaky stringer updip
308 Reid

Rel K Diagram
100
Pore Rock grain

10
Above Ps So = 0.75
∴ Gas Dissolved Kro

Kr
Kro = 0.6
1.0

0.1
1 So 75% 100

Resumed deceleration

Ps saturation pressure DST chart


Early acceleration

Pre- ISI 2nd flow FSI


flow Result: GTS - TSTM
Rec. OCM

100

10
Below Ps So = 0.20 Krg
∴ Solution Gas
Kr

Kro = 0.07 Kro


Breakout 1.0
but
Sg = 0.8
Krg = 0.7
0.1
125% S 100
(75% gas) o
100 Sg 1

Figure 14. Diagram showing the development of an S curve in the case of a test of an oil zone with a pore
geometry of big pores and small throats.

of a pinch-out and we know k, we can estimate the (Petroleum Research Corp. A-5, 1959). Using the
displacement pressure and from this value estimate chart shown in Figure 27, we enter k as 0.4, then Pd =
how much the downdip hydrocarbon height, Z, is in 0.25–1.5 psi. Next, we find the subsurface oil and
the reservoir (Figure 26). In order to find the dis- water gradients, which in this example are 51° API
placement pressure using a DST in a leaky stringer oil at 4000 psi and 160°F = 0.246 psi/ft, and 200,000
updip of a “sure shot” pinch-out, an estimate, n, of mg/L water at 4000 psi and 160°F = 0.481 psi/ft
the pore-size distribution index is made (Burdine et (Petroleum Research Corp., 1960, figures 3,4).
al., 1950). Since in this example the lithology is a Finally, we calculate the maximum oil column (Zmax)
silty, dirty sand, n will vary between 2.25 and 3.5 that can be supported under these conditions,
Evaluating Seal Facies Permeability and Fluid Content from Drill-Stem Test Data 309

'S' shape more


pronounced as
permeability Subtle 'S'
decreases (good permeability)

Broad
sweeping 'S'
(lower
permeability)

Only part of 'S'


developed
(apparent poor
permeability)

Figure 15. Diagram relating a range of S curves to permeability.

accounting for any variations in interfacial tension, This prospect is unlikely to hold an economic oil
which in this case is column unless the hydrodynamics are favorable,
Pd 1.5 i.e., there is strong downdip flow (e.g., as in the
Zmax = = = 7 ft Joffre oil field, in the Viking Sand, Alberta), or the
∇ w – ∇ o (0.481 – 0.246) leaky stringer is not continuous to the updip edge of
the reservoir. (The chart assumes an oil/water inter-
where Zmax = maximum oil column (ft), Pd = dis- facial tension of 10 dynes/cm; since this could vary
placement pressure (psi), ∇w = subsurface water gra- from 3 times (30) to 0.5 times (5.0), P d and Z also
dient (psi/ft), and ∇0 = subsurface oil gradient (psi/ft). vary directly.)
310 Reid

5168 5172

3520

1303

60 88

REC 1438' GAS


120' GAS AND WATER CUT MUD
SAMPLE CHAMBERS CONTENTS 1100 cc OIL, 1220 cc MUD,
DST: Rec 55 ft mud. VWIP dead after 15 min. 1.5 cu ft GAS 525 psi

Figure 16. A typical, poor DST with development of


only one part of the S curve from a marginally pro- Figure 17. Another DST with a poorly developed S
ductive unit. VWIP = very weak initial puff. curve from a commercially viable reservoir rock.
Arrows indicate early accelerating and late deceler-
ating shut-in curve displaying the S shape.

3500
direction of flow
Horner Plot
3000
Too tight for straight
line to be developed Water Gas Water
2500
large capillary force,
Pressure (psi)

small pore radius


2000 small capillary force,
large pore radius

1500 Figure 19. Schematic diagram illustrating large


Final pores and small pore throats; these are ideal condi-
tions for gas block.
1000
Initial

500 REFERENCES CITED


Almon, W. R., and H.W. Reid, 1990, Geology and DST
analysis: an integrated approach to identify forma-
0
10 7 5 4 3 2 1 tion damage and bypassed pay using pore geome-
try, clay mineralogy, and pressure behaviour:
Calgary, Canada, Canadian Society of Petroleum
Time (T + ∆t)/∆t Geologists, v. 17, p. 1–2.
Burdine, N., L.S. Gournay, and P.P. Reichertz, 1950,
Figure 18. Horner Plot of the DST shown in Figure 17. Pore size distribution of petroleum reservoir rocks:
AIME Petroleum Br. Trans., v. 189, T.P. 2893,
p. 195–204.
Crawford, G.E., A.E. Pierce, and R.M. McKinley, 1977,
Evaluating Seal Facies Permeability and Fluid Content from Drill-Stem Test Data 311

Intergranular
50.0

Pore Throat Size Distribution


40.0

Percent pore space


30.0 29.3

20.0 19.1

11.3
10.0 9.12
5.35 5.88 5.90
2.51 2.52 2.51
1.00 2.03 0.83 0.32 1.00
0.0
0.00 .021 .024 .027 .030 .035 .042 .053 .061 .071 .085 .107 0.14 0.21 0.35 0.52
Pore throat radius, microns

Figure 21. Pore-throat size distribution of the forma-


tion tested.
Microgranular Intragranular

Figure 20. Pitman diagram for the rock tested by the


DST shown in Figure 17. DST Chart, Retest

5168 5188
1.0 485
preflow
Krg second flow
0.8 high kh
Productivity 2860 2930 ISI buildup 2923
Increase by
0.6 0.6 1.6 cushion press 1834
Krv
Krg

above bubble point 1565


0.4 Kro to
cushion
bleed-off 440
0.2 ~0.2 5
times RECOV: 713' HIGHLY GAS CUT OIL
(if Gas Block 360' GAS CUT SALT WATER
0 removed)
0 20 40 60 80 90 100 HORNER PLOT SHOWS kh = 45 md ft
Liquid saturation (Sw + So)
Figure 23. Drill-stem test chart of the retest, indicat-
ing that 713 ft (217 m) of recoverable, highly gas-cut
Figure 22. Relative permeability diagram for the zone oil and 360 ft (110 m) of gas-cut saltwater are present
tested, indicating that the permeability and deliver- in the unit. The Horner Plot shows kh = 45 md ft.
ability values predicted by standard DST computa-
tions are two to five times less than the actual values.

Type curves for McKinley analysis of drill-stem text Petroleum Research Corp., 1959, Reservoir pinch-
data: SPE paper 6754 presented at 52nd Annual Fall outs—sieves or seals?: Denver, Colorado, Researach
Conference of SPE of AIME, Denver, Colorado, Oct. Report A-5 (unpublished), September, 58 p.
9–12. Petroleum Research Corp., 1960, Geologic significance
Hill, V.G., W.A. Colburn, and J.W. Knight, 1961, of pressure gradients, capillarity, and relative per-
Reducing oil finding costs by use of hydrodynamic meability: Denver, Colorado, Research Report A-0
evaluations, in Economics of petroleum explo- (unpublished), 60 p.
ration, development, and property evaluation: New Society of Petroleum Engineers, 1988, Energy focus:
Jersey, Prentice Hall, p. 38–69. Journal of Petroleum Technology, v. 40, no. 2, p. 168.
312 Reid

'S' SHAPED CURVE OF WATER ZONE 'S' SHAPED CURVE OF DAMAGED GAS ZONES
(RESIDUAL OIL) – 'CLEANING UP'

Recovery: 64 m mud; no gas to surface


DST Flow Rate: 12 bbl/d (1.9 m3/d)
Post-Completion Rate: Swabbed 'dead' SW @ 40 bbl/d (6.4 m3/d) Recovery: 19 m mud
DST Flow Rate: Gas flow rate gradually increasing
to 25 Mscf/D (700 m3/d)
Figure 24. S-shaped curve of water zone caused by Post-Completion Rate: 1.7 mmscf/d (47x103 m3/d)
residual, nonproducible oil in the tested unit. after 5000 gal. acid

y,zyyy ,
y
Figure 25. S-shaped curve caused by gas in the reser-

,


y
z
{
voir, not oil (i.e., damaged gas zone cleaning up).


,,,
zzz

z
,


y
{
| ,



y
z
{
| |
|,
z{
y

z

y
, y
,


z
{
|
50.0


z


zzz
||| ,,,
yyy


|||
{{{
,



y
z
{
| ,



y
z
{
| ,



y
z
{
|
,



y
z
{
| ,

y
{
Pd
10.0
n=
10

,


y
z
|
hydrocarbon 5.0
Zh n=

,



y
z
{
|
height n=7
Pd


||| 
{{{
6
n=
buoyancy 5

|{ |
n=
4
n=
Pd 3.5
Z= 1.0 n=
3

0.5
n=
2.5

n=
2.2
Figure 26. Forces controlling hydrocarbon height 5
in pinchout-static (not hydrodynamic) conditions.
0.05 0.1 0.5 1.0 5.0
K (md)

Figure 27. Chart showing the relationship between


displacement pressure, Pd, and rock permeability,
K, for various values of n, the pore-size distribu-
tion index.

You might also like