You are on page 1of 468

Copyrig ht © 2006

By the American Association of Perroleum Cc~>logists (AAPG) and SEPM (Society fo r Sedimentary Geology)
All Rights Reserved

!SllN: 978-0-89181-704-8
0-89181-704-2

AAPC a nd SEPM grant permission for a single pho tocopy of an item fro m this publicatio n fo r personal use .
Authorization for add itio nal copies o f items from this publication for personal or internal use is g ran ted
by AAJ>G and SEPM provided that the base fee of $3.50 per copy and $.50 per page is paid d irectly to the
Co pyrig ht Cleara nce Center, 222 Ros<>wood Drive, Danvers, Massachllsetts 01923 (phone: 978/750-8400).
Fees are subject to change. Any form of electron.ic or d igital sca n.niJ>g or ot her d igital transformatio n of
portions ot' t·his publication i.nto computer-readable M>d/or t·ra nsm ittable form for persoMI or corporate use
re<.1 uires specia l permission f rom, and is subject to fee charges by, the 1\APG ru\d SEPM.

AAJ'>G Editor: Ernest A. Mancini SEJ>M Special P ublicatio n Ed itors:


Geoscience Director: )ames B. Blankenship Laura Crossey and Donald Me~'leill
SEPM Publications Coordinator: Robert Clarke

Tf\is publicat io t\ is available fro m:

The AAPG Bookstore SEPM


P.O. Box 9·79 6128 E.1St 381h St«->et, Suite 308
T,~so. OK U.S.A. 74101-0979 Tulsa, OK U.S.A. 74135-5814
Phone: l -918-584-2555 Phone: l-&J0-865-9765 (North America)
or '1-1:!00-364-AAPG (U.S.A. only) or ·t-9'1 8-610-3361
Fax: 1-918-260-2652 Fax: 1-918-62H685
o r 1-800-898-2274 (U.S.A. only) E ~ 1na iJ: eeUisW'sepm.org
E-mail: bookt>tore®aapg.org www.sepm.org
w-..\1"\v,aapg.org

Ceolosic.ll Society Publi•hiog 1-toU>e


Unit 7, Brnssmill Enterpl'ise Centre
BrassmiU Lane, Bath BA13JN
U.K.
Phone: +44-1225-445046
Fax: +44-'1 225-442836
E~roai J : $ale$@geolwc.org.uk
W\\11\'.geolsoc.org. tr.k

The American Associatio n of Petroleum Geologists (AAPG) does no t endorse or recommend products or
services that may be cited_, used, o r discussed in AAPG publications o r in presentations a t e vents associated
with AAJ'>C.
Table of Con tents
Acknowledgments ······································ ~~-~~--..!<.

About the Edjtors vj

Introductj on

Giant Hydrocarbon Reservoirs of the World: From Rocks to


Reservoir Characterization and Modeling
P. M. Harris and L. [. Weber

Chapterl ................................. ··----------------------------L----


Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup, Precaspian
Basin, Kazakhstan: Depositional Evolution and Reservoir Development
[ . A . M. Kenter, P. M. Harris, [ . F. Collins, L. [. Webe1; G. Kuanyslwva, and D . [. Fischer

Chapter 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
Facies and Reservoir-quality Variations in the Late Visean to Bashkirian Outer Platform,
Rim, and Flank of the Tengiz Buildup, Precaspian Basin, Kazakhstan
[.F. Collins,[. A.M. Kenter, P.M. Harris, G. Kuanysl1eva, D . J. Fischer, and K. L. Stef(en

Chapter 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Ghawar Arab-D Reservoir: Widespread Porosity In Shoaling-upward
Carbonate Cycles, Saudi Arabia
Robert P. Lindsay, Dave L. Cantrell, Geraint 111. Hugl1es, Thomas H. Keith,
Harry W. Mueller Ill, and S. Duffy f<usse/1

Chapter 4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
High-resolution Sequence Stratigraphy and Reservoir Characterization of Upper Tha mama
(Lower Cretaceous) Reservoirs of a Giant Abu Dhabi Oil Field, Un ited Arab Emi rates
Christian[. Stmhmeuge1; L.jim Weber, Ahmed Ghani, Klwlil AI-Mehsin, Omar Al-jeelani,
Abdulla AI-Mansoori, l'aha AI-Dayyani, Lee Vaughan, Sameer A . Khan, and
john C. Mitchell

Chapter 5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir,
Lower Cretaceous, Abu Dhabi (Un ited Arab Em irates)
Lyndon A . Yose, Amy S. Ru(, Christian f. Strohmenger, jim S. Schuelke, Andy Combos,
Ismail AI-Hosani, Slwmsa AI-Maskary, Gerald Bloch, Yousuf 111-Mehairi, and
Imelda G. jolmson
Ch apter 6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
Sequence Stratigraphy and Reservoir Architecture of the Burgan and Mauddud Formations
(lower Cretaceous). Kuwait
Christian f. Strolmrenger; Pemrv E. Patterson, Glraida AI-Salrlan, John C. Mitchell,
Howard R. Feldman, Timothy M. Demko, E<obert W. Wellner, Patrick f. Lelrmann,
G. Glen McCrimmon, E<obett H( Broom/rail, and Neama ;1/.J\jmi

Ch apter 7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
The Sequence Stratigraphy of the Maastrichtian (Upper Cretaceous) Reservoir at Wafra Field,
Partitioned Neutral Zone, Saudi Arabia and Kuwait: Key to Reservoir Modeling and Assessment
Dennis 111. Dull, /laynumd A. Garber, and 111. Scott Meddaugh

Ch apter 8. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Stratigraphic Organization and Predictability of Mixed Coarse- and Fine-grained Lithofacies
Successions in a Lower Miocene Deep-water Slope-channel System, Angola Block IS
M. L. Porter, II. II. G. Sprague, M. D . Sul/ivmr, D. C. Jennette, II. T. Beaubouet T. R. Gar{ield,
C. Rossen, D. K. Sicktt(oose, G. N./ensen, S. /. Friedmann, and D. C. Mohrig

Ch apter 9. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Multiscale Geologic an d Petrophysical Modeling of the Giant Hugoton Gas field (Permian),
Kansas and Oklahoma. U.S.A.
Martin K. Dubois, 111/ttn P. BJ•rrres. Geoffrey C. Bo/1/irrg, and john H. Dovet011

Ch apter 10. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355


Key Role of Outcrops and Core.~ In Carbonate Reservoir Characterization and Modeling, Lower
Permian Ful.lerton Field. Perm ian Basi.n. United States
Step/ten C. Ruppel cmd Rebecca H. Jones

Chapter 11. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395


Cart>onate Sequence Stratigrap hy and Petroleum Geology of the j urassic Deep Panuke Field,
Offshore Nova Scotia. Canada
John A. W Weissettberger, Richard 11. Wierzbicki, and Nancy f . Harland

Chapter 12. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433


Sedi mento logy, Sequence Stratigraphy, and Reservoir Ao:chltecture o f the Eocene 1-Hrador
Formation, Cupiagua Field, Llanos Foothills, Colombia
Juan Carlos Ramon cmd Anclres Fajardo

tv
Acknowledgments
AAPG and SEPM
are grateful and appreciative
for the support of the AAPG Founda t io n .

The follow ing are gratefully


acknowledged fo r their financial support:

Chevron
EJf{onMobil

Contributions are applied against the production


costs o f publicatio n, thus directly reducing the hoo k's
purchase price and ma king the volu me
available to a greater a udience.

v
About the Editors
Paul M. (Mitch) _f:l.a rris is n. Carbonate Reservoir Consultan t with Chevron Energy
T('chnolog:y Compa ny in San Ramon, Califo rnia. He performs <·a rbona te research,
1echnical support projects, consulting, and train ing for the various operating u nils
of Chevron. His \\10tk during t he last~ years has centered o n facies-re lated, strati-
graphic. and diagenetic probl(;'ms th at p€'rtain lO carbom:ne reservoirs and expiO·
ra tion play,o; in most carbo nate basins ,vorldwide.
Mitch rt-ceived h is B.S. and M.S. degre,o.s from West Virginia University a nd Ph.D.
from the University o f l\ttiami, Florida. He has published nu merous pa pers, edited sev-
eral books, a nd is active In AAI'G a nd SEI'I\1. lie has been a Dlstlng11ished Lecturer and
Inte rnational Distinguished Lecturer fo r AAPG. was awarded 11onorary lvlembershlp
from SEPM, and has received the Wallace E. Pratt Memorial Award and Robert H. Dott
Sr. Memorial Award fro m AAPG. Mitch is also ad junct faculty at Rlce University, the
University o f Miami, and the University o f Southern California .

L. J. Qim) W eber works for Ex.• onMobil ExploratJon Company as a Catbonate


Stratigraphy Expert specializing in seque nce/seismic stratigraphy and reservoir chmac·
tedzalion. liis curfent wo rk assignment invoJves techn ica l leadership an d pro ject
toordination as a stratigraphic advisor. J im h<as been employed in th(' oil industry fo r
more than 17 years, working for both heritage Exxon and ~·fobil in production, d eveJ.
o prnent, exploration, and research o rganizations. Pas t· work assignmen ts include SE
Asia, North a nd South Am(' rica, th (' Middle E.:1st, a nd th~ North Caspian rl'gion.
Jim received a bachelo r 1s degree in geology fro m DePau w Un iversity, a master's
d egree from the New Mexico Institute or Mini ng a nd Technology, aud his Ph.D.
from th e University of Ten nessee. He is ac tive in various geological socie ties includ-
in g SEPM and AAPG. He has published vario us papers and abstrac ts ranging from
t he Cambro.Ordovicia n to th e Holocene.

vi
_,. M., and 1.. J. Wehc:r, 2006, Giant hydrocarbon re..'icrvoirs of
Harris, C
the world: from rods to reservoir e.hara<.'1CrizatiOH and modeling,. iu
P. M. Harris and 1.. J, Weber. eds., Giant hydrocarbon reservoirs of

Introduction the world: From rocl-:s to reservoir <:haracterlz.atlon and mo<lellng;


AAPCi Memoir 88/ Sti1M Special PublicatiOn, p. 1- 6.

Giant Hydrocarbon Reservoirs of


the World: From Rocks to Reservoir
Characterization and Modeling
P.M. Harris
Chevron E11ergy Tectmology Company, Sm1 Ramon, California, U.S.A.

L. J. Weber
fuonMo/Jil /Jxploration Company, Houston, Texas, U.S.A.

INTRODUCTION It is our goal that the technical content o f the chapters


presented here result in discussion d irected toward
The SEPM/AAPG core workshop "Giant Hydroca r·
bon Reservo irs o f the World: From Rocks to Reservoir • fundamental concepts and methods of reservo ir
Characterization and Modeling" and th is co mpan ion characterization, which include accurate represen·
publication are an attempt to assemble information tation of internal and external reservoir geology,
on giant (>500 MOEB recoverable reserves) reservoirs precise and quantitative description, and level of
that is of value to a wide audience. Various examples detail that satisfies simulation capabilities
and methods of reservoir characterization, develop· • an understanding of the scales of rock hetero·
ment, and modeling practices are documented in this geneity and its effect o n petrophysical and engi·
volume. Although far from exhaustive, this compila· neering properties and their relationships to fluid
lion includes a wide range of reservoirs when exam· flow and hydrocarbon recovery
ined from any perspective, i.e., location, geology, pro· • improved methods o f reservoir description an.d de-
duction history, and characterization. Figure 1 shows velopment through application of high-resolution
the geographic distribution of the reservoir examples sequence and seismic stratigraphy and seismic vi-
that are Included In this volume. sualization techniques
Agood understanding of geologic variability in time • identifying a lternative approache.~ to mo re e ffec-
and space is a prereq uisite for any successful reservoir tive reservoi r management practices
description. Geologic and engineering data obtlli ned
from core are fundamental buildi.ng blocks in reservoir The reservoir examples described in this volume
characterization. A com mo n goal is to describe the (!) explore historical and alternative approaches to
reservoir in sufficient detail to identify remain ing hy· reservoir description, characterization, and manage·
drocarbons and then to produce these reserve_~ effi. ment and (2) examine appropriate levels and timing
ciently. To this end, a focus in this volume centers on of data gathering, technology applications, evalua·
aspects of geologic model ing as they relate to het· tion techniques_, risk assessment, and nt.a nagement
erogeneity In facies, which typic.a lly controls vari· practices In various stages In the life of Individua l
ability In porosiry, permeability, and fluid saturations. development projects.

COp)'rigllt 0 2006 by 1'hc American Associat:iou of PNrolcum CicoJogim.


001:10. 1306/1 21S872M8824?

1
2 I liarris and Weber

FIGURE 1 . Reservoirs included


in Giant Hydrocarbon Reser-
voirs of the World. Caspian
Basin: (1) Tengiz platform and
(2) Tenglz rim; Middle East:
(3) Ghawar Arab-D, (4) Abu
Dhabi Thamama, (5) Abu Dhabi
Shualba, (6) Kuwait Burgan -
Mauddud, and (7} Wafra
Maastrichtian; West Africa:
(8) Angola Block 15; North
America: ( 9) Hugoton,
(1 0) Fullerton, and (11) Deep
Panuke; South America:
(12) Cuplagua.

The giant reservoirs of th is volume account for the B<lShkirian . Visean and Serp ukhovian cycles are
approximately 0.5 trillion bbl recoverable hydrocar- genera lly easy to correlate fro m well to we.ll over sev-
bons. Reservo ir examples a re both ca rbonate and si- eral kilometers distance, whereas Bash kirian cycles
liciclastic, and collectively, they accoun t for a wide are incomplete and are more difficu lt to correlate.
range of variability in reservoir parameters (e.g., gross The distribution of reservoir rock types in the central
rock volume, net-to-gross, porosity, and permeabi i· platform is determined by buria l diagenetic modifl·
ity). Most of the reservoirs in this volu me occur o ut· cation of an earlier reservoir system that includes
side the United States and Canada, a nd until now, meteoric alteration and po rosity enhancement be-
core from many of these reservoirs have not been low major sequence boundaries and reduced disso-
widely observed . lu tion along higher order sequence boundaries asso-
Enhanced recovery of hydrocarbons requires a criti· ciated with the presence of volcanic ash. The lateral
cal understanding of reservoir heterogeneity by both continuity of tight layers at sequence boundaries prob-
geoscientists and engineers. Spatial heterogeneity that ably grea tly affected later fluid flow as well as t he
affects fluid now occurs over a ll scales of investiga- ultimate distribution o f cementS, dissolution, and bi·
tion from the intr<!well, in terwell, reservoir, and ba- lumen in the central pl<ltfo rm reservoir. The buri<Jl
sin scales of exam ination. We feel that the giant fields diagenetic overprint includes two majo r phases of
discussed herein address issues important to reservo ir reservo ir modification: (1) a corrosion and cementa-
description, characterizatlon, and manage ment from tion phase signillcantly enhanced existing matrix po-
both geologic and engineering perspeclives. rosity In the interior central platform while reduci ng
porosity in the exterior central and outer platform by
SUMMARY OF RESERVOIR EXAMPLES pore-filling equant calcite cement, and (2) bitumen
emplacement and associated corrosion.
Caspian Basin Authors Collins, Kenter, Harris, Kuanysheva, Fischer,
Kenter, Harris, Collins, Weber, Kuanysheva, and and Steffen examine the outer platform, rim and fla nk
fischer discuss the t:entral pla tfo rm part oftheTengiz. of the Tengi z field in " Facies and Reservoir Qual-
field in "L<!te Visean to Bashkirian Platform Cyclicity ity Vari<ltions in th e L;ote Visean to Bashkiri;m Out-
in the Centra l Tengiz Buildup (Precaspian Basin): er Platform, Rim, <lnd Fl<mk of the Tengiz Buildup,
Depositional Evolution and Reservoir Development." Precaspian Basin, Kazakhstan." Platform backstep-
The Tengiz buildup, an inten sely cored and studied ping fro m Tournaisian through late Visean resulted
iso lated ca rbonate platform in the Precaspian Basin, in approximately 800 m (2624 ft) o f bathymetric
contains a succession of shallow-water deposits ra ng- relief. This topography was enve loped by as much as
ing fro m Famenn ian to Bashkirian in age. The upper 2 km (1.2 mi) of Serpukhovian progradation, which
Visean, Serpukhovian, a nd Bashkirian form the main formed a depositiona l wedge a round the o lder plat-
h ydrocarbon-bearing interval in the platform. Depo- forms referred to as the Serpukhovian rim a nd flank.
sitio nal q•cles (high-frequency sequences) in this in· Lower-slope facies include mudstone, volcanic ash,
terval arc several to tens of meters thick fo r the Visean and platform-derived skeletal packstone to grainstone
and Serpukhovian, and decimeter to meter scale for interbedded with boundstone breccia; m iddle-slope
Introduct ion I 3

facies are poorly bedded to massive boundstone brec- mama Group) o f Abu Dhabi in wb icb important hy-
cia with subtypes based on clast co mposition, size, droca.rbon accumu lations occur in p latform carbon-
and packing; and upper-slope facies con sist of in-situ ates. The Kharaib and Lower Shuaiba form ations
microbial bo undstone. Facies of the outer platform contain th ree reservoir units separated by low-porosity
are shallow-platform skeletal, coated-grain, and ooid and permeability dense zones. Core and well-Jog data
packsto ne to grai n.stone. Periodic large-sca le failure fro m a gia nt oil field in Abu Dhabi and outcrop data
of the rim during both the Serpukhovian and Bash· fro m Wadi Rah abah in Ras Al-Khaimah were used to
kiria n resulted in a high degree of lateral facies dis- establish a seq uence-s tratigraphic framework and a
continuity. Solution-en larged fractures, large vugs, litl1ofacies scheme. The Lower a nd Upper Kharaib
and lost circlllation zones produced mainly during late Reservoir Units, as well as the upper dense zone, are
diagenesis form a high-permeability, well-connected part of a late transgressive sequence set of a scc.ond-
reservo ir in the rim and flank. order supersequence, built by two third-ord er com-
posite sequences. The overlying Lower Shuaiba Res-
Middle East ervoir Unit belongs to the la te transgressive sequence
Lindsay, Cantrell, Hughes, Keith, Mueller, and Rus- set and the early highstand sequence set o f this second-
sell describe Ghawar field, which is the world's largest order supersequence a nd comprises one third-order
and most prolific o il field, in "Ghawar Arab-0 Res- composite sequence. The th ree third-o rder compos-
ervoir: Widespread Porosity in Shoaling-upwa rd Car- ite sequences a re composed o f fourth-ord er parase-
bonate Cycles, Saudi Arabia." Production occurs fro m quence sets that show p redo minantly aggradational
the j urassic Arab-D carbonates. The upper half of the and progradational stacking patterns, typical of green-
reservoir is dominated by exceptiona lly high reser- house cycle.~. Reservoir lithofacies range from lower-
voir quality, and t he lower half conta ins interbeds ramp to shoal crest to near backshoal open-platform
of less porosity. Th e reservoir is com posed o f two deposits, whereas nonreservoir (dense) lithofacies rep-
composite sequences. The upper composite sequence resent an inner-ramp, restricted shallow lagoonal set-
boundary is the top o f Arab-D carbonate, which is ting. Integration of subsurface and o utcrop data leads
locally characterized by colla pse breccia. The lower to more insightful and realistic geological models of
composite sequence boundary between tbe Arab-D the subsurface stratigraphy, and the geological model
and the underlying j ubaila is m arked by deeper water realizations based on core, outcrop, well-log, and seis-
cycles occu rring over gra in-dom inated cycles. Several mic data co nstrain now-simulation models.
h igh-frequency sequences are each composed o f cy- Yose, Ruf, Strohmenger, Sch uelke, Gombos, AI-
cle sets that each contai n approximately five individ- Hosani, AI Maskary, Bloch, Al-Mehairi, a nd Johnson
ual carbonate cycles. These carbonates formed upon integrate high-resolution three-dimensional (3-D) seis-
a broad, arid storm -domina ted ram p with a variety mic with geologic and production data to describe
of rock types deposited . Diagenesis tha t is common the Lower Cretaceous (Aptian) reservoir in Abu Dhabi in
within the Arab-D reservoir includes severa l dissolu- "Volume-based Characterization of a Heterogeneous
tion even ts, recrystallization, and physical compac- Carbonate Reservoir, Lower Cretaceous, Abu Dhabi
tion. The resultant limesto ne porosity is a m ixture of (United Arab Emirates)." The. reservoir is positioned
interparticle (dom inant), moldic (c.ornmon), in tra- over a platform-to-basin transition and rec.ords a di-
particle (common), and microporosity (common) pore verse range of depositional facies and stratal geom-
types. Less common dolostone porosity is a mixture o f etries. A second-order sequence set is divided into
moldic (less common), intercrystalline (less common). five depositiona l sequen ces. Sequences 1 and 2 are a
and intracrystalline (least common) pore types. The transgressional phase showing the initial formation
vertical seal for the reservoi r is the overlying Arab C-D of buildup ma rgins and do minated by a lgaJ.prone
an hydrite. f•cies. The subsequent h i.ghstand phase of Sequence 3
In "High -resolution Seq uence Stratigraph y an d is mainly aggradational and records the proliferation
Reservoir Characterization o f Upper Thama ma (Low- of rudists across t he platfo rm !OJ>. A late highsta nd
er C retaceous) Reservoirs o f a Giant Abu Dhabi Oil phase of sequences 4 and 5 is progradational showing
Field, United Arab Emirates" authors Strohmenger, th e progressive downstepping of the platform margin
Weber, Ghan i, AJ-Mehsln, Al-Jeelani, AJ-Mansoori, onto a low-angle slope. Three-dimen sional seismic
Al-Dan•ani, Vaughan, Khan, and Mitchell describe data in the southern field area show a com plex mo-
the Lower Cretaceous Kh araib (Barrem ian and early saic of tidal channels, high-energy rudist shoals, and
Aptian) and Shuaiba (Aptian) formations (upperTha- intershoal ponds. The geometry and reservoir-quality
4 1 Harris and \Veber

variations of these geologic features have a strong Neutral Zone, Saudi Arabia and KuwaH: Key to Res·
impact on reservo ir S\oveep and co nformance in the ervoir Modeli ng and Assessme nt." Oi l production is
platform interior. ln the northern field area, seismic largely from subtidal dolomite formed on a very gently
images of prograding slope clinoforms reveal system - dipping, shallow, arid, and restricted ramp setting
atic variations in architecture and reservoir qualitythat that transitioned between no rma l ~marine conditions
reflect multiple scales of stratigraphic cyclicity. A pat· to restricted lagoonal environments. The key to mod·
tern gas llood has been Implemented in the clinoforms ellng the reservoir was the construction of an appro·
to add pressure support and improve recovery. Busi· priately d etailed sequence -stratigraphic fram ework
ness applications of the reservoi r framework include for use in building the geostatistical reservoir model.
(1) 3-0 seismic visua lization as a tool for optimizing Within the sequence-stratigraphic framework, 10 ltigh·
well placement, identifying bypassed reservoirs and frequency sequences are correlated, a lbeit with some
evaluating reservoir connectivity; (2) integra tion of dllficulty, across the entue field. Ultimately, diagen·
quantitative, volume-based seismic informatio n into esis is a major factor in the distribution of po rosity
reservoir models; (3) m<Jximjzing recovery through full and permeability. Dolomitjzation is pervasive, but fa·
integration of a ll subsurface data; and (4) enh<~nced cies exerts some control on the distribution of pO·
communicatio n a mong geoscientists and e ngmeers, rosity. Average porosity o f the reservoir interval is
leading to improved reservoir management practices. 15%, with values as much as 45%, a nd permeability
In "Sequence Strati.graphy and Reservo ir Arch itec- averages30 md wiUl core-plug measurements as much
ture of the Burgan and Mauddud Formations (Lower as 1200 md. The geostatistical model of the Maas-
Cretaceous), Kuwait" authors Strohmenger, Patter· trichtian reservoir demonstrates the layered a nd com ·
son, AI-Sahlan, Mitchell, Feldman. Demko, Wellner, partmentalized nature of the reservoir a nd clearly
Lehmann, McCrimmon, Broomhall, and AI·Ajml pro· shows that the location of the reservoir facies is con·
pose a new seq uence-stratlgraphic framework for the trolled by the o riginal depositional fabric and sub·
Burgan and Mauddud formations (Albian) of Kuwait, sequent dolomitization, both of which have been
which resu lts in a predictable distribution of reset· influenced by the paleotopography. Th is study w<Js
voir and seal fades. The Burgan and Mauddud for· undertaken to determine reservoir volumetrics, un·
mations form two second-order composite sequences, derstand the distributio n of inte rvals likely to yield
the o ldest of which constitutes the lowstaod, trans· h igher volumes of better quality oil, and provide a
gressive, and h ighstand sequence sets of the Burgan reservo ir property model fo r use in fluid-flow simu-
Formation. This composite sequence is subdivided into lation. Such an understanding is critica l to efficiently
high-frequen cy sequences that a re characterized by develop the 1.5 billion bbl oil Maastricthian resource
tidal-inlluenced, marginal-marine deposits in norUl · at Wafra field .
east Kuwait grading into more fluvial-dominated, con·
tinental deposits to the southwest. The younger com· West Africa
posite sequen ce consists of the lowstand sequence set In "Stratigraphic Organization and Predictability
of the uppermost Burgan formation and the trans· of Mixed Coarse· and fine-grained Lithofacies Succes·
gressive. and highstand sequence sets of the overly· sions io a Lower Miocene Deep-water Slope-channel
jng Maud dud l'ormation. This composite sequence is System, Angola Block 15," Porter, Sprague, Sullivan,
sand and mud prone in southern and southwestern Jennette, Beaubouef. Garfield, Rossen, Sickafoose, )en·
Kuwait and is carbonate prone in northern and north· sen, Friedmann, and Mohrig describe the lower lvlio·
eastern Kuwait. The lowstand se<Juence set of the Bur· cene slo pe-channel systems from Angola Block 15,
gan is subdivided into five high-frequency sequences. which regional seism ic mapping, exploration drilling,
and the Mauddud transgress ive and highstand se- and appraisal drilling have established as a world-
q ue nce sets a re subdivided into eight h igh-frequency class development opportunity. One o f the major de·
sequences. n 1e traditional lithostratigraphic Burgan - velopment targets in Block 15 is Burdigalian-aged
Mauddud contact is time transgressive. The upper slo pe-channel reservoirs, which are part of a system
Mauddud h ighstand sequence set is carbonate prone that traverses across the block in a n east - west direc.
and thins southward because of depositional thinning. tion and can be continuously mapped on adjacent
AuU10rs Dull, Garber, and Meddaugh describe the seism ic datasets for more than 30- 40 km (18- 24 ml).
Maastrichtian (Upper Cretaceous) reservoir in the giant The channel system was a sediment fairway for the
Wafra o il field in "The Sequence Stratigraphy of the delivery of coarse-grained turdidites and mixed mud·
Maastrichtian Reservoir at Wafra field, Partitioned dy and sandy debrites in to the Lower Congo basin.
Introduction I 5

Map patterns show distinctive changes in sinuosity, for understanding and modeling reservoir systems
channel confi nement, and degree of amalgamation worldwide that have similar geologic age and reser-
bro<tdly rel<tted to concurrent growth of salt-related voir architectu re.
structures. The episodic fill of the s lope-channel sys- In "Key Role o f Outcrops and Cores in Carbonate
tem can be better understood by a hierarchical ar· Reservoir Characterization and Modeling, Lower Perm -
rangement of unco nformity-bounded stratal units. ian Fullerton field, Permian Basin, United States" au-
Nested channels form composite channel complexes thors Ruppel and jones discuss the rock-based model
that show distinctive trends in lithofacies type and con.struction for fulle rton Clear fork field, which is
vertical facies succession . Conventional cores calibrat- a shallow-water platform carbonate reservoir of mid-
ed to well logs and high-resolution seismic data show dle Permian age in the Permian Basin of west Texas.
that the lower parts of the channel complexes are Fundamenta l steps in their study included (1) creat-
dominated by sandy-muddy debrites, slumps, and in- ing and applying an analogous outcrop depositional
jected sa ndsto nes. These facies are typically overlain model; (2) describing and interpreting subsurface core
by coarse-grained, gravelly, and well-amalgamated and log data in terms of this initial model; (3) defin-
sandy turbidites. The ove rlying facies succession is ing the sequence-stratigraphic arch ite~ture of the res-
more variable, but commonly consists o f interbedded ervoir section; (4) developing a ~)'de-based reservoir
sandy and muddy turbidites, injected sandstones, framework; and (5) defin ing controls, interrelation-
and a range of both muddy and sandy debrites. ships, and distribution o f porosity and permeability.
Data used in th is analysis included co res. th in sec-
North America tions, 3-D and two-dimensional (2-D) seism ic data,
Dubois, Byrnes. Bohllng, a nd Doveton docume nt borehole image Jogs, and outcrop models. Stratal ar-
the reservoir characterization and modeling from chitecture, differential dolomitization, karst fill, min-
pore to field scale of the giant Hugoton field in "Mul· eralogical variations, and rock-fabric distribution were
tlscale Geologic and Petrophysical Modeling of the incorporated into the model. 'lh ese components were
Giant Hugoton Gas field (Permia n), Kansas and Ok- used to con strain interpretation and d efinition of
lahoma." Their work on this mature Permian gas sys- flow units, permeability distribution, and saturation.
tem is aiding in defining original gas in place, the The rock-based methods demonstrated in this study
nature and distribution of gas saturation, and reser- provide key insights with broader application into
voir properties. The Kansas- Oklahoma part of the field the formatio n, characterization, and interpre tation
has yielded 963 billion m 3 (34 teO gas throughout a of cMbonate platform reservoirs.
70-year period from more than 12,000 wells. Most Authors Weissenlle rger, Wierzbicki, and Harland
remaini ng gas is in lower permeabiUty pay zones o f the describe the dee p Pan uke field in "Carbonate Se-
170-m (557-ft)·thick, differentially depleted, layered quence Stratigraphy and Pe troleum Geology of the
reservoir system. The main pay zones have remarkable jurassic Deep Pa nuke f ield, Off~hore Nova Scotia,
lateral continuity. They represent 13 shoaling-upward, Canada ." Deep Panuke, which is located 250 km
fourth-order marine-continental cycles, comprising (!55 mi) offshore of Halifax, Nova Scotia, Canada,
thin-bed ded (2-10-m; 6.6-33-ft), marine cMbonatc contains gas-fiUed cavernous porosiry in jurassic car-
mudstone to grainstone and siltstones to very fine sand- bonates of the Abenaki !'ormation. The Abena ki car-
stones. The pay zones are separated by low-reservoir- bonates range from Bathonian to Neocomian in age
quality eolian and sabkha redbeds. Petrophysical prop- and were deposited on broad carbonate platform at-
erties vaty among major lithofacies classes. Neu ral tached to a siliciclastic hinterland. Seven third-order
n etwork p rocedures, stochastic modeling, and auto- depositio nal sequences recogn ized in the Abenaki
mation fadlitated building a detailed full-field 3-D cel- are correlated with geology and a 2-D seism iC grid;
lular reservo ir model using a fou r-step workflow: 3-D seismic data are used (Ot deli neation drilling and
(1) define lithofacies in co re and correlate to electric reservoir characterization. Lithologies range fro m fo re-
log cu rves (training set); (2) train a neu ral network slope to reefal and shoal deposits near the platform
and predict li thofacies at noncored wells; (3) po p- margin. Open and restricted lagoon or tidal-flat de-
u late a 3-D cellu lar mode l wi th lithofacies using posits occur in th e platform interior. Siliciclastics
stochastic methods; and (4) populate m odel with are concentrated near sequence boundaries or are
lithofacies-specific petrophysical properties and fluid distributed along stri ke, close to point sources such
saturations. Both the knowledge gained and the tech- as rive rs cutting through the platform . Th e reser-
niques and workflow employed have implications voir occurs n car the platform margin in coral and
6 1 Harris and \Veber

stromatoporoid reef and associated skeleta l-peloidal sists of tluee onlapping cycles composed of a suc-
and occasionally oolitic shoal deposits. It comprises cession of aggradatio nal channel deposits, p rograda-
a range of porosity types, from vuggy limestone and tional bay-head delta and bay-fill deposits with a
dolo mite to microporosity in limestone. Geochemi- landward -stepping stacking pattern. The Mirador is
cal, isotOJ>e, and petrographic data suggest that the capped by restricted marine shales of the Carbonera
dolomitization and dissolution can be attributed to Formation.
deep burial and hydrothermal fluids.
ACKNOWLEDGMENTS
Sout h America
In "Sedimentology, Sequence Stratigraphy, and As organizers of the core workshop and editors of
Reservoir Architecture of the Eocene Mirador !'or- the companion voltune, we acknowledge several peo-
mation, Cupiagua Field, Llanos Footh.ills, Colombia," ple and our respec.tive companies for their assistance,
Ramon and Fajardo document the stratigraph.ic archi- without which the workshop and volume would not
tecture and facies distribution in a h igh-resolution have happened. Chevron and ExxonMobil provided
ti me-space framework to define the 3-D reservoir generous financial assistance that enabled numerous
zonati on of the Mirador Formation (Eocene) in the students to attend the core workshop and subsidized
Cupiagua field. The Cupiagua structure is a large, east- printing costs o f the worksl•op volume. SEPM staff
verging. asymmetric anticlinal fold that trends no rth- helped us o ve rco me n umerous o rganizational and
northeast in the hanging wall o f the frontal fault. The logistica l issues that confronted us as we organized
Mirador Formation accounts for approximately 55% the workshop. AAPG staff and, in particular, Beverly
of the recoverable oil in Ute field. Three scale.~ of strati· Molyneux provided us invaluable help in the editing
graphic cycles are recognized based on stacking pat· and publication of the workshop volume. We thank
tern and general lrend of facies successions: short- all of the authors wh o worked bard to prepare poster
term cycles or progradational and aggradational units and core displays, oral presentations, and the de-
stack systematically into intermediate-term cycles, tailed manuscripts t11at describe their respective fields.
which, in turn, are grouped into long-term cycles. Finally, we thank the many who helped us in the
The lower half of the Mirador consists of flood-plain technical review of the manuscripts for the workshop
facies with channel, crevasse splay/ and swamp and publication: Bob Alway, Steve Bachtel, Sherry Becker,
flood-plain facies successions. Bay-head delta and bay- Kelley Bergman, j ohn Bova,joel Colli ns. Bob Dalrym-
fill facies occur in the upper half of the Mirador Forma- ple, Laurence Droz. Ch.ip Feazel, Sean Guidry, Jurgen
tion. The Lower 1\·firador consists or two intermediate- Grotseh, jean Hsieh, jon Kaufman, Mike Kozar, Dale
sca le cycles s howing a seaward-stepping stacking leckie, j ose Matos, jim 1\kGovney, Gary Parker, Carlos
pattern overla in by a third cycle with a landward- Plrmez, George Pemberton, Linda l>rJce, Gene Rankey,
stepping pattern, and the upper Mirador conti nues Rick Sarg, Toni Simo, Krishnan Srinivasan, and Niall
the landward-stepping pattern . This upper unit con- Toomey.
1
Kenter, J. A. M., P. M. Harris, J. F. Collins, L. J. Weber, G. Kuanysheva, and
D. J. Fischer, 2006, Late Visean to Bashkirian platform cyclicity in the cen-
tral Tengiz buildup, Precaspian Basin, Kazakhstan: Depositional evolu-
tion and reservoir development, in P. M. Harris and L. J. Weber, eds., Giant
hydrocarbon reservoirs of the world: From rocks to reservoir characteriza-
tion and modeling: AAPG Memoir 88/SEPM Special Publication, p. 7 – 54.

Late Visean to Bashkirian Platform


Cyclicity in the Central Tengiz
Buildup, Precaspian Basin,
Kazakhstan: Depositional Evolution
and Reservoir Development
J. A. M. Kenter1 L. J. Weber
Vrije Universiteit, Amsterdam, Netherlands ExxonMobil Development Company,
Houston, Texas, U.S.A.
P. M. Harris
Chevron Energy Technology Company, G. Kuanysheva
San Ramon, California, U.S.A. TengizChevroil, Atyrau, Kazakhstan

J. F. Collins D. J. Fischer
ExxonMobil Development Company, TengizChevroil, Atyrau, Kazakhstan
Houston, Texas, U.S.A.

ABSTRACT

T
he Tengiz buildup, an intensely cored and studied isolated carbonate
platform in the Precaspian Basin, contains a succession of shallow-water
deposits ranging from Famennian to Bashkirian in age. From a reservoir
perspective, Tengiz can be subdivided into platform (central and outer) and rim-
slope (flank) regions. The upper Visean, Serpukhovian, and Bashkirian form the
main hydrocarbon-bearing interval in the platform. Depositional cycles (high-
frequency sequences) in this interval are several to tens of meters thick for the
Visean and Serpukhovian, and decimeter to meter scale for the Bashkirian.
Cycles are made up of a succession of lithofacies overlying a sharp base that
locally shows erosion, calcretes, meteoric diagenesis, and other evidence for
subaerial exposure. At the base of the succession, tight peloidal mudstone and
ash beds are associated with sequence boundaries and are thought to reflect low-
energy conditions developed in deeper platform areas at lowstand and during

1
Present address: Chevron Energy Technology Company, San Ramon, California, U.S.A.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215873M88374

7
8 / Kenter et al.

initial flooding. Above this, beds with in-situ articulated brachiopods signal ini-
tial open-marine but still low-energy conditions. Succeeding crinoid-dominated
intervals represent maximum marine flooding and overlying skeletal-peloidal
grainstones highstand shoaling phases.
Visean and Serpukhovian cycles are generally easy to correlate from well to
well over several kilometers distance. Volcanic ash beds are identified by gamma-
ray spikes, and flooding intervals show as low-porosity zones. In contrast, Bash-
kirian cycles are thinner and incomplete, dominated by thin, peloidal mudstone
intervals alternating with high-energy coated-grain and ooid grainstone, and are
more difficult to correlate. High-frequency icehouse sea level fluctuations ex-
posed the platform during each fall of sea level, and rapid flooding resulted in
incomplete cycles and complex lateral facies changes that may explain relatively
poor lateral continuity of log character.
The distribution of reservoir rock types in the central platform is deter-
mined by burial diagenetic modification of an earlier reservoir system that includes
meteoric alteration and porosity enhancement below major sequence bound-
aries and reduced dissolution along higher order sequence boundaries associated
with the presence of volcanic ash. The lateral continuity of tight layers at se-
quence boundaries probably greatly affected later fluid flow as well as the ulti-
mate distribution of cements, dissolution, and bitumen in the central platform
reservoir.
The burial diagenetic overprint included two major phases of reservoir modi-
fication. First, a corrosion and cementation phase significantly enhanced existing
matrix porosity in the interior central platform while reducing porosity in the
exterior central and outer platform by pore-filling equant calcite cement. This was
followed by bitumen emplacement and associated corrosion. These processes not
only exerted an overall porosity-reducing effect prior to and associated with
bitumen invasion toward the exterior central platform, but also dampened or
flattened the initial cyclic porosity variations and obscured relationships between
pore types and permeability. The bitumen overprint is nearly absent in the
innermost platform wells; bitumen concentrations are highest near the bases of
the cycles, which may imply that the first fill of hydrocarbons migrated through
the flanks laterally into the platform cycles.

INTRODUCTION Tengiz field produces a light, intermediate-sulfur,


stabilized tank oil of approximately 478 API. As of mid-
Tengiz Field History year 2005, more than 115 wells have been drilled on
The Tengiz field, located in western Kazakhstan, near Tengiz. The highest rate wells are located in the plat-
the northeastern shore of the Caspian Sea (Figure 1), form margin and slope in fractured carbonates with
produces oil from an isolated carbonate platform low (<6%) matrix porosity. Platform wells exhibit
(aerial extent of >110 km2 [42 mi2]) of Devonian and higher porosity (as much as 18%), but matrix perme-
Carboniferous age. Tengiz was discovered in 1979 by ability is typically low (<10 md).
the Ministry of Oil Industry of the Soviet Union. The
discovery well Tengiz 1 (T-1) reached a total depth Regional Paleogeography
of 4095 m (13,435 ft). Development drilling of the The Precaspian Basin is a large Paleozoic basin that
Tengiz field commenced in 1983, and onsite con- occupies much of the area known today as the north-
struction of plant-processing facilities began in 1987. ern part of the Caspian Sea and adjacent landmass.
Field production officially began in April 1991. Since During the Early Carboniferous, the Precaspian Basin
1993, TengizChevroil, an in-country joint venture was a major equatorial sag basin in the western part of
company run by Chevron, has operated Tengiz and Kazakhstan (Ross and Ross, 1985). Carbonate shelves
the adjacent Korolev field. formed along the basin margins, and isolated, broad
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 9

FIGURE 1. Index map locating North Caspian region and outlining area of the Tengiz field in blue, which is shown in
more detail in Figure 2.

carbonate platforms grew on preexisting basement associated with the closure of the Precaspian Basin.
highs (Cook et al., 1997). Together with other nearby Initially, a veneer of Moscovian through Artinskian
fields, Tengiz forms an archipelago of isolated car- deep-water carbonates and volcaniclastics accumulat-
bonate platforms that grew on a regional pre-middle ed around and on top of the buildups. These deposits
Devonian structural high near the southeastern mar- were followed by thick Kungurian evaporites, includ-
gin of the basin (Figure 2). The initial carbonate plat- ing halite that completely encased the platforms.
forms in the archipelago formed sometime after the
Middle Devonian and continued more or less unin-
terrupted until the early Bashkirian. The total thick- SUPERSEQUENCE FRAMEWORK
ness of carbonate that accumulated at Tengiz during
that time has not been drilled but is represented by In 2001, a joint study by ExxonMobil, Chevron,
approximately 1.2 s two-way traveltime on the most and TengizChevroil established a stratigraphic frame-
recent three-dimensional seismic survey (Figure 3). work for the Tengiz platform based on seismic, bio-
Following platform termination in the early Bashkir- stratigraphic, core, and well-log data. This study was
ian, the platforms were encased in a series of lithologies published in late 2003 (Weber et al., 2003) and stands
10 / Kenter et al.

FIGURE 2. Paleogeographic map showing the archipelago of isolated carbonate platforms, including Tengiz, that grew
on a regional premiddle Devonian structural high near the southeastern margin of the Precaspian Basin.

as the definitive stratigraphic reference for Tengiz. constitute the platform reservoir part of the Tengiz
The framework consists of a hierarchy of sequences buildup.
that have continued to be maintained as Tengiz dril- Thin subaerial exposure caps on subtidal lithofa-
ling proceeds. cies form the most common type of cycle top in the
Through the main hydrocarbon-bearing interval Lvis_SSB to Bash_SSB succession on the Tengiz plat-
at Tengiz (Famennian – Bashkirian), seven bounding form. The implication of this type of cycle boundary
discontinuity surfaces (sequence boundaries and max- is that rapid base-level fall exposed subtidal litho-
imum flooding surfaces [MFSs]) are recognized on facies without development of an intervening tidal
seismic data (Weber et al., 2003) (Figure 4; Table 1). flat or peritidal succession. Alternatively, anchor points
Two additional MFSs, Lvis1_MFS and Lvis2_MFS, are for tidal flats may not have been available because the
identified primarily from well control (cores and depositional profile near the margin gradually deep-
well logs). Four supersequences (Tournaisian – lower ened toward the platform break without the pres-
Visean, lower Visean – upper Visean, upper Visean – ence of a shallow barrier. In other words, the Tengiz
Serpukhovian, and Bashkirian) extend from the Fa- platform was not conducive to producing extensive
mennian supersequence boundary (Fame_SSB) to the tidal-flat facies. Each cycle is more appropriately
Bash_SSB (Weber et al., 2003). Each supersequence termed a ‘‘high-frequency sequence,’’ capped by a high-
contains a transgressive sequence set (TSS) and a high- frequency sequence boundary. Cycle boundaries also
stand sequence set (HSS) separated by an MFS. The occur where subaerial exposure is not evident from
upper two supersequences are the focus of this chapter. core and thin-section description. Without evidence of
Three reservoir zones (Visean A, Serpukhovian, and subaerial exposure, upper bounding surfaces of these
Bashkirian) are contained in these supersequences and cycles define parasequences.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 11

FIGURE 3. (A) Map of Tengiz showing the location of seismic line shown in (B). The map shows top reservoir (=top
Bashkirian) structure with a contour interval of 100 m (330 ft). (B) Seismic profile showing general depositional areas of
Tengiz buildup (central platform, outer platform, rim slope, or flank) and key reservoir zones (RZ). Vertical scale is
two-way traveltime in milliseconds. See text for discussion.

Upper Visean–Serpukhovian Supersequence 2006) may exceed 600 m (1968 ft), where in-situ mi-
(Visean A and Serpukhovian Reservoir Zones) crobial boundstone and allochthonous debris aprons
The upper Visean – Serpukhovian supersequence of platform- and slope-derived grainy carbonate and
is defined as the interval between the Lvis_SSB and breccia accumulated (Figure 3). The upper Visean –
the Serp_SSB (Figure 4). It includes the Mikhailovsky, Serpukhovian supersequence spans 14 m.y. (Gradstein
Venevsky, Tarussky, Steshevsky, Protvinsky, and Za- et al., 2004).
paltyubinsky regional horizons (Weber et al., 2003). The Lvis2_MFS separates the TSS of the super-
The Zapaltyubinsky zone is not generally recognized sequence from the overlying HSS (Figure 5). The TSS
on the Tengiz platform but is observed in wells that consists of five composite sequences that are domi-
penetrate the slope and basin. The Zapaltyubinsky nantly aggradational and do not extend beyond the
may therefore represent a basinward shift in sedi- platform break of the underlying lower Visean–upper
mentation and form a lowstand sequence set (Weber Visean supersequence. Small-scale backsteps are pos-
et al., 2003). However, rim failure processes (see Collins sible, but not proven, at or near the platform break
et al., 2006) have obscured stratal patterns that clear- through the TSS. Deeper water, low-energy open-marine,
ly define the formation of such a depositional body. platform lithofacies in these sequences include crinoid
On the platform, the upper Visean – Serpukhovian grainstone-packstone and poorly sorted, thick-walled
supersequence is approximately 250 m (820 ft) thick; brachiopod grainstone-packstone. Shallower water,
however, equivalent slope deposits (see Collins et al., high-energy, open-marine to back shoal lithofacies
12 / Kenter et al.

FIGURE 4. (A) Map of Tengiz showing the location of cross section shown in (C). The map shows top reservoir (=top Bashkirian) structure with a contour interval
of 100 m (330 ft). (B) Legend for log curves shown in cross section of (C). Gamma-ray (SGR and CGR) is shown on the left and porosity (PHIE) on the right. (C) Cross
section through Tengiz showing major stratigraphic surfaces and reservoir zonation scheme. Stratigraphic surfaces, in ascending order, are Tournaisian (Tour_MFS);
early Visean (Evis_SSB); late Visean (Lvis1_MFS, Lvis_SSB, and Lvis13_csb); Serpukhovian (Serp_SSB); and Bashkirian (Bash_SSB). Reservoir zones bracketed by
these surfaces are Visean D (VisD), Visean C (VisC), Visean B (VisB), Visean A (VisA), Serpukhovian (Serp), and Bashkirian (Bash). The Visean A, Serpukhovian, and
Bashkirian are the subject of this study.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 13

Table 1. Summary of sequence-stratigraphic nomenclature for Tengiz.*


Stage Reservoir Zones Sequence Top Surface Sequence Number

Bashkirian Unit 1 Bashkirian Bash6 Bash_SSB 1


Bash4 Bash4_csb 2
Bash2.5 Bash2.5_csb 3
Bash1 Bash1_csb 4
Serpukhovian Serpukhovian Serp5 Serp_SSB 5
Serp4 Serp4_csb 6
Serp3 Serp3_csb 7
Serp2 Serp2_csb 8
Serp1 Serp1_csb 9
Late Visean Visean A Lvis13 Lvis13_csb 10
Lvis12 Lvis12_csb 11
Lvis11 Lvis2_MFS 12
Lvis10.5 Lvis10.5_sb 13
Lvis10 Lvis10_csb 14
Lvis9.5 Lvis9.5_sb 15
Lvis9 Lvis9_csb 16
Visean B Lvis8 Lvis_SSB 17
Lvis7 Lvis7_csb 18
Lvis6 Lvis6_csb 19
HRZ Lvis5_csb 20
Unit 2 Visean C Lvis4 Lvis1_MFS 21
Lvis3 Lvis3_sb 22
Lvis2 Lvis2_sb 23
Lvis1 Lvis1_sb 24
Early Visean Evis9 Evis9_sb 25
Evis8 Evis8_sb 26
Evis7 Evis7_sb 27
Evis6 Evis6_sb 28
Visean D Evis5 Evis_SSB 29
Evis4 Evis4_sb 30
Evis3 Evis3_sb 31
Evis2 Evis2_sb 32
Evis1 Evis1_sb 33
Tournasian Tournasian Tour6 Tour_MFS 34
Tour5 Tour5_sb 35
Tour4 Tour4_sb 36
Tour3 Tour3_sb 37
Tour2 Tour2_sb 38
Tour1 Tour1_sb 39
Famennian Unit 3 Famennian A Fame2 Fame_SSB 40
Famennian B Fame1 Fame_MFS 41

*This framework was established by Weber et al. (2003), and picks in wells have, since then, been refined as a result of new wells and
progressed knowledge of the field.

include well-sorted grainstone and grainstone-pack- (composite sequence boundary) is composed of three
stone consisting of algae, foraminifera, peloids, and composite sequences (Figure 5). Significant progra-
coated grains. dation is observed in four subsequent composite se-
The upper Visean–Serpukhovian HSS is subdivided quences from the Serp1_csb to the Serp_SSB as shallow
into an aggradational and progradational phase. The platform lithofacies fill available accommodation
aggradational interval from Lvis2_MFS to Serp1_csb space and prograde to the present-day location of the
14 / Kenter et al.

FIGURE 5. North – south cross section through wells T-5044, T-5246, T-5447, T-220, T-6246, and T-6846; well locations
are shown in Figure 4A. Cross section show detailed lithofacies and cycle correlations in the central platform region in
the Visean A (VisA), Serpukhovian (Serp), and Bashkirian (Bash) supersequences. The lithofacies legend was simplified
from that of Weber et al. (2003) by reducing the number of facies types and relating them to the level of energy
during deposition. Sequence stratigraphy is adopted from Weber et al. (2003); triangles identify higher order sequences
(fourth and higher order) correlated from well to well. See text for further explanation and discussion.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 15
16 / Kenter et al.

platform break. As much as 400–500 m (1600–1900 ft) several hundreds of meters thick. The remaining
of depositional topography was filled on the slope thickening could be explained by differential com-
as shallow-water platform lithofacies prograded over paction between the massive and mechanically strong
predominantly slope boundstone of microbial origin boundstone and the compacted central platform. The
(Figure 3). During the progradational part (Serpuk- latter is supported by the petrographic observation
hovian) of the upper Visean – Serpukhovian HSS, the that most grainstones are overcompacted.
Tengiz buildup was an overall flat-topped platform
with a substantial slope apron collecting debris primar- VISEAN A, SERPUKHOVIAN, AND
ily sourced from autochthonous boundstone growing BASHKIRIAN PLATFORM ARCHITECTURE
on the outermost platform and upper slope and, to a
lesser extent, from the platform top. The platform in- The Visean A, Serpukhovian, and Bashkirian reser-
terior is dominated by grainy textures (packstone and voir zones are significant because they account for
grainstone), with packstone dominating in the deeper the bulk of the oil production from the platform part
subtidal areas and well-sorted grainstone in the higher of the buildup. The platform is subdivided into differ-
energy shallow subtidal parts of the platform. ent paleogeographic or depositional areas: central plat-
form, outer platform, and rim slope (flank) (Figure 3).
Bashkirian Supersequence This chapter focuses on details of the central plat-
(Bashkirian Reservoir Zone) form part of the Visean A, Serpukhovian, and Bash-
The Bashkirian supersequence occurs between the kirian intervals.
Serp_SSB and the Bash_SSB (Figures 4, 5) and spans To refine the central platform stratigraphy, a care-
approximately 10 m.y. (Gradstein et al., 2004). The ful comparison of core and thin sections in a series of
oldest Bashkirian horizon (i.e., Bogdanovsky) and north –south wells (T-5044, T-5246, T-5447, T-220,
most of the overlying Syuransky horizon do not ap- T-6246, and T-6846) produced a detailed lithofacies
pear to be present on the Tengiz platform (Brenckle correlation across the central platform region in the
and Milkina, 2001). The Akavassky and Askynbashsky Visean A, Serpukhovian, and Bashkirian reservoir zones
horizons are generally present on the Tengiz platform (Figure 5). Several general observations can be made
above the Syuransky horizon. Although more work is on the cycle stacking pattern, overall depositional
necessary, results suggest that the early Bashkirian at framework, and facies patterns visible in Figure 5; they
Tengiz is transgressive, i.e., TSS of the Bashkirian su- are summarized here and discussed in more detail in
persequence (Figure 5). The HSS of the Bashkirian subsequent sections.
supersequence is very thin and condensed. Biostrati- The succession of Visean A, Serpukhovian, and
graphic analysis of core and cuttings typically dates Bashkirian reservoir zones are, respectively, the TSS,
the Bash_SSB as the top of the early Bashkirian. The HSS, and TSS of the two supersequences previously
late Bashkirian is condensed and represents starved mentioned, and each consists of several composite
sedimentation on an isolated platform, probably de- and higher order sequences (cycles). Cycle stacking
posited in deep water (Weber et al., 2003). patterns (shoaling cycles as indicated by triangles in
Throughout most of the central part of the Tengiz Figure 5) in the Visean A and Serpukhovian reservoir
platform, the Bashkirian supersequence is uniform zones (and, respectively, TSS and HSS intervals) show
in thickness (80 – 100 m; 262 –330 ft). The Bashkirian a repetition of a thick cycle (15– 20 m; 49 – 66 ft) with
platform break is generally aggradational through commonly a basal crinoid- and brachiopod-rich sand
three composite sequences (Bash 1, 2.5, and 4) and followed by a series of thinner cycles (2 – 8 m; 6.6 –
coincident with the underlying platform break at the 26 ft). Those crinoid- and brachiopod-rich facies are
Serp_SSB. In isolated areas along the eastern margin interpreted as generally deeper water and relatively
of the platform, the Bashkirian section thickens to low energy and commonly extend all across the plat-
about 150 m (492 ft) (e.g., T-5056 and T-7252), with form, which would suggest significant flooding fol-
some of this thickening resulting from a facies change. lowing a lowstand of relative sea level. Ten to twelve
Well-log correlations in these areas indicate that the composite sequences, each containing two to six high-
uppermost composite sequence is locally 30–60 m er order sequences (cycles), are shown in Figure 5.
(98–196 ft) thick (Weber et al., 2003), seismic data The boundaries of these composite sequences do not
shows locally mounded seismic facies, and cores show always exactly match those by Weber et al. (2003),
skeletal packstone and local microbial boundstone on although the number of composite sequences is es-
top of Serpukhovian microbial cement boundstone sentially the same.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 17

The composite sequences described herein bundle ate but substantial differences from the central plat-
into four larger sequences that each have a thick and form area and, therefore, a potentially discontinuous
laterally continuous brachiopod or crinoid interval transition from platform to flank. The situation is more
near the base. Minor differences are observed be- complicated toward the eastern margin, where appar-
tween wells, but in general, the stacking pattern de- ent top-lap geometry of the Serpukhovian and Bash-
scribed above is very consistent. In the Bashkirian in- kirian platform develops over a Serpukhovian bound-
terval (mostly interpreted as TSS), the stacking pattern stone slope. In the underlying Visean A platform, the
is more complex and shows a lower interval with thick transition may be comparable to that recognized from
cycles (10 m; 33 ft), followed by an interval domi- the western margin, wherein the Serpukhovian and
nated by thin (2–4-m; 6.6–13-ft) cycles, and an interval Bashkirian central to outer platform transition seems to
near the top of intermediate cycles (4 – 8 m; 13 – 26 ft). be gradual and changes from deeper outer platform with
No clear stacking pattern is observed, and although intercalated boundstone intervals into boundstone
lateral thickness variations are significant, the succes- upper slope. As a consequence, the lateral facies asso-
sion of thick, thin, and intermediate cycle intervals ciations below the pseudo-top-lap may be significantly
seems to be laterally continuous. different from those above. See Collins et al. (2006)
The moderate differences between the sequences for further discussion related to these transitions.
established by Weber et al. (2003) and those described
herein may lead to slight modifications of the plat- Lithofacies Types in the
form sequence-stratigraphy framework. One poten- Central and Outer Platform
tial modification is changing the MFS in the Visean Core penetrations available for the Tengiz sequence-
A – Serpukhovian supersequence from its position at stratigraphic framework developed by Weber et al.
Lvis2_MFS as previously discussed to the base of the (2003) were limited to TengizChevroil wells T-220,
thick brachiopod interval observed between Lvis12_csb T-5246, T-6846, and incomplete cores from numer-
and Lvis13_csb, which marks a shift from a series of ous older Russian wells (figure 3 of Weber et al., 2003).
thinning cycles below to thick cycle above and an over- The present study builds on that of Weber et al. (2003)
all change of generally deeper facies below to higher by adding the information from cores that nearly com-
energy shallow-water facies above (Figure 5). pletely span the Visean A, Serpukhovian, and Bashkir-
These sequences of the Visean A, Serpukhovian, and ian in several key wells in the central (T-5044, T-5447)
Bashkirian intervals span a transition from greenhouse and southwestern outer platform (T-6246) (Figure 4A).
to icehouse cyclicity as suggested by observed trends in These complete core penetrations for the first time
cycle thickness and general depositional environment. allowed a detailed comparison of each individual cy-
Another important observation is that sequences and cle (and facies types within) in the study window.
cycles (third order and higher order) display a high Evaluation of core observations, petrography, and
degree of lateral continuity and constant thickness wire-line and core gamma logs (porosity was exclud-
across the central platform area into the outer platform ed during this phase) resulted in a coherent catalog of
zones over distances of more than 10 km (6 mi) north 10 key lithofacies types that are ranked primarily based
to south. Careful inspection of this railroad pattern on the relative hydrodynamic energy during deposi-
suggests the presence of a very low-angle top lap and tion (Table 2). This information was extracted through
possible truncation of underlying cycles at the Serp_SSB environmental indicators like grain size, sorting, ma-
surface (the cross section of Figure 5 is flattened on this trix type, sedimentary structures, and grain types (bio-
level). The truncation is variable but generally increas- ta). The resulting 10 facies or rock types can be grouped
ing to the south, indicating northward paleodip. into three or four composite rock types, reflecting
In general, it can be stated that the central plat- major differences in energy conditions during depo-
form area as documented in this chapter is several sition. The catalog was also designed to assist in the
kilometers inboard from the physical break between characterization of plugs and petrography of the key
platform and flank. The outer platform contains areas wells at 1-ft (0.3-m) intervals and serve as a direct link
that remain poorly defined because of a lack of core between geological parameters to reservoir properties.
penetrations and/or uncertainty of the physical and
sedimentological character of the central platform-to- Cyclic Lithofacies Variations
flank transition. Core penetrations close to the western In the central platform, lithofacies stack into
edge of the platform are uncommon, but those avail- shallowing-upward cycles. The two supersequences
able, wells T-44, T-6846, and T-6743, suggest moder- and three reservoir zones making up the Lvis_SSB to
18 / Kenter et al.

Table 2. Summary of lithofacies, pore types, porosity, and bitumen content for central platform wells T-220 and
T-6246.*

Facies Fabric EoD

Abbreviation General Description New Facies Simplified


Code Number Fabric Code

Algal grainstone Algal dominated silt to fine sand G, 46 Grainstone High-energy reef flat,
(AG) very well sorted, with minor coated grains backshoal-inner platform
(smaller than 500 mm)
Skeletal intraclast Mixed skeletal-(peloidal) G, moderately well 44 Grainstone High-energy reef flat,
grainstone (SIG) sorted, fine to medium sand, varying shell backshoal-inner platform
hash, minor crinoids, abundant intraclasts,
oncoids, coral (massive) and Chaetetes
fragments, forams, algae
Coated grain Ooid-dominated G, well to very well sorted, 43 Grainstone High energy above
grainstone-rudstone minor crinoids or shell hash, intraclasts, wave base-intertidal
(CgGR) forams; locally rundstone with intraclasts; platform
cross-lamination
Ooid grainstone Skeletal G, fine to medium sand, well to very 42 Grainstone High energy above
(OG) well sorted, coated grains, minor crinoids or wave base-intertidal
shell hash, forams, intraclasts, algal fragments, platform
locally overpacked; locally rudstone with
intraclasts; cross-lamination
Well-sorted Mixed skeletal-(peloidal) G, moderately well 41 Grainstone Moderate to High
skeletal sorted, fine to medium sand, minor to energy above wave
grainstone (SPG) abundant shell hash, moderate crinoids, base-open platform
intraclasts, forams, algae
Poorly sorted Skeletal-peloidal GP, coarse sand, moderately 32 Grainstone- Moderate to high
skeletal-peloid to poorly sorted, mixed crinoids and packstone energy below or around
grainstone-packstone brachiopod shell hash, minor in-situ wave base-open platform
(SPP) (thick-walled) brachiopods, peloids, algae,
forams, whole and fragments of stick corals
Crinoid Skeletal-peloidal GP, medium to coarse sand, 31 Grainstone- Moderate to low energy
skeletal-peloid moderately to poorly sorted, dominant packstone below or around wave
grainstone-packstone crinoids (1 – 2 mm or larger), minor whole base-open platform
(CSP) brachiopods and hash, peloids, forams,
(green and tubular) algal fragments
Brachiopod Skeletal-peloidal GP, coarse sand, poorly 2 Grainstone- Moderate to low energy
skeletal-peloid sorted, brachiopods in-situ (thin vs. packstone below or around wave base
grainstone- thick-walled; 1-5/20-50), bedded, variable open platform (abundant
packstone crinoids, peloids, divers algal fragment, crinoids an thick-walled
(BSP) abundant and divers forams, whole and brachiopods) versus shallow
fragments of stick corals protected platform interior
(minor crinoids and
thin-walled brachiopods)
Peloid Ranging from mud to peloidal PW, 13 Packstone- Low-energy restricted;
packstone- silt-fine sand size with small oncoids, grainstone restricted lagoon to
wackestone calcispheres, thin-walled bivalve supratidal
(PPW) fragments and ostracods
Volcanic ash Volcanic ash 12 Clay mudstone Low-energy restricted
(VA) inner platform to
postdrowning
All
Grain-supported mud-lean; high energy 41 – 46
Grain-supported muddy; intermediate 2, 31, 32
energy
muddy; low energy 13

*Lithofacies, also shown in Figure 5, are based on Weber et al. (2003), but simplified by reducing the number of facies and relating them to the level
of energy during deposition. The vertical arrangement of lithofacies in this table resembles the ideal and most complete succession of lithofacies in a
shoaling cycle or sequence. Pore types are 1 = interparticle; 2 = intraparticle; 3 = fenestral; 4 = shelter; 5 = growth framework; 6 = intercrystalline; 7 =
moldic; 8 = microporosity; 9 = fracture; 10 = channel; 11 = vug; and 12 = enhanced dissolution. Note that bitumen analyses for well T-6246 are not
yet complete. See text for discussion.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 19

T-220 T-6246

N Porosity Bitumen (% Rock Volume) N Porosity

Dominant Range Mean Range Mean Dominant Range Mean


Pore Types (%) (%) Pore Types (%) (%)

24 11, 1, 6, 2 2.25 – 17.59 13.08 0.16 – 0.31 0.23 54/48 1, 7, 6, 11, 2, 9 0.84 – 14.80 4.90

18 0.18 – 16.35 4.46 0.08 – 0.41 0.28 22/22 1, 2, 7 0.97 – 12.07 4.58

78 11, 1, 7 0.42 – 23.66 10.09 0.13 – 6.86 1.35 36/36 1, 7, 11, 6, 9, 2 0.37 – 14.91 4.43

100 11, 1, 7, 2 0.78 – 21.36 11.79 0.01 – 2.73 0.40 103/78 1, 7, 6, 11, 2, 9 0.41 – 20.11 4.74

188 11, 1, 7, 6 0.19 – 23.18 11.64 0.04 – 6.36 0.23 217/114 1, 11, 2, 6, 7, 8 1.21 – 16.18 6.81

217 11, 1, 6, 7, 2 0.50 – 17.33 12.56 0.06 – 4.04 0.35 262/90 1, 11, 2, 9, 6, 7 0.97 – 15.30 6.91

83 11, 6, 1, 7, 9 0.34 – 14.79 5.37 0.05 – 5.54 2.58 79/14 1, 2, 11, 6, 7, 8 1.30 – 18.48 8.28

199 11, 1, 6, 2 1.45 – 17.53 10.51 0.09 – 5.49 0.68 207/27 11, 1, 6, 9, 2, 7 0.63 – 16.50 8.42

53 11, 9, 1, 6, 7, 2 0.07 – 17.55 3.62 0.11 – 3.72 0.67 49/14 1, 2, 11, 6, 7, 9 1.02 – 15.50 8.16

971 11, 1, 6, 7, 2, 9 0.07 – 23.66 10.44 0.01 – 6.87 0.74 1031 1, 11, 2, 6, 7, 9 0.37 – 20.11 6.89
408 11, 1, 7, 2, 6 0.18 – 23.66 11.15 0.01 – 6.87 0.24 432 1, 11, 7, 6, 2, 9, 8 0.37 – 20.11 5.77

510 11, 6, 1, 2, 7 0.34 – 19.34 10.58 0.05 – 5.54 0.84 549 11, 1, 2, 6, 7, 9, 8 0.63 – 18.48 7.67

53 11, 9, 1, 6, 7, 2 0.07 – 17.55 3.62 0.11 – 3.72 0.67 48 1, 2, 11, 6, 7, 9 1.02 – 15.50 8.16
20 / Kenter et al.

Bash_SSB interval comprise more than 50 cycles rang- Peloid Packstone-wackestone (PPW) and
ing in thickness from 2 to 15 m (6.6 to 49 ft) (Table 1; Volcanic Ash (VA)
Figure 5). Commonly, cycles are bounded by thin vol- Peloid packstone-wackestone caps many shallowing-
canic ash layers less than a few centimeters in thick- upward carbonate cycles and can overlie any of the
ness that have been altered by burial diagenesis to other lithofacies. Usually less than 0.5 m (1.6 ft) thick,
resemble shale partings. Closely spaced volcanic ash this lithofacies exhibits rhizoliths, laminated crusts,
partings are sometimes observed near bounding sur- alveolar fabric, uncommon fenestrae, low-diversity
faces. Thicker ash beds do occur in some parts of the microfauna (ostracods and calcispheres), and uncom-
section (e.g., just above the Lvis_SSB and the Serp_SSB) mon megascopic biota (Figure 6). Peloid packstone-
and may exceed 1 m (3.3 ft) in thickness. Laminated wackestone includes the development of a pedolithic
ash beds were likely deposited (and preserved) during zone that occurs at most cycle boundaries, which is
periods of low energy, either during exposure of the interpreted to include both the top of a cycle prior to
platform top or during the initial flooding and devel- exposure and the exposure event itself, as well as the
opment of low-energy lagoonal deposits. The pres- initial reflooding event with intermittent exposure.
ence of thick (decimeter to meter) intervals of vol- Peloid packstone-wackestone nearly always has volca-
canic ash indicates major periods of exposure and/or nic ash interbeds or dispersed volcanic ash (Figure 6A)
deep initial flooding. These are therefore key sur- and, probably as a result, is tight and well cemented.
faces that guided the selection of the supersequence Volcanic ash is locally also dispersed within high-
boundaries. energy facies immediately below the Serp_SSB and
When presented in the context of their deposi- below the Serp1_csb in well T-5447 (Figure 5). Those
tional setting related to energy (i.e., bottom agitation) intervals are similarly tight and well cemented. The
at the time of deposition, lithofacies roughly form a frequent occurrence of volcanic ash interbeds argues
succession of low to higher energy from the base to for nearly continuous fallout during deposition of
top of a depositional cycle (see Table 1). One caveat in the sequences; ash layers accumulated only during
relating the lithofacies with energy level is that there subaerial or low-energy conditions, thereby marking
is no one-to-one correspondence. For example, simi- cycle boundaries on the gamma-ray logs. The pres-
lar low-energy facies occur both in deeper subtidal as ervation of volcanic ash is related to terrestrial fall-
well as in more restricted shallow depositional envi- out deposition during exposure and during initial
ronments. For discussion purposes, such a succession low-energy conditions while the platform was sub-
is shown in Table 1 and partly demonstrates litho- merged, probably in a shallow, restricted lagoon.
facies stacking from base to top. A brief description
is provided in the sections immediately following, Brachiopod and Crinoid Skeletal-peloid
whereas the implications for the general depositional Grainstone-packstone (BSP and CSP)
system (spatial distribution of lithofacies types and An inferred rise in relative sea level resulted in
cycle thickness) are discussed in subsequent sections. slightly deeper conditions and deposition of skeletal
In brief, the description of lithofacies types and grainstone to packstone at the base of a cycle. The basal
associated depositional energy, from base to top, is part of the cycles can be either brachiopod skeletal-
as follows: cycles generally start with a thin peloid peloid grainstone-packstone or crinoid skeletal-peloid
packstone-wackestone that exhibits subaerial expo- grainstone-packstone, but brachiopod skeletal-peloid
sure features and are commonly interbedded with grainstone-packstone is dominant.
thin volcanic ash layers. These lithofacies are over- Brachiopod skeletal-peloid grainstone-packstone is
lain by brachiopod and/or crinoid skeletal-peloid interpreted to represent the continued flooding of the
grainstone-packstone, which are, in turn, overlain by platform in relative low-energy conditions. Thick accu-
poorly sorted skeletal-peloid grainstone-packstone mulation of such deposits could indicate deeper water
and/or well-sorted skeletal grainstone. This succes- or ongoing restriction caused by emergent outer plat-
sion is, in turn, overlain by coated-grain grainstone- form shoals or barriers. The facies is initiated during ini-
rudstone and/or ooid grainstone and locally a skeletal tial flooding of the platform and is most robust during
intraclast grainstone or algal grainstone. Obviously, times of maximum flooding. Brachiopod skeletal-peloid
deviations from this generalization exist, and cycle grainstone-packstone is poorly sorted, composed of large,
thickness, stacking patterns, and relative contribution commonly in growth position, thick-walled brachio-
of each facies type vary within and across the major pods, variable crinoids, diverse algal fragments and
sequences. foraminifera, and/or corals (Figure 7). Brachiopod
FIGURE 6. (A) Core photos from well T-5246 show peloid packstone-wackestone (PPW) alternating with volcanic ash layers (VA) below the Lvis_SSB. The boundary
that shows evidence for subaerial exposure is overlain by poorly sorted skeletal-peloid grainstone-packstone (SPP) with elevated bitumen cement in the lower
70 cm (27 in.) (see Table 2 and Figure 5 for additional lithofacies information). Photomicrographs show the range of PPW textures. Common skeletal grains are
tubular algal fragments (B, D), calcispheres (B, E), and minor coated skeletal grains and aggregrate grains (C, D). Width of photomicrographs is 4.20 and 1.58 mm
(0.16 and 0.062 in.) (B, C – E, respectively).
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 21
22 / Kenter et al.

FIGURE 7. (A) Core photos from well T-5246 show shoaling succession of skeletal-peloid grainstone-packstone (SSP) and well-sorted skeletal grainstone (SPG)
immediately below the Lvis_9.0_csb. The boundary is overlain by peloid packstone-wackestone (PPW) and a thick interval of brachiopod-skeletal-peloid
grainstone-packstone (BSP). Concentrations of in-situ brachiopods and local development of intraparticle and shelter porosity are clearly visible in the BSP
interval. Photomicrographs (B – E) show the generally poorly sorted texture and microfractured nature of BSP. Brachiopods are dominant constituents and range
from small thin-walled (less than 2 cm [0.8 in.] long and 1 mm [0.04 in.] thick) to large and thick-walled species (respectively, 5– 15 cm [1.9– 5.9 in.] and 1– 10 mm
[0.04 and 0.40 in.]). Note the sutured, stylotized contacts of the brachiopods. Other common grain types are crinoid ossicles, peloids, algal fragments, and
benthic foraminifera. Width of photomicrographs is 4.20 mm (0.16 in.).
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 23

skeletal-peloid grainstone-packstone intervals are gen- which are interpreted as high-energy shallow subtidal
erally continuous across the platform, appear to thin open-platform to intertidal deposits (Figures 11, 12,
toward the outer platform, and have the highest fre- respectively). Ooids are generally small, ranging from
quency and volume in the upper part of the Visean A 0.5 to 1.0 mm (0.02 to 0.04 in.), and were most prob-
(between Lvis_13 and Lvis_10.5), where they generally ably calcitic in origin from their radial structure.
form the basal parts of most cycles (Figure 5). Intervals Coated grains are commonly skeletal fragments with
of brachiopod skeletal-peloid grainstone-packstone coatings that are micritic and mainly constructional.
occur sporadically in the lower part of the Visean A Coated-grain grainstone has uncommon occurrences
platform. Two very distinct and thick-bedded (4–10-m; in the Visean A but is common in the Serpukhovian
13–33-ft) intervals, one right above the Lvis9_csb and reservoir (Figure 5), where it is concentrated in the
one below the Lvis13_csb, blanket the entire platform. northern part of the platform amplifying a general
Crinoid skeletal-peloid grainstone-packstone is coarse trend from higher energy in the north to lower en-
and poorly to moderately sorted crinoid-dominated ergy (and/or open marine) toward the south. Coated-
skeletal-peloidal sand. The crinoids commonly have grain and/or ooid grainstones are locally overlain by a
destructionally micritized outer surfaces; other com- relatively thin interval of coarse-grained lithoclastic
mon components are mostly broken brachiopods, green rudstone-grainstone, skeletal-intraclast grainstone,
and red algal fragments, and foraminifera (Figure 8). with centimeter-size subangular to well-rounded intra-
Crinoid skeletal-peloid grainstone-packstone inter- clasts and Chaetetes (demosponge) fragments, which
vals are generally not continuous across the platform are commonly coated by micrite (Figure 13). Some in-
and seem to preferentially occur near the platform tervals display clast-supported flat pebbles, indicative
margin (wells T-6846, T6246, T-8, and T-44). Crinoid of beach environments. Dominant grain types range
skeletal-peloid grainstone-packstone is interpreted as from coated grains to ooids and algal grains.
being deposited in moderate to low energy below or
around wave base and may represent the time of maxi- Algal Grainstone (AG)
mum flooding. These intervals represent the maxi- Finally, very well-sorted and cross-bedded, silt to
mum open-marine influx during the cycle, and only fine sand-size, algal grainstone occurs near the top of
one interval (right above Lvis_10.5) covers the entire the Bashkirian sequences and locally in the Visean A
platform (Figure 5). sequence (Figures 5, 14). Algal grainstone intervals
were also observed in the outermost platform (Serpu-
Poorly Sorted Skeletal-peloid Grainstone-packstone khovian) of the T-4556 well and may indicate the oc-
(SPP) and Well-sorted Skeletal Grainstone (SPG) currence of this facies in a high-energy outer platform
Poorly sorted skeletal-peloid grainstone-packstone setting. Elsewhere on the platform, algal grainstone
and well-sorted skeletal grainstone generally occur alternates with more open-marine well-sorted skel-
above brachiopod skeletal-peloid grainstone-packstone etal grainstone or coated-grain grainstone-rudstone;
and crinoid skeletal-peloid grainstone-packstone, exhibit in these instances, algal grainstone is interpreted as
progressively better sorting and less mud, and therefore being deposited in the intertidal zone near the top of
have been interpreted as shallow-water wave-agitated cycles, perhaps sourced from the outer platform and
sand shoal deposits (Figures 9, 10, respectively). These filling the remaining accommodation space.
grainstone lithofacies dominate the cycles for most of
the Visean A reservoir and commonly show lateral Lithofacies Mosaics
shifts from one facies into the other. In the T-5246 well, Figure 15 shows hypothetical depositional models
well-sorted skeletal grainstone dominates the cycles, of higher order depositional cycles for both the Bash-
and there seems to be a general subtle increase away kirian sequence (Figure 15A) as well as the Visean A
from the central platform (Figure 5). Immediately above and Serpukhovian sequences (Figure 15B). The mod-
the Lvis2_MFS, a southward thinning interval of well- els show grain compositional and textural trends from
sorted skeletal grainstone covers the entire platform. the central platform through the outer platform and
into the upper flank and indicate some associated
Coated-Grain Grainstone (CgGR), Ooid Grainstone, bathymetric relief during the Visean A and Serpu-
(OG), and Skeletal-intraclast Grainstone (SIG) khovian. In the central platform, Visean A and Serpu-
In a complete cycle, the previously described litho- khovian platform cycles are vertically and laterally
facies are overlain by cross-bedded or uniformly lami- relatively predictable and show a succession, from
nated, well-sorted coated-grain and/or ooid grainstone, base to top, that reflects depositional energy. Near
24 / Kenter et al.

FIGURE 8. (A) Core photos from well T-5246 show succession of well-sorted coated-grain grainstone (CgGR) overlain by peloid packstone-wackestone (PPW)
and capped by the Lvis_10.5_csb. The boundary is, in turn, overlain by crinoid skeletal-peloid grainstone-packstone (CSP) and poorly sorted skeletal-peloid
grainstone-packstone (SPP) (see Table 2 and Figure 5 for additional lithofacies information). The CSP interval is only moderately stained by bitumen here in
T-5246, whereas toward the outer platform, it is commonly completely stained black. Crinoid skeletal-peloid grainstone-packstone is coarse and poorly to
moderately sorted crinoid-dominated skeletal-peloidal sands (B) with destructionally micritized outer surfaces (C) and common components as (mostly broken)
brachiopods, (green and red) algal fragments and foraminifera (B – E). Common grain types are crinoid ossicles, brachiopod hash and whole valves, peloids, algal
fragments (notably Koninckopora), benthic foraminifera. Width of photomicrographs is 4.20 mm (0.16 in.).
FIGURE 9. (A) Core photos from well T-5246 show a thick interval of poorly sorted skeletal-peloid grainstone-packstone (SPP) at the top. The lower half of the core
contains a shoaling succession of well-sorted skeletal grainstone (SPG) capped by Lvis_9.5_csb. The boundary is overlain by peloid packstone-wackestone (PPW)
with a thin volcanic ash ( VA) interbed near the top (see Table 2 and Figure 5 for additional lithofacies information). The skeletal-peloid grainstone-packstone (SSP) and
well-sorted skeletal grainstone (SPG) are generally located above the B- and CSP; because they exhibit progressively better sorting and less mud, they have been
interpreted as shallow-water wave-agitated sand shoal deposits. Common grain types found in SPP shown in (B – E) are crinoid ossicles, brachiopod hash, peloids, algal
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 25

fragments, and benthic foraminifera. Grains are commonly coated by thin micrite rims and are poorly sorted. Width of photomicrographs is 4.20 mm (0.16 in.).
26 / Kenter et al.

FIGURE 10. (A) Core photos from well T-5246 show shoaling succession of well-sorted skeletal grainstone (SPG) capped by Lvis_10.0_csb. The boundary is overlain
by thin peloid packstone-wackestone (PPW), well-sorted skeletal grainstone (SPG), and peloid packstone-wackestone (PPW), with large thin-walled brachiopods
(see Table 2 and Figure 5 for legend). Well-sorted skeletal grainstones (SPG) are generally located above the BSP and CSP lithofacies, exhibit better sorting and less
mud, and have been interpreted as shallow-water wave-agitated sand shoal deposits. Photomicrographs (B – E) show common grain types of SPG: crinoid ossicles,
brachiopod hash, peloids, algal fragments, and benthic foraminifera. Width of photomicrographs is 4.20 mm (0.16 in.) (B, C, E) and 1.58 mm (0.06 in.) (D).
FIGURE 11. (A) Core photos from T-5246 show shoaling succession of a thick interval of cross-beds, but laminations are obscured by bioturbation, well-sorted
coated-grain grainstone (CgGR), which is overlain by a thin peloid packstone-wackestone (PPW). The Lvis_10.5_csb is overlain by coarse and poorly to moderately
sorted crinoid-dominated skeletal-peloidal sands (CSP) (see Table 2 and Figure 5 for legend). In a complete cycle, mixed skeletal-peloidal sands (mostly SPG) are
overlain by cross-bedded or uniformly laminated, well-sorted coated-grain and/or ooid grainstone (CgGR and OG), interpreted as high-energy shallow subtidal
open platform to intertidal deposits. Common grain types in CgGR shown in (B – E) are well-sorted peloids, coated skeletal grains, algal fragments (tubular), and
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 27

benthic foraminifera; crinoid ossicles and brachiopod hash are minor grains. Width of photomicrographs is 4.20 mm (0.16 in.) (D) and 1.58 mm (0.06 in.) (B, C, E).
28 / Kenter et al.

FIGURE 12. (A) Core photos from T-5246 show part of an approximately 5 – 6-m (16 – 19-ft)-thick interval of cross-bedded, well-sorted ooid grainstone (OG). An
exposure surface, with small intraclasts, separates a dark, bitumen-rich interval above from the underlying better cemented, yellow-cream-colored interval
in the rightmost core slab (see also Table 2 and Figure 5 for legend). In the Bashkirian, a complete cycle of mixed skeletal-peloidal sands (mostly SPG) is generally
overlain by cross-bedded ooid grainstone lithofacies (OG) interpreted as high-energy shallow subtidal open platform to intertidal deposits. Photomicrographs
(B – E) show the dominant grain types of ooids and coated skeletal grains; algal fragments (tubular) and benthic foraminifera also occur. Width of photomicrographs
is 4.20 mm (0.16 in.) (B, D) and 1.58 mm (0.06 in.) (C, E).
FIGURE 13. (A) Core photos from T-5246 show a shoaling succession of well-sorted skeletal grainstone (SPG), overlain by a relatively thin interval of coarse-
grained skeletal-intraclast grainstone (SIG), and capped by the Serp1_csb. The boundary is overlain by thin intervals of PPW and SPG, brachiopod-skeletal-peloid
grainstone-packstone (BSP), and another interval of SPG (see Table 2 and Figure 5 for legend). (B– D) Skeletal-intraclast grainstone (SIG) typically has centimeter-
size subangular to well-rounded intraclasts and Chaetetes fragments, which are commonly coated by micrite. Other grains are coated grains, ooids, and algal fragments.
Width of photomicrographs is 4.20 mm (0.16 in.).
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 29
30 / Kenter et al.

FIGURE 14. (A) Core photos from T-5246 show a thick interval of very well-sorted and cross-bedded, silt- to fine sand-size algal grainstone (AG) underlain
and overlain by well-sorted coated-grain grainstone (CgGR); an interbed of volcanic ash (VA) occurs at the bottom (see Table 2 and Figure 5 for legend).
Photomicrographs (B-E) show dominant grain types of AG: algal fragments, peloids, and uncommon benthic foraminifera. Typical algal species are Donezella
(B, C), paleoberesellids such as Kamaena (D), and beresellids such as Beresella (E). Paleoberesellids are typically Visean and beresellids and Donezella occur in the
Serpukhovian to Bashkirian. Width of photomicrographs is 4.20 and 1.58 mm (0.16 and 0.06 in.) (B, C-E, respectively).
FIGURE 15. Models of hypothetical higher order depositional cycles for the Bashkirian (A) and the Visean A – Serphuhovian (B) showing grain composition and
textural trends from the central Tengiz platform through the outer platform and into the flank. TST = transgressive systems tract; HST = highstand systems tract.
See text for discussion.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 31
32 / Kenter et al.

the base, a thin, peloid packstone-wackestone that trast, the Bashkirian sequence shows a flat-topped plat-
exhibits subaerial exposure features is commonly in- form with parallel facies zones dominated by higher
terbedded with thin volcanic ash layers. These litho- energy grainy lithofacies types. High-frequency cycles
facies are overlain by brachiopod bioherms and/or are much thinner in the Bashkirian, at meter scale,
crinoid skeletal-peloid grainstone-packstone, which and show mostly alternating coated grain grainstone,
are, in turn, overlain by poorly sorted skeletal-peloid ooid grainstone, and algal grainstone with sedimen-
grainstone-packstone and/or well-sorted skeletal grain- tary structures. The Bashkirian rim was probably flat
stone. These lithofacies reflect lowstand and early during deposition and shows thinning of platform
transgressive restriction providing the environment facies that border a nearly starved slope system.
for trapping volcanic ash fallout (peloid packstone-
wackestone) and maximum flooding during the cri-
noid skeletal-peloid grainstone-packstone. This succes- Visean A and Serpukhovian:
sion is succeeded by highstand deposits that indicate Trends and Depositional Environments
shoaling to (nearly) sea level: coated-grain grainstone- Visean A cycles consist almost entirely of grainy
rudstone and/or ooid grainstone and locally a skeletal facies (packstone and grainstone) that feature depo-
intraclast grainstone or algal grainstone. In the outer sitional environments ranging from restricted low-
platform toward the east, cycles lack the lagoonal base energy brachiopod skeletal-peloid grainstone-packstone
and have a microbial boundstone interval (type 3 micro- to moderate- to low-energy deeper subtidal envi-
bial boundstone; see Collins et al., 2006) instead, which ronments characterized by more poorly sorted and
is superceded by thicker poorly sorted mixed skeletal- open-marine poorly sorted skeletal-peloid grainstone-
peloidal and crinoid sand bodies. Some cycles include packstone as well as crinoid skeletal-peloid grainstone-
thin deposits of high-energy coated-grain grainstone- packstone and high-energy well-sorted grainstone shoals
rudstone and/or ooid grainstone and skeletal intra- (well-sorted skeletal grainstone and uncommon coated-
clast grainstone or algal grainstone. Brachiopod bio- grain grainstone-rudstone).
herms, crinoid skeletal-peloid grainstone-packstone, General trends in lithofacies distribution, cycle
and poorly sorted skeletal-peloid grainstone-packstone thickness, and stacking suggests the presence of three
and/or well-sorted skeletal grainstone are absent. intervals in the Visean A and Serpukhovian reservoir
The overlying Bashkirian cycles (Figure 15A) are zones, each displaying certain depositional trends
thinner and, in the central platform, show alternating (Figure 5). These are the intervals between the Lvis_SSB
lagoonal peloid packstone-wackestone with subaerial and Lvis10.5_csb, Lvis10.5_csb and Lvis13_csb (which
exposure features and interbedded with thin volcanic straddles the second-order Lvis2_MFS), and Lvis13_csb
ash layers, coated-grain grainstone-rudstone and/ to the Serp_SSB.
or ooid grainstone and skeletal intraclast grainstone. In the Lvis_SSB to Lvis10.5_csb interval, the plat-
Above the Bash2.5_MFS, coated-grain grainstone- form cycles are dominated by the poorly sorted, open-
rudstone and/or ooid grainstone alternate with algal marine poorly sorted skeletal-peloid grainstone-
grainstone intervals, and volcanic ash beds are nearly packstone, with minor intercalations of brachiopod
absent. However, toward the outer platform areas, low- skeletal-peloid grainstone-packstone and/or crinoid
energy transgressive deposits (brachiopod skeletal- skeletal-peloid grainstone-packstone. Crinoid skeletal-
peloid grainstone-packstone, crinoid skeletal-peloid peloid grainstone-packstone intercalations generally
grainstone-packstone, poorly sorted skeletal-peloid occur near the base of cycles, whereas well-sorted skel-
grainstone-packstone) thicken and may even include etal grainstone intercalations appear near the top.
microbial boundstone intervals (T-4556), while high- Cycle caps are thin, decimeter-thick, micropeloidal
energy cycle tops (coated-grain grainstone-rudstone, packstone-mudstone intervals that have, in nearly all
ooid grainstone) are thinning or absent (T-6743). cases, one or two thin volcanic ash interbeds, but
Obviously, deviations from the generalizations of generally display moderate subaerial exposure. Cycle
Figure 15 exist, as will be discussed in the following thickness ranges from 1–2 m (3.3–6.6 ft) near the
sections. Cycle thickness, stacking patterns, and relative base to more than 15 m (49 ft) in the middle of the
contribution of each facies type vary within and across interval with an average of 5–8 m (16– 26 ft). Poorly
the major sequences. As an example, the bathymetric sorted skeletal-peloid grainstone-packstone intervals
profile changes toward the Visean A–Serphukhovian are dominant in the central platform around wells
outer platform to a deeper setting that brings microbial T-220 and T-5848 but gradually are partially replaced
boundstone on the platform during flooding. In con- by higher energy sand shoals (well-sorted skeletal
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 33

grainstone) toward the outer platform areas (well exposure toward the Serp_SSB. These observations
T-5246 to the north and wells T-6246 and T-6846 to suggest the development of a slightly deeper and lower
the southwest). Crinoid-rich intervals represent maxi- energy outer platform toward the southeastern mar-
mum, open-marine flooding and the deepest litho- gin (possibly caused by structural tilting). The thin-
facies. These occur preferentially in the outer platform ning of cycles, increasing exposure of cycle tops, and
toward the southwest and seem to thin and pinch out lateral facies differentiation during the Serpukhovian
toward the central platform. One interval, however, coincide with the development of an unusual deeper
runs contrary to this trend and appears to blanket water outer platform toward the northeastern to east-
the entire platform at the top of the Lvis_SSB to ern margins of the Tengiz platform, where upper
Lvis10.5_csb interval. Similarly, one nearly 10-m (33-ft)- slope facies consisting of algal-microbial, skeletal-
thick and very continuous brachiopod skeletal-peloid microbial, and microbial boundstone prograded sev-
grainstone-packstone interval covers the platform right eral kilometers out into the basin on a preexisting
above the Lvis9_csb boundary. This general deposi- Visean A and B shelf edge. In these regions of the outer
tional picture suggests an overall low-energy platform, platform, which generally have lower energy and
subtidal or below wave base, with slightly shallower probably deeper water facies, exposure expression is
outer platform sand shoals and a dominant southern reduced or absent (see Figure 15).
open-marine influence. The cycles in this succession
never filled the total accommodation space, despite
the relatively low-amplitude sea level oscillations. Bashkirian: Trends and
The Lvis10.5_csb to Lvis13_csb interval, which strad- Depositional Environments
dles the supersequence flooding surface (Lvis2_MFS), Bashkirian cycles are thinner and less systematic
shows a similar central versus outer platform facies than those recognized in the Visean A and Serpu-
pattern as the previous interval but has much thinner, khovian, most likely because of rapid sea level fluc-
2–5-m (6.6–16-ft)-thick cycles and frequent brachio- tuations produced during global icehouse conditions.
pod skeletal-peloid grainstone-packstone intervals. The Bashkirian consists of three intervals, of which
These brachiopod interbeds seem to thin away from the lower (Serp_SSB to Bash1_csb) is essentially a 15–
the platform center and are replaced by higher energy 20-m (49–66-ft)-thick ooid sand body covering the
mixed skeletal-peloidal sands to the south and high- entire platform (Figure 5). This unit is nearly tabular in
energy, well-sorted sands to the north. In addition, shape, has locally thin, mixed skeletal-peloidal open-
the section has a crinoid-rich base, a thick interval of marine intercalations, and is cyclic with thin mud-
high-energy sand shoals in the middle, and a thick stone cycle caps and associated exposure surfaces. In
brachiopod interval (more than 10 m) near the top. the central platform, the lower half of the interval
Each of these intervals is continuous across the entire commonly contains high bitumen content and has
platform. The crinoid-dominated base indicates max- low porosity; toward the southwestern outer platform
imum deepening of the platform coinciding with the (T-6246) and north (T-5246), nearly the entire section
Lvis_MFS surface. Although relative depositional con- has bitumen and strongly reduced porosity. The mid-
ditions did not change and accommodation space dle part of the Bashkirian is bounded above by a
was still underfilled, it suggests that sea level fluctua- stratigraphic surface (Bash2.5_MFS) that was not part
tions increased in frequency. of the original framework published by Weber et al.
The Lvis13_csb to the Serp_SSB interval in the HSS (2003). The Bash1_csb to Bash2.5_MFS is dominated
of the supersequence has similar thin cycles, 2– 5 m by thin (a few meters) shoaling cycles of sorted skel-
(6.6– 16 ft) in thickness, but those show more pro- etal subtidal open-marine sands, overlain by high-
nounced subaerial exposure near the cycle tops. Fur- energy intertidal coated grain to ooid sand shoals,
thermore, there is an overall trend of higher energy and capped by thin peloidal packstone-mudstone in-
coated-grain and ooid sand shoal deposits with ele- tervals with general extensive subaerial exposure and
vated amounts of dispersed volcanic ash toward the intercalated centimeter- to decimeter-thick volcanic
north (wells T-5447, T-5044, and T-5444) and slightly ash beds. In the uppermost part of the Bashkirian
deeper, lower energy, environments toward the south, (Bash2.5_MFS to Bash_SSB), ooid to coated grain sands
near the outer platform (wells T-6246 and T-6846). occur in the basal part of shoaling cycles and are
Cycles were nearly completely filled in the northern overlain by silt-size (high-angle), cross-bedded, well-
platform area, multiple ash layers occur near sequence sorted algal sands. Lagoonal packstone-mudstone inter-
boundaries, and there is an increased expression of vals and associated volcanic ash beds are uncommon or
34 / Kenter et al.

lacking completely. The algal fragments are silt-size, and Lvis_SSB, a nearly 5-m (16-ft) thicker section be-
dominant, and probably tubular red algae in origin. tween the Lvis13_csb and Lvis10.5_csb, probable minor
The waning of ooid domination and introduction erosion at the Lvis13_csb, and several additional meters
of algal silt above the Bash2.5_MFS may indicate a missing section below the Serp_SSB. Although not
general deepening stratigraphic trend toward the proven, these differences may indicate that the south-
Bash_SSB, which is consistent with an interpretation western margin tilted downward some 10 m (33 ft) or
of the subsequent drowning of the platform (Weber more prior to the Serp_SSB and was eroded at the time
et al., 2003). Cross-bedding was possibly generated by of the Serp_SSB following minor tilting to the north.
oceanic currents sweeping the deeper water platform
top during the progressively longer highstands; the RESERVOIR QUALITY AND DIAGENESIS
platform was still exposed during shorter lowstands
associated with high-frequency and high-amplitude Pore Types
(as much as 40–60-m [131–196-ft]) sea level changes Air permeability data from approximately 11,000
of the icehouse world. Expressions of subaerial ex- plugs clearly suggest the presence of three general
posure, such as dissolution, recrystallization, and pres- reservoir environments based on plug-scale matrix
ervation of ash layers are also nonuniform from cycle properties: central platform, outer platform, and rim
to cycle. Despite this, Bashkirian cycles retain some slope (flank) (Collins et al., 2006). The Visean A, Ser-
similarities to Visean A and Serpukhovian cycles, such pukhovian, and Bashkirian central platform reservoir
as the presence of high-energy shoals and deeper can be regarded as having distinctly different gross
platform skeletal grainstone and the association of reservoir properties that are generally well behaved as
semirestricted to restricted facies near sequence bound- a whole. Porosity-permeability crossplots and poros-
aries. In the Bashkirian, high-energy shoals are rep- ity histograms suggest different behavior between
resented by oolites, pisolites, and coated grains, whereas the northern wells (T-220 and T-5246) and southern
the deeper and restricted facies are similar to those wells (T-6246 and T-6846) in the central platform
present in the Serpukhovian and Visean A. (Figure 16). Trends of generally lower mean porosi-
ties and better behaved porosity-permeability relation-
Interplay between Minor Tilting ships occur in the southern wells, whereas bimodal
and Deposition porosity distributions occur in the northern wells.
A remarkable observation from the cross section in Pore-type characterization of nearly 2000 plugs (and
Figure 5 is that the railroad-track pattern of sequence associated thin sections) from the nearly bitumen-
and cycle boundaries below the Serp_SSB is nearly free T-220 well and the moderately bitumen-cemented
perfectly parallel, with the exception of the section in T-6246 well show the presence of varied pore types
T-6846. Another observation is the presence of a minor like moldic, vugs, microporosity, interparticle poros-
angular unconformity of several degrees between the ity, intercrystalline, and intraparticle porosity and
Serp_SSB and the underlying cycle breaks, wherein the general absence of fractures (Figure 17).
both the Serp_SSB and cycle breaks dip to the south a The porosity classification scheme of Choquette
few degrees. In addition, the Serp_SSB appears to have and Pray (1970) defined numerous carbonate pore
variable erosional relief of as much as 5 m (16 ft) types, either fabric selective (including interparticle,
across the platform as indicated by the apparent in- intraparticle, intercrystalline, and moldic) or not fab-
creased erosion in well T-5447. These observations sug- ric selective (including fracture and vug). There were
gest several things. First, fairly gradual and constant implications for porosity-permeability relationships,
vertical aggradation across the central platform re- of course, but this was not emphasized. Lucia’s (1995,
sulted in a flat platform top with minimal local relief as 1999) landmark study showing the direct link between
a result of immediate compensation of possible varia- rock type, porosity type, and reservoir quality (porosity-
tions in depositional topography. In addition, the plat- permeability relationships or transforms) changed the
form was tilted to the south prior to the Serp_SSB, sub- context of some of these classic pore types. Lucia (1995,
sequently eroded (based on the missing Bogdanovsky 1999) recognized interparticle pore types and vuggy
and Suryansky biozones), generating as much as 5-m pore types. Interparticle pores, which have a more
(16-ft) relief, and finally, tilted a few more degrees straightforward porosity-permeability transform that
along the same axis during the post-Bashkirian. Well can be tied to rock type, grain size, and sorting, are
T-6846 shows about a 5-m (16-ft) thicker section rela- instances where the pore space occurs between par-
tive to other wells of Figure 5 between the Lvis10.5_csb ticles and is not significantly larger that the particles.
FIGURE 16. Subtle porosity-permeability differences are observed between four wells in the central platform; see Figure 4 for well locations and Figure 5 for
their facies and stratigraphy. K-Phi plots (top) and porosity histograms (bottom) suggest different behavior between the northern wells (T-220 and T-5246) and
southern wells (T-6246 and T-6846) with generally lower mean porosities and better behaved permeability-porosity relationships in the southern wells and
bimodal porosity distributions in the northern wells.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 35
36 / Kenter et al.

FIGURE 17. Pore and cement types recognized in the T-220 and T-6246 wells are generally similar. However, relative
contributions by vuggy, intercrystalline, and moldic pore types are significantly higher in T-220 for all facies types
than in T-6246, which may be related to the larger average grain size and higher contribution by the mud-lean rock
types in T-6246. Similarly, cement types in T-6246 show a higher abundance (frequency) of equant calcite and syntaxial
cements than in T-220. This may suggest a larger volume of deep burial cements filling pores at T-6246, but clearly, more
quantitative data are required for such assessment. Pore types across the bottom of the figures are 1 = interparticle;
2 = intraparticle; 3 = fenestral; 4 = shelter; 5 = growth framework; 6 = intercrystalline; 7 = moldic; 8 = microporosity; 9 = fracture;
10 = channel; 11 = vug; and 12 = enhanced dissolution. Cement types are 1 = fibrous; 2 = bladed; 3 = equant to rhombic;
4 = coarse crystalline; 5 = botryoidal; 6 = syntaxial; 7 = micritic; 8 = meniscus; 9 = microstalactitic; 10 = poikilotopic; 11 =
neomorphic spar; and 12 = grain rims. Plotted are relative dominance of pore or cement types in terms of visual presence
with the number indicating their ranking. See text for discussion.

Vuggy pore types are typically larger than the grains or be modified. Clearly, real-world reservoirs in lime-
occur within the grains, and these can be separate or stone can be even more complex with the addition of
touching. Separate vugs are connected through the in- microfractures, intercrystalline porosity, or micropo-
terparticle pore system, whereas touching vugs them- rosity connecting separate vugs, and/or the problem
selves form an interconnected system. of assessing in two dimensions with petrography the
At Tengiz, separate vugs that add to porosity but three-dimensional complexity of porosity. The fol-
not permeability include moldic and intraparticle lowing sections describe the pore types commonly
porosity, so these must be subtracted from total po- observed in Tengiz samples and their relative abun-
rosity to use interparticle transforms. Touching pore dance in the central platform.
types include some vugs (solution-enlarged interpar- Pore types were identified in the Tengiz cores and
ticle pores) and fractures, which add porosity and thin sections generally following the terminology
permeability independently from the interparticle of Choquette and Pray (1970) to first assess their
component, such that interparticle transforms must relation to lithofacies and then explore the relation
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 37

FIGURE 18. Photomicrographs illus-


trating the dominant pore types in
the central platform area. (A) Vugs
are nonfabric-selective pores com-
monly larger than the dominant grain
size (although different interpreta-
tions exist in the literature). Examples
of vugs shown here are the result of
enhanced dissolution in a grainstone
following early equant rim cementa-
tion. (B) Vugs developed in a pack-
stone-grainstone, dissolving both
grains and matrix but leaving crinoid
ossicles and brachiopod fragments
intact. (C) Overpacked ooid grain-
stone with vugs that postdate early
cementation but predate blocky cal-
cite and bitumen emplacement. Vugs
are a dominant pore type in the
central platform and seem to be
mostly related to deep-burial disso-
lution instead of initial meteoric pro-
cesses. (D) Advanced stage of en-
hanced dissolution of moldic and
interparticle porosity (and minor in-
traparticle) leading to the destruction
of cement bridges and development
of vugs. (E) Skeletal and coated grain
grainstone dominated by interparti-
cle porosity with minor cement fill.
(F) Overpacked ooid grainstone with
minor blocky calcite and bitumen
occluding solution-enhanced inter-
particle porosity. (G) Skeletal and
peloid grainstone with rims of equant
spar lining interparticle pores and
minor moldic porosity. (H) An exam-
ple where the interparticle porosity is
completely occluded by equant to
rhombic spar and bitumen cement.
Width of photomicrographs is 4.20
and 1.58 mm (0.16 and 0.06 mm)
(A – G, H, respectively). See text for
discussion.

between lithofacies and reservoir quality (Table 2). forms (Lucia, 1995, 1999). Figure 19C – E shows suc-
Common pore types in the Tengiz central platform are cessive stages of the development of moldic porosity
vugs, interparticle, microporosity, intercrystalline, at Tengiz: in its early phase where micrite grains are
moldic, and intraparticle with generally the highest gradually being dissolved following partial fill of in-
permeability occurring in vuggy-dominated matrix. terparticle porosity by equant to rhombic spar ce-
The most common pore type is vuggy, which is, in part, ment (Figure 19C); a phase of progressive dissolution
fabric selective, being found in matrix, grainy areas shown by corrosion of the mold boundaries, nearly
(solution-enhanced interparticle pores), or associated breaking through at contacts and corroding earlier
with stylolites (Figure 18A–C). The second most common cement (Figure 19D); and an advanced stage where
pore type is interparticle porosity (Figure 18D–H), enhanced dissolution left only micrite envelopes. In-
followed by microporosity (Figure 19A, B). traparticle porosity, which is also included as vugs in
Vuggy porosity includes moldic porosity when terms of porosity-permeability transforms, is shown
considered in terms of porosity-permeability trans- in Figure 19D–F.
38 / Kenter et al.

FIGURE 19. Photomicrographs illus-


trating additional dominant pore types
in the central platform area. (A, B)
Examples of microporosity, pore size
smaller than 25 Mm, developed in
grains (A) and matrix (B). This type of
pore is generally difficult to observe
in thin sections because of their thick-
ness, commonly about 25 – 45 Mm.
(C) Moldic porosity in its early phase:
micrite grains are gradually being
dissolved following partial fill of inter-
particle porosity by equant to rhombic
spar cement. (D) Moldic porosity
showing evidence for enhanced disso-
lution shown by corrosion of the mold
boundaries, nearly breaking through
at contacts and corroding earlier ce-
ment. (E) Moldic porosity showing
further stage of enhanced dissolution
leaving only micrite envelopes. Bitumen-
stained micrite appears to slow down
dissolution. Interparticle porosity is
partially filled with calcite cement.
(F) Intraparticle porosity developed
in benthic foraminifera. (G) Intraparticle
porosity developed in a green alga,
Koninckopora. (H) Intraparticle porosity
developed in a Chaetetes fragment (not
all pores are filled by blue epoxy). Width
of photomicrographs is 4.20 mm (0.16 in.)
(A, B, F, G) and 1.58 mm (0.06 in.)
(C, D, E, H). See text for discussion.

generally beyond normal petro-


graphic resolution and, therefore, dif-
ficult to distinguish. Fractures and
fissures are generally uncommon in
the central platform, and where pre-
sent, they are mostly occluded by
cement. In some few cases where the
fractures remain open, they range
from microhairline fissures to sharp
Examples of an advanced stage of dissolution of microfractures (Figure 20A–F). Stylolites are generally
moldic and interparticle porosity leading to the de- stained by bitumen and commonly associated with
struction of cement bridges and development of vugs small vugs (Figure 21A – C). Examples demonstrate
with minor blocky calcite and bitumen occluding that such dissolution postdates stylolite formation.
porosity are shown in Figure 18D and F. Commonly,
moldic porosity co-occurs with interparticle poros- Parameters Controlling Porosity
ity as shown in Figure 18G; interparticle porosity is and Permeability
completely occluded by equant to rhombic spar and The relative contributions by vuggy, moldic, and
bitumen in some samples (Figure 18H). Microporos- intercrystalline porosity are significantly higher for
ity in grains or matrix is difficult to observe in Tengiz all facies types in well T-220 than in well T-6246. In
samples because pore sizes are much smaller than the contrast, the relative contribution by interparticle po-
thickness of a typical thin section. Microporosity in- rosity is higher at T-6246, a change that is most likely
cludes fine intercrystalline porosity because both are related to the higher percentage of grainy, mud-lean
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 39

FIGURE 20. Photomicrographs illus-


trating dominant pore types in the
central platform area. (A, B) Generat-
ing thin sections from core plugs
with open fractures is difficult if not
impossible. Shown here are examples
of thin fractures that are filled by
blocky calcite cement. (C, D) Examples
of hairline fissures or fractures filled
by bitumen. Note that the fissure in
(C) postdates blocky calcite interparti-
cle cement fill but predates bitumen
emplacement and confirms the very
late timing of bitumen invasion.
(E, F) An exception to the rule: vague
outline of open fractures that have
been enhanced by late-stage dissolu-
tion in a skeletal-peloidal grainstone
with high interparticle porosity. Frac-
tures may connect vugs as shown
in these samples. Width of photomi-
crographs is 4.20 mm (0.16 in.). See
text for discussion.

ooid grainstone, skeletal intra-


clast grainstone, and algal grain-
stone) that make up the cyclic
pattern shown in Figure 5 do show
subtle but significant differences
(Figure 22). Pore type contribu-
tions in the grainy facies groups
are very similar, but muddy la-
rock types in that well (Figure 22). Nevertheless, at a goonal facies have a higher contribution of intra-
gross scale, none of the individual pore types or com- particle and fracture porosity. The grainy facies in
bination of pore types seem to have a significant first- well T-6246 have a higher contribution of interpar-
order relationship with primary depositional facies ticle porosity but also differ in the overall distribu-
nor have any significant control on porosity or per- tion of porosity. These distributions are remarkably
meability. This suggests that the central platform area different between the two wells with bimodal and
could be regarded as one single reservoir region wherein generally higher porosities in grainy facies and low
changing pore types and lithofacies do not significantly porosities in lagoonal facies in well T-220, whereas
impact true reservoir connectivity. This suggestion is the porosity distributions in well T-6246 show highly
rather surprising and seemingly contradicts Lucia’s (1995, reduced and flattened distributions for the mud-lean
1999) reservoir rock-type classification that suggests facies compared to T-220 (Figure 23). Similarly, the
pore types control permeability behavior. mud-rich grainy facies have reduced porosity, but the
Although the observations above suggest minimal lagoonal mudstone has higher porosity, displaying
first-order control by pore type or facies, several second- a bimodal distribution that is possibly related to a
order relationships should be acknowledged. Porosity higher contribution of fracture porosity. The above
distributions and pore-type contributions for the suggests that T-220 porosity has a relatively higher
muddy lagoonal facies (peloid packstone-wackestone), contribution by pore types that are the result of cor-
moderately sorted grainy facies with mud matrix rosive processes, like vug and moldic porosity, an ob-
(brachiopod skeletal-peloid grainstone-packstone, servation that can be made for T-5246 as well.
crinoid skeletal-peloid grainstone-packstone, and In addition to pore type, observations were made
poorly sorted skeletal-peloid grainstone-packstone) on the type of pore-filling agents, which resulted in
and mud-lean well-sorted grainy facies (well-sorted two significant trends. First, the distribution of pore-
skeletal grainstone, coated-grain grainstone-rudstone, filling cement types for both wells T-220 and T-6246
40 / Kenter et al.

FIGURE 21. Photomicrographs illus-


trating the dominant pore types in the
central platform area. (A) Pressure
solution stylolite with dark-brown
micrite and possibly bitumen. (B) As
in (A), but dissolution associated with
stylolite. (C) As in (A, B), but here,
dissolution postdating stylolite forma-
tion is clearly visible; note that the vug
is not breaking through the bitumen
seam. (D) Poikilotopic syntaxial over-
growth cements on crinoid ossicles are
locally dominant and reduce interpar-
ticle porosity. However, invasion by
bitumen and associated dissolution was
able to corrode cement and crinoid
ossicles. (E) Coated grainstone where
equant to rhombic spar is occluding in-
terparticle pore space, probably re-
placing earlier marine cements, and
only sparse molds and remaining
interparticle pores remain. (F) Micri-
tized coated grainstone with evidence
for minor corrosion of grain boundaries
followed by burial blocky calcite that
occludes part of the interparticle po-
rosity. Micrite is locally invaded by bi-
tumen. (G) Detail of partially open
fracture that was occluded by blocky
calcite cement and followed by bitumen.
(H) Porosity in poorly sorted grain-
stone was occluded by cement and
completely recrystallized to fine cal-
cite spar cement. Width of photomi-
crographs is 4.20 mm (0.16 in.)
(A – C, E – H) and 1.58 mm (0.06 in.)
(D). See text for discussion.

Figure 25 better shows the influ-


ence of bitumen volume on the po-
rosity versus permeability relation-
ship in well T-220 (no such data are
yet available for T-6246). Increasing
bitumen content above a threshold
of approximately 0.5% rock volume
appears to be comparable, with the exception of a rel- is negatively correlated with K-Phi. Although bitumen
atively higher contribution by equant and syntaxial commonly occurs in grainy facies with matrix pore types,
calcite at T-6246 (Figure 17). Another obvious difference it also occurs with late diagenetic pore types like vugs
between the T-220 and T-6246 wells is the amount of as well as with nonmatrix pore types like fractures.
bitumen; Figure 5 shows qualitatively higher amounts Bitumen content significantly impacts the cyclic and
of bitumen in T-6246 (and T-6846) than in T-220. For stratigraphic porosity variations as shown by Figure 26;
well T-220, bitumen volume is mostly associated with it also causes the observed decrease in porosity toward
muddy grain-rich to mud-rich facies (Figure 24A) and the outer platform. Porosity logs corrected for bitumen
dominant pore types like interparticle, intercrystalline, volume are easier to correlate and confirm that the dis-
and vug with minor contributions by intraparticle, tribution of bitumen increases toward the outer plat-
moldic, and fracture (Figure 24B). Bitumen content form (Figure 26). Clearly, bitumen reduced the porosity
shows a negative relationship with porosity (Figure 24C) toward the outer part of the central platform and even
and permeability (Figure 24D). more so into the outer platform itself (Figures 4, 5).
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 41

FIGURE 22. Pore-type contributions for muddy lagoonal facies (13 = PPW) and moderately sorted grainy facies with mud
matrix (2 = BSP, 31 = CSP, 32 = SPP) and mud-lean, well-sorted grainy facies (41 = SPG, 42 = CgGR, 43 = OG, 44 =
SIG, 46 = AG) in both T-220 and T-6246 wells. Pore types are 1 = interparticle; 2 = intraparticle; 3 = fenestral; 4 = shelter;
5 = growth framework; 6 = intercrystalline; 7 = moldic; 8 = microporosity; 9 = fracture; 10 = channel; 11 = vug; and
12 = enhanced dissolution. Plotted are relative dominance of pore types in terms of visual presence with the number
indicating whether pore types were first, second, or third in importance. See text for discussion.

Volcanic ash also produces low-porosity (tight) zones. determined by the combined effect of late diagenetic
These tight zones tend to occur around cycle bound- modification, specifically bitumen invasion and as-
aries and within intervals that have dispersed volcanic sociated corrosion, of an earlier reservoir system whose
ash. One mechanism explaining these observations porosity and permeability distribution was established
associates volcanic ash with inhibited downward fluid by highly cyclic depositional and early diagenetic
percolation during early diagenesis associated with processes (Figure 27). Early diagenesis includes me-
sequence and cycle boundaries. This condition gener- teoric alteration in cyclic platform facies associated
ally reduced the effect of meteoric dissolution. In ad- with major sequence boundaries and reduced disso-
dition, silica leached from the volcanic ash was, in lution along sequence and high-order cycle bound-
some cases, redeposited as chert and silica cement be- aries associated with the presence of volcanic ash.
low the sequence boundaries, thereby adding to the Diagenetic processes associated with subaerial expo-
porosity reduction of the associated lagoonal inter- sure produced both an increase and reduction of po-
vals. Finally, the dissolution associated with the bitu- rosity (Figure 27). Brown fibrous, mycrostalactitic,
men emplacement did not affect those tight zones. meniscus, and pendant cements are commonly as-
sociated with meteoric vadose and marine vadose
Processes and Timing of Diagenesis diagenesis, and these are observed in Tengiz cores to
The present-day distribution of reservoir quality in several meters below a suspected exposure surface
the central and (southwestern) outer platform in the (Figure 28A – D). Possible examples of alveolar fabrics
Lvis_SSB to Bash_SSB platform succession (Visean A, indicating soil formation below an exposure surface
Serpukhovian, and Bashkirian reservoir zones) was are shown in Figure 28E. Early dissolution related to
42 / Kenter et al.

FIGURE 23. Porosity distributions for the major facies types in T-220 and T-6246: muddy lagoonal facies (13 = PPW), moderately sorted grainy facies with mud
matrix (2 = BSP, 31 = CSP, 32 = SPP), and mud-lean well-sorted grainy facies (41 = SPG, 42 = CgGR, 43 = OG, 44 = SIG, 46 = AG). See text for discussion.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 43

FIGURE 24. Relationship between bitumen (Bit rock), rock fabric (PP fabric), pore type (pore 1), porosity (Phi), and
permeability (K ) in well T-220 (no such data are yet available for T-6246). Bitumen content is mostly associated with
grainy mud-lean to mud-rich facies (A) and dominantly in interparticle, intercrystalline, and vug pore types with minor
contributions by intraparticle, moldic, and fracture pores (B). Bitumen contents above 0.5% rock volume shows a
negative relationship with porosity (C) and permeability (D). Pore types are 1 = interparticle; 2 = intraparticle;
3 = fenestral; 4 = shelter; 5 = growth framework; 6 = intercrystalline; 7 = moldic; 8 = microporosity; 9 = fracture; 10 =
channel; 11 = vug; and 12 = enhanced dissolution. Textures are 1 = rudstone; 2 = grainstone-rudstone; 3 = grainstone;
4 = grainstone-packstone; 5 = packstone-grainstone; 6 = packstone; 7 = packstone-wackestone; 7 = wackestone-packstone; 8 =
wackestone; 9 = wackestone-mudstone; 10 = mudstone-wackestone; 11 = mudstone; and 12 = boundstone. See text for
discussion.

exposure may have led to early compaction as shown this process, which includes cementation of pore space
by the frequent occurrence of fitted fabrics (Figure 28H), created by fracturing (Figure 21G).
where the thin cement rim is included in the com- Burial dissolution events that added porosity are
paction and postdated by blocky calcite cement and of great interest and paramount importance to the
bitumen. Tengiz reservoir and its interpretation with respect
Later stage diagenesis includes pressure solution to porosity modeling (Figure 29A–F). A very common
culminating in stylolites with concentrations of dark- observation in samples that have interparticle poros-
brown micrite and bitumen (Figure 21A, B). Dissolution ity is enhanced (late-burial?) dissolution that enlarges
is sometimes observed postdating the stylolite forma- and connects interparticle pore space into larger vugs.
tion (Figure 21C). Poikilotopic syntaxial overgrowth These vugs are lined by bitumen, attesting to the timing
cements on crinoid ossicles are locally a dominant of the dissolution relative to bitumen emplacement
cement, reducing interparticle porosity (Figure 21D). (Figure 29A). Grain boundaries are commonly corrod-
Later burial processes also reduced porosity by cemen- ed by this dissolution event, as is syntaxial overgrowth
tation and recrystallization (Figure 21E – H). Exam- cement in some cases (Figure 29B, C). Extensive late
ples show progressive occlusion of pore space during deep burial dissolution prior to oil migration has been
44 / Kenter et al.

FIGURE 25. Influence of bitu-


men volume on the porosity
(PHI) versus permeability (K )
relationship in well T-220 (no
such data are yet available for
T-6246). Increasing bitumen
content is negatively correlat-
ed with K-Phi above a certain
threshold of about 0.5% rock
volume. Although bitumen is
common in grainy facies with
matrix pore types, it also oc-
curs with late secondary pore
types like vugs indicative of
dissolution as well as with non-
matrix pore types like fracture
porosity. See text for discussion.

occluded by blocky calcite


(see micrite envelope), where-
as this later dissolution phase
generated the open molds that
are lined by bitumen. More
common are finely recrystal-
lized fabrics associated with
intercrystalline and minor
moldic porosity, as well as
the presence of scattered bi-
tumen (Figure 29E, F). Al-
though the recrystallization
phase seems to postdate (cor-
reported as a major reservoir porosity generation mech- rode and truncate) blocky calcite cementation, the
anism in, at least, several other carbonate fields (Moore, exact timing remains unclear because of the destruc-
2001; Esteban and Taberner, 2003; Zampetti et al., tive nature of the process.
2003; Sattler et al., 2004). Most of these authors sug- The important diagenetic overprint in the Tengiz
gest corrosion by mixing of formation fluids with an central platform as outlined in Figure 27 can be sum-
external fluid at higher temperatures, and such pro- marized as follows. Syn- and postdepositional disso-
cess may be responsible for the burial corrosion in the lution in meteoric environments generated moldic,
Tengiz central platform as well. Similarly, the cor- enlarged interparticle and intraparticle, and micro-
rosion event postdating the bitumen invasion (and porosity pore types, as well as vuggy microkarst. Ce-
cementation) may represent yet another such leach- mentation of lagoonal intervals associated with the
ing phase with different fluids. Fluid-inclusion and presence of discrete and dispersed volcanic ash signifi-
stable isotope studies will shed light on the origin of cantly reduced porosity and created permeability baffles
the substantial burial corrosion at Tengiz; see Collins around cycle and sequence boundaries. Intermediate
et al. (2006) for discussion of preliminary geochem- shallow burial produced commonly observed fitted
ical data and a probable dissolution model. fabrics and low- to medium-amplitude stylolites. The
Recrystallization is also an important process in the lateral continuity and persistence of tight layers around
central platform diagenesis. As an example, Figure 29D sequence and cycle boundaries as well as in local high-
is a grainstone that was recrystallized to equant and energy intervals probably greatly affected later fluid
blocky spar. This recrystallization followed early par- flow and, thus, the ultimate distribution of cements,
tial cementation of interparticle porosity and was, dissolution, and bitumen in the central platform res-
in turn, followed by late dissolution that produced ervoir. The late diagenetic overprint was responsible
moldic porosity. Note that early moldic porosity was for the enhanced dissolution, increasing any existing
FIGURE 26. Subtle but significant differences in petrophysical and geological properties are observed between these four wells in the central platform (see Figure 4
for well locations and Figure 5 for additional correlations). K-Phi plots and porosity histograms suggest different behavior between northern central platform wells (T-220
and T-5246) and southern central platform wells (T-6246 and T-6846), with generally lower mean porosities and better behaved permeability-porosity relationships
in the southern wells and bimodal porosity distributions in the northern wells. Bash = Bashkirian; Serp = Serpukhovian; VisA = Visean A. See text for further discussion.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 45
46 / Kenter et al.

FIGURE 27. Paragenetic sequence for the central platform. Significant syndepositional, intermediate (shallow)-burial,
and deep (late)-burial events are listed in chronological order. In the lower part of the figure, those events that are
particularly important to porosity creation and destruction are listed. MO = moldic; BP = interparticle; IP = intraparticle;
MP = microporosity; MVUG = micro vug. See text for discussion.

porosity in the matrix and forming vugs and minor Figure 30 presents a model that summarizes po-
fractures, whereas equant calcite cement reduced po- rosity generation and reduction events in the cen-
rosity away from the interior central platform. The sub- tral Tengiz platform from the standpoint of gamma-
sequent bitumen invasion exerted an overall porosity- ray and porosity logs. Porosity changes are linked to
reducing effect increasing toward the outer platform. the major diagenetic phases following deposition of
These late-burial events that increased and reduced alternating thin, low-energy lagoonal intervals with
porosity significantly dampened or flattened initial volcanic ash beds and thick intervals of high-energy
vertical, nearly cyclic, porosity variations linked to grainstone. Detailed log and core correlation shows
primary depositional properties. that lagoonal interval and ash beds are not always
associated and/or continuous between well locations,
and cycles may vary slightly in thickness over kilo-
Spatial Distribution of Central meters distance. Three important diagenetic stages
Platform Reservoir Quality that affect the reservoir quality and distribution are
In the absence of any significant first-order con- observed. First, porosity curves with a blocky shape
trol by pore type or lithofacies, the following sec- are the result of early-burial processes, including
tions describe the reservoir quality in the context of low porosity around cycle boundaries related to the
the transition from central to outer platform envi- presence of thin ash beds and corrosion and cemen-
ronments. This lateral change is likely a result of the tation related to meteoric diagenesis (Figure 30A).
early cyclic depositional and diagenetic processes, in- Locally thick grainstone intervals have low porosity
cluding the preservation of volcanic ash, coupled with and show the presence of elevated levels of dispersed
late-burial dissolution and bitumen emplacement. ash concentrated in frequent stylolites. Second, a
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 47

FIGURE 28. Photomicrographs illustrat-


ing some of the dominant cement
types in the central platform area. (A,
top to the right) Brown fibrous pendant
cement attached to lower part of bra-
chiopod fragments is interpreted as
the result of meteoric diagenesis and,
therefore, suggests subaerial exposure
above. (B) Mycrostalactitic or pendant
cements, which are commonly associ-
ated with vadose diagenesis, are ob-
served to several meters below a suspect
exposure surface. (C, D) Meniscus and
pendant cements, indicating vadose
to marine vadose diagenesis occlude
much of the primary porosity in these
samples. Burial dissolution and/or cor-
rosion of preserved interparticle porosity
are followed by blocky calcite cement
and bitumen. (E) Possible example of
alveolar fabrics indicating soil formation
below exposure surface. (F) Possible
vadose cementation followed by en-
hanced corrosion or dissolution of grain
boundaries and cement, blocky calcite
cement, and bitumen. (G, H) A combi-
nation of minor cementation and com-
paction leads to fitted fabrics or over-
packed fabrics. Here, a thin cement
rim is included in the compaction pro-
cess and followed by burial blocky
calcite spar cement and bitumen. Width
of photomicrographs is 4.20 mm
(0.16 in.) (A, C, D, E, F) and 1.58 mm
(0.06 in.) (B, G, H). See text for
discussion.

rim and flank area, observations indi-


cate a corrosion event following the
bitumen cement (Collins et. al.,
2006).
Packstone to mudstone intervals
around the cycle boundaries in the
central platform wells typically dis-
burial event led to corrosion and enhanced blocky play significantly lower porosity and permeability rela-
porosity curves in the interior central platform and tive to the overlying grainstone-packstone lithofacies
porosity reduction in the exterior central platform; (Figure 31, sequences between Lvis10.5_csb-Lvis9.5_csb
alternatively, corrosion in the exterior central plat- in wells T-5246 and T-220). The latter have porosity
form was lower compared to the interior central plat- ranges between 0 and 24% with a mean of 10.5 and
form (Figure 30B). Finally, bitumen that was emplaced 11.0% (Table 2). The packstone to mudstone inter-
during a late-burial event acts as cement and reduces vals are generally associated with gamma-ray spikes
porosity (Figure 30C). Bitumen volume increases from caused by the presence of volcanic ash and are gen-
the interior to the exterior central platform and margin erally tight with a mean porosity of 3.5%. Within the
and obscures correlation using the porosity curves. central platform, the cyclic signature of alternating
Some evidence suggests that another corrosion phase blocky high-porosity and low-porosity zones associ-
is associated with the bitumen emplacement; in the ated with gamma-ray spikes is dominant. Lateral and
48 / Kenter et al.

FIGURE 29. Photomicrographs illustrating additional pore types of the central platform area. (A) Enhanced late-burial(?)
dissolution joining interparticle pores into larger vugs prior to bitumen lining the pore walls. The primary grain
boundaries appear to have been corroded by the dissolution. In addition, following bitumen emplacement, grains were
partially dissolved by an additional dissolution event. (B) Bitumen corroding both the crinoid ossicle boundaries as well as
those of the syntaxial overgrowth cement. (C) As in (B), but advanced stage of bitumen emplacement corroding grain
boundaries and leading to pseudopackstone to wackestone fabric. (D) Recrystallization is a dominant process in the
central platform diagenesis. Here, an example where the grainstone fabric was recrystallized to equant and blocky spar
following early partial cementation of interparticle porosity and followed by late dissolution generating the moldic
porosity. Note that early moldic porosity was occluded by blocky calcite (see micrite envelop), whereas a later disso-
lution phase generated the open molds that are lined by bitumen. (E, F) Finely recrystallized fabrics associated with
intercrystalline and minor moldic porosity, as well as the presence of scattered bitumen, are very common in the
central platform. Although the recrystallization phase seems to postdate (corrode and truncate) blocky calcite cemen-
tation, the exact timing remains unclear because of the destructive nature of the process. Width of photomicrographs
is 4.20 mm (0.16 in.) (C, F) and 1.58 mm (0.06 in.) (A, B, D, E). See text for discussion.

vertical changes in distribution and type of the in- A remarkable deviation in the central platform
termediate grainstone to packstone lithofacies have occurs in the lower part of the Bash1_csb, where
only a minor effect on the present-day porosity. The mostly well-sorted ooid grainstone is invaded and
interval on Figure 31 in well T-5246 represents high- cemented by bitumen all across the platform. The re-
energy grainstone, and although the interval in well sulting porosity range is 0–10%, with a mean of 4%,
T-220 is lower energy packstone-grainstone facies, whereas in bitumen-poor ooid grainstone, the poros-
no visible difference is observed on the porosity ity ranges between 0 and 21%, with a mean of 12%
log. (Table 2). The fact that bitumen (as much as 3.5%
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 49

FIGURE 30. Hypothetical log correlation scenarios in the central platform, with gamma ray shown on the left and porosity
on the right for each well(s). Porosity generation and reduction events that are linked to the major diagenetic phases
are indicated for a stratigraphy that includes thin, low-energy lagoonal intervals alternating with volcanic ash beds (yellow
and green horizons) and thick intervals of high-energy grainstone (white). Detailed correlation shows that lagoonal
intervals and ash beds are not always associated and/or continuous between well locations, and cycles (triangles) may
vary slightly in thickness over kilometers distance. (A) Porosity curves with a blocky shape are the result of early-burial
processes, including porosity loss around cycle boundaries related to the presence of thin ash beds and corrosion and
cementation related to meteoric diagenesis (porosity is blue). Locally thick grainstone intervals have low porosity and
show the presence of elevated levels of dispersed ash concentrated in frequent stylolites. (B) Late-burial events leading
to corrosion and enhanced blocky porosity curves in interior central platform (red) and porosity reduction in the
exterior central platform (yellow); alternatively, corrosion in the exterior central platform was lower compared to the
interior central platform. (C) Bitumen, which is emplaced during late burial, acts as cement and reduces porosity.
Bitumen volume increases from the interior to the exterior central platform and margin and obscures correlation using
the porosity curves.

rock volume) invaded this far into the central plat- Bitumen occurs within fine matrix and as a pore-
form may be related to the continuous underlying filling phase, and two spatial trends are observed (see
pressure baffle, at the Serp_SSB, which preexisted Figure 5, 25): (1) a general increase from exterior to
prior to the bitumen and associated dissolution phase interior central platform wells and (2) elevated bitu-
(Figure 32A). men levels near the base of grainy facies between
50 / Kenter et al.

FIGURE 31. Cross section showing lithofacies, gamma-ray (left), and porosity (right) logs; well locations are shown in Figure 4, and lithofacies legend is shown in

and T-6846 wells. The cyclic and blocky high-low porosity character is visible in nearly bitumen-free wells (T-5246 and T-220) and represents, respectively, alternating
tight lagoonal intervals. In addition, some ooid grain-

grainy facies and lagoonal mud intervals with volcanic ash. Although the interval in T-5246 represents high-energy and essentially mud-free grainstone, and that
Figure 5. The lateral facies and thickness continuity of the sequences between Lvis10.5_csb and Lvis9.5_csb are shown between the T-5246, T-220, T-6246,
stone (right above the Serp_SSB) and crinoid grain-
stone-packstone intervals (mostly between the Lvis_SSB
and Serp_SSB) show increased bitumen content all
across the central platform. The late timing of bitu-
men relative to other diagenetic overprint suggests
the introduction through the flanks and into the
platform part of the Tengiz buildup, with the largest
lateral extent in the most permeable units and resid-

in T-220 is lower energy and packstone-grainstone facies, no visible difference is observed on the porosity log. See text for discussion.
ual deposition controlled by gravity, that is, on top of
the immediate underlying impermeable cycle bound-
ary (see also Figure 26). Additional intervals with
elevated bitumen extend to T-220 but disappear up-
section and in the direction of T-5246 (Figures 5, 32B).
In the north (well T-5044) and southwest (wells
T-6246 and T-6846) parts of the central platform,
overall bitumen content in the Lvis_SSB to Bash_SSB
grainy facies reduces porosities to a mean of 5.8–
7.6%, which is significantly reduced from the 10.6 –
11.2% recorded in well T-220 (Figure 5). In general,
increased bitumen content is commonly confined to
the lower part of a cycle. The bitumen-cemented
interval exhibits a sharp, lower contact with under-
lying, well-cemented peloid packstone-wackestone
lithofacies, has a gradational upper contact, and ap-
pears dark gray to black in core. Although the blocky
log character changes to a more irregular and spiky
shape with increasing bitumen content, no obvious
relationship exists between the amount of porosity
reduction and primary lithofacies type. Instead, the
presence of the tight peloid packstone-wackestone
baffles seems to control the occurrences of bitumen-
cemented interval (see Figures 31, and 33, well T-220
versus well T-6246, no bitumen but lateral facies
change). The bitumen effect progressively increases
away from the platform center and, at the scale of
the plug sample, reduces the porosity and perme-
ability as suggested by the relationship shown in
Figure 24.
In addition to porosity and permeability reduction
by bitumen and volcanic ash layers, the presence of
dispersed volcanic ash in high-energy grainstone fa-
cies locally also has an effect (Figure 34). Volcanic ash
dispersed in the well-sorted skeletal grainstone and
coated-grain grainstone-rudstone lithofacies of the
Serp1_csb sequence at well T-5447 reduces porosity
to nearly zero and generates a nearly flat porosity
curve, whereas the same facies interval at well T-220
retains high porosity. Similarly, the tight and later-
ally continuous interval associated with the Serp_SSB
is probably also associated with elevated levels of dis-
persed volcanic ash (see Figure 5).
FIGURE 32. Cross sections showing lithofacies, gamma-ray (left), and porosity (right) logs; well locations are shown in Figure 4, and lithofacies legend is shown
in Figure 5. The lateral facies and thickness continuity of sequences between the Serp_SSB and Lvis13_csb are shown for the T-220 and T-6246 wells in (A) and
between the Bash2.5_MFS to Bash4_csb for the T-5246, T-5447, T-220, and T-6246 wells in (B). The lower part of the Bash1_csb, mostly well-sorted ooid grainstone
(OG), is invaded and cemented by bitumen all across the platform. Additional intervals with elevated bitumen extend to T-220 but disappear upsection and toward
T-5246. See text for discussion.
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 51
52 / Kenter et al.

section increases (as much as 2 – 4% rock volume bitumen in some parts of T-6246), and generally, elevated levels of bitumen are commonly confined to the lower
FIGURE 33. Cross section showing lithofacies, gamma-ray (left), and porosity (right) logs; well locations are shown in Figure 4, and lithofacies legend is shown in

southwestern (interpreted as outer platform) margin and away from the central platform to the north (T-5044), overall bitumen in the Lvis_SSB to Bash_SSB
IMPLICATIONS FOR RESERVOIR MODELING

Figure 5. The lateral facies and thickness continuity of sequence Lvis9.5_csb to Lvis9_csb are shown for wells T-5246, T-5447, T-220, and T-6246. Toward the
The stratigraphic framework introduced for the
central platform part of the Tengiz buildup in the
preceding sections is the basis for the layering scheme
that is being used in reservoir models for the field. Full-
field models use the layering portrayed on Figure 4,
whereas more detailed models focusing on parts of
the platform incorporate more detailed layers as
shown in Figures 5 and 25. The lithofacies summa-
rized in Table 2 have been grouped and used to
subdivide layers in the reservoir models into regions
where porosity and permeability might be expected to
vary. The complex diagenesis that occurs in the central
platform and the outer platform has overprinted the
original (depositional) porosity to an extent, as sum-
marized in Figure 30, that it cannot be ignored when
trying to interpret reservoir quality from wire-line logs
and correlate it within a reservoir model. Ongoing studies
of the detailed analysis of the types and amounts of
porosity, as well as pore-filling agents are being used to
better understand the log signatures of porosity and
improve porosity-permeability transforms. Much work
remains in the detailed characterization of the Tengiz
reservoir and in correctly portraying the significant
geologic characteristics of the Tengiz platform in
reservoir models.

SUMMARY AND CONCLUSIONS

An integrated (and ongoing) study of core, well


logs, and discrete petrophysical measurements provid-
ed a first comprehensive assessment of depositional
evolution and reservoir quality in the late Visean to
Bashkirian interval in the Tengiz central platform.
The study resulted in the following observations and
tentative conclusions.
Depositional cycles (=higher frequency sequences)
are up several to tens of meters in thickness and made
part of the cycle. See text for discussion.

up of a succession of lithofacies overlying a sharp


base with variable evidence for subaerial exposure. At
the base of the cycles, tight peloidal mudstone and
volcanic ash beds are associated with sequence bound-
aries and represent flooding events. They are overlain
by brachiopod- and crinoid-dominated intervals that
represent maximum marine flooding and near the top
by skeletal-peloidal grainstone that are interpreted as
the highstand-shoaling phase. Nonskeletal grainstones
are most common in the Bashkirian. Visean and Ser-
pukhovian cycles are generally easy to correlate lat-
erally, whereas Bashkirian cycles are best explained by
the transition to high-frequency icehouse sea level
Late Visean to Bashkirian Platform Cyclicity in the Central Tengiz Buildup / 53

FIGURE 34. Cross section showing lithofacies, gamma-ray (left), and porosity (right) logs; well locations are shown in Figure 4, and lithofacies legend is shown

bitumen-reducing porosity, dispersed volcanic ash in high-energy grainstone facies locally has an even greater effect on porosity in the central platform. Volcanic
fluctuations, which generate incomplete cycles and

ash dispersed in the SPG and CgPR lithofacies in the Serp1_csb sequence at T-5447 reduces porosity to nearly 0%, generating a flat porosity curve, whereas
in Figure 5. The lateral facies and thickness continuity of sequences between Serp_SSB and Lvis13_csb are shown for wells T-5447 and T-220. In addition to
complex lateral facies changes.
The distribution of reservoir quality in the central
platform was initially determined by early diagen-
esis associated with cyclic deposition, which resulted
in a stacked system of high-porosity intervals bor-
dered by tight higher order sequence boundaries with
reduced dissolution associated with the presence of
volcanic ash. This diagenetic phase includes low-
porosity, thick (discontinuous and not platformwide)
grainstone intervals with high amounts of dispersed
volcanic ash. The lateral continuity of tight cycle
boundaries probably greatly affected later fluid flow
as well as the ultimate distribution of cements, dis-
solution, and bitumen in the central platform res-
ervoir. Burial diagenetic overprint, an initial corrosion
phase followed by a pore-filling phase, and finally,
bitumen invasion and associated corrosion, produced
an overall porosity reduction toward the exterior cen-
tral platform. Bitumen content decreases from the
exterior to the central platform and is highest near the
bases of the cycles. This suggests that the first fill of
hydrocarbons migrated through the flanks laterally
into the platform cycles. Both corrosion and bitumen
emplacement flattened the initial cyclic porosity va-
riations and further complicated porosity and perme-
the same facies interval at T-220 retained high porosity. See text for discussion.

ability relationships.

ACKNOWLEDGMENTS

The authors thank the following individuals, who


provided insightful comments or suggestions and,
in some cases, data to greatly improve this manuscript.
Perhaps most importantly, Michael Clark (Chevron)
is thanked for initiating an intense coring and core
analysis program that formed the basis for most of
our geologic studies and for his strong support of the
sequence-stratigraphic framework. Ben Robertson
(Chevron) provided important discussion during the
continued coring program and supporting geological
analyses of outcrop analogs and literature reviews.
Numerous coworkers in Chevron, ExxonMobil, and
TengizChevroil assisted with our core and log studies
and participated in lively discussions on many topics.
Giovanna Della Porta (Cardiff University) and Frans
van Hoeflaken (independent consultant) are acknowl-
edged for critically reviewing the petrographic and
lithofacies descriptions. Steve Jenkins (TengizChev-
roil) offered numerous comments and observations,
which tested the ideas expressed in this study in
their practical application for three-dimensional
54 / Kenter et al.

geologic modeling of the Tengiz platform. Comments development during late burial in carbonate reservoirs
by AAPG reviewers helped to greatly improve the as a result of mixing and/or cooling of brines: Journal of
Geochemical Exploration, v. 78 – 79, p. 355 – 359.
manuscript. Finally, we thank TengizChevroil and its
Gradstein, F. M., J. G. Ogg, and A. G. Smith, 2004, A
shareholder companies (Chevron, ExxonMobil, Kaz-
geologic time scale: International Commission on
munaigaz, and BPLukArco) for permission to publish Stratigraphy (ICS) under: www.stratigraphy.org.
this study. Lucia, F. J., 1995, Rock-fabric/petrophysical classification
of carbonate pore space for reservoir characterization:
REFERENCES CITED AAPG Bulletin, v. 79, p. 1275 – 1300.
Lucia, F. J., 1999, Carbonate reservoir characterization:
Brenckle, P. L., and N. V. Milkina, 2001, Foraminiferal New York, Springer-Verlag, 222 p.
timing of carbonate deposition on the Mississippian – Moore, C. H., 2001, Carbonate reservoir-porosity evolu-
early Pennsylvanian Tengiz platform, Kazakhstan: tion and diagenesis in a sequence stratigraphic frame-
Paleoforams 2001 — International Conference on work: Developments in Sedimentology, v. 55, p. 444.
Paleozoic Benthic Foraminifera, Abstracts, Ankara, Ross, C. A., and J. R. P. Ross, 1985, Carboniferous and Early
August 20 – 24, p. 2001. Permian biogeography: Geology, v. 13, p. 27 – 30.
Choquette, P. W., and L. C. Pray, 1970, Geologic no- Sattler, U., V. Zampetti, W. Schlager, and A. Immenhau-
menclature and classification of porosity in sedimen- ser, 2004, Late leaching under deep burial conditions:
tary carbonates: AAPG Bulletin, v. 54, p. 207 – 250. A case study from Miocene Zhujiang carbonate reser-
Collins, J. F., J. A. M. Kenter, P. M. Harris, G. Kuanysheva, voir: South China Sea: Marine and Petroleum Geology,
D. J. Fischer, and K. L. Steffen, 2006, Facies and v. 21, p. 977– 992.
reservoir-quality variations in the late Visean to Bash- Weber, L. J., B. P. Francis, P. M. Harris, and M. Clark, 2003,
kirian outer platform, rim, and flank of the Tengiz Stratigraphy, lithofacies, and reservoir distribution,
buildup, Precaspian Basin, Kazakhstan, in P. M. Harris Tengiz field, Kazakhstan in W. M. Ahr, P. M. Harris, W. A.
and L. J. Weber, eds., Giant hydrocarbon reservoirs of Morgan, and I. D. Somerville, eds., Permo-Carboniferous
the world: From rocks to reservoir chacracteriza- carbonate platforms and reefs: SEPM Special Publica-
tion and modeling: AAPG Memoir 88/SEPM Special tion 78 and AAPG Memoir 83, p. 351 – 394.
Publication, p. 55 – 95. Zampetti, V., W. Schlager, J. H. Van Konijnenburg, and
Cook, H. E., W. G. Zempolich, V. G. Zhemchuzhnikov, A. J. Everts, 2003, Architecture and growth history of a
and J. J. Corboy, 1997, Inside Kazakstan: Cooperative Miocene carbonate platform from 3D reflection data:
oil and gas research: Geotimes, v. 42, no. 11, p. 16 – 20. Luconia Province, offshore Sarawak, Malaysia: Marine
Esteban, M., and C. Taberner, 2003, Secondary porosity and Petroleum Geology, v. 21, p. 517 – 534.
2
Collins, J. F., J. A. M. Kenter, P. M. Harris, G. Kuanysheva, D. J. Fischer, and
K. L. Steffen, 2006, Facies and reservoir-quality variations in the late Visean to
Bashkirian outer platform, rim, and flank of the Tengiz buildup, Precaspian
Basin, Kazakhstan, in P. M. Harris and L. J. Weber, eds., Giant hydrocarbon
reservoirs of the world: From rocks to reservoir characterization and mod-
eling: AAPG Memoir 88/SEPM Special Publication, p. 55 – 95.

Facies and Reservoir-quality


Variations in the Late Visean to
Bashkirian Outer Platform, Rim,
and Flank of the Tengiz Buildup,
Precaspian Basin, Kazakhstan
J. F. Collins G. Kuanysheva
ExxonMobil Development Company, TengizChevroil, Atyrau, Kazakhstan
Houston, Texas, U.S.A.
D. J. Fischer
J. A. M. Kenter1 TengizChevroil, Atyrau, Kazakhstan
Vrije University, Amsterdam, Netherlands
K. L. Steffen
P. M. Harris ExxonMobil Development Company,
Chevron Energy Technology Company, Houston, Texas, U.S.A.
San Ramon, California, U.S.A.

ABSTRACT

T
engiz field is an isolated carbonate buildup in the southeastern Precaspian
Basin, containing a succession of shallow-water platforms ranging in age
from late Famennian to early Bashkirian. Platform backstepping from Tour-
naisian through late Visean resulted in approximately 800 m (2625 ft) of bathy-
metric relief above the Famennian platform. This was followed by as much as
2 km (1.2 mi) of Serpukhovian progradation, which formed a depositional wedge
around the older platforms referred to as the Serpukhovian rim and flank.
Rim and flank facies include lower slope mudstone, volcanic ash, and platform-
derived skeletal packstone to grainstone interbedded with boundstone breccia;
middle-slope poorly bedded to massive boundstone breccia with subtypes based
on clast composition, size, and packing; upper-slope in-situ microbial bound-
stone; and outer-platform to shallow-platform skeletal, coated-grain, and ooid
packstone to grainstone. The upper-slope microbial boundstone represents the

1
Present address: Chevron Energy Technology Company, San Ramon, California, U.S.A.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215874M881469

55
56 / Collins et al.

dominant source of clasts in the middle- and lower-slope breccias. Periodic large-
scale failure of the rim during both Serpukhovian and Bashkirian time resulted in
a high degree of lateral facies discontinuity.
Solution-enlarged fractures, large vugs, and lost circulation zones produced
mainly during late diagenesis form a high-permeability, well-connected reser-
voir in the rim and flank. This diagenetic overprint is associated with the pres-
ence of bitumen and extends upward into overlying Serpukhovian and Bash-
kirian platform facies and inward into adjacent late Visean platforms, where it
has substantially altered reservoir properties that remained after early diagenesis
related to cyclic depositional processes.

INTRODUCTION STRATIGRAPHY

The Tengiz field is located in onshore western Ka- In 2001, a joint study by ExxonMobil, Chevron,
zakhstan at the northern end of the Caspian Sea’s and TCO established a stratigraphic framework for
eastern shoreline, among an archipelago of isolated the Tengiz platform based on seismic, biostratrigraphic,
carbonate buildups that grew on a regional pre-Middle core, and well-log data that existed up to that time.
Devonian structural high (Figure 1). Tengiz was dis- This study was published (Weber et al., 2003) and
covered in 1979 by the Oil Ministry of the former stands as the definitive stratigraphic reference for Ten-
Soviet Union and, since 1993, has been operated by giz. The published framework includes a hierarchy
TengizChevroil (TCO), an in-country joint-venture of second-order, third-order, and fourth-order sequences
company. The field is covered by a three-dimensional that have been maintained as Tengiz drilling pro-
(3-D) seismic survey encompassing about 1000 km2 ceeds. This architecture is a locally applicable, fit-for-
(385 mi2), acquired in 1998. As of year-end 2005, more purpose framework, although it more or less con-
than 115 wells have penetrated the reservoir at Tengiz forms to the second-order eustatic cycles of Ross and
(Figure 2), which produces an intermediate sulfur Ross (1987).
oil (478 API). The highest initial production rates are This chapter describes facies and reservoir-quality
from wells located in the buildup rim and flank areas, variations in the Visean A through Bashkirian plat-
in fractured carbonates with low (<6%) matrix poros- form margins at Tengiz (Figure 3B). This interval forms
ity. Wells located toward the center encounter higher the late highstand of a second-order supersequence
porosity (as much as 18%) but lower matrix perme- set that began with deposition of a series of back-
ability (<10 md) and lower initial rates. stepping platforms in the Tournaisian. By the end
The Tengiz buildup formed sometime after the Mid- of the Visean, the platform area had decreased from
dle Devonian and grew more or less uninterrupted 210 km2 (81 mi2) in the late Famennian to about
until the early Bashkirian by aggradation of a series of 90 km2 (35 mi2). During the Serpukhovian, the Tengiz
grain-dominated platforms. The total thickness of car- platform prograded as much as 2 km (1.2 mi) basin-
bonate is unknown, but is represented by approxi- ward, filling much of the accommodation space cre-
mately 1.2 s TWT on 3-D seismic data. The deepest ated during backstepping, and forming a microbial
borehole to date bottomed in the middle Famennian at boundstone-cored depositional wedge known as the
6032 m (19,790 ft) subsea, more than 2100 m (6890 ft) rim and flank around the late Visean and older plat-
below the top of the reservoir and more than 700 m forms. This wedge attained a maximum thickness of
(2297 ft) below the top of the Famennian platform. more than 800 m (2625 ft). Following a brief deposi-
Following platform demise in the Bashkirian, the tional hiatus on the platform (Brenckle and Milkina,
Tengiz buildup was buried initially by a thin veneer 2003), early Bashkirian platform carbonates accumu-
of Moscovian through Artinskian carbonates and vol- lated more or less aggradationally above the Serpu-
caniclastics and then by thick Kungurian evaporites khovian. Because this chapter also discusses reservoir
(mainly halite) that completely encased the platform. connectivity between the Serpukhovian rim and flank,
Post-Permian salt diapirism produced a series of salt the overlying Bashkirian outer platform region, and
pillars and clastic-filled withdrawal basins above the the adjacent Visean A platforms, a brief description of
platforms, which locally adversely affect 3-D seismic how platform facies correlate to rim and flank facies
image quality (Figure 3A). is included here. For a detailed account of the nature
FIGURE 1. North Caspian region and location of the Tengiz field.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 57
58 / Collins et al.

FIGURE 2. Structure contours


on top of the Tengiz buildup.
Current well locations and the
path of the seismic profile in
Figure 3 are indicated. Contour
interval = 100 m (330 ft).

than 10 m (33 ft) to several


tens of meters. They consist
of dipping, bedded to mas-
sive skeletal packstone to
rudstone containing abun-
dant large crinoids and bra-
chiopod fragments, minor
colonial coral detritus, and
occasional platform-derived
breccias. In the central plat-
form, fourth-order sequence
boundaries are associated
with subaerial exposure and
deposition of low-energy or
restricted facies containing
thin volcanic ash layers (Ken-
ter et al., 2006). In the Ser-
pukhovian outer-platform
cycles, these features are re-
duced or absent at the se-
quence boundaries.

of the Visean A, Serpukhovian, and Bashkirian cen- DEPOSITIONAL ENVIRONMENTS


tral platform regions, the reader is referred to Kenter
et al. (2006). General Features of the Tengiz Rim and Flank
The terms ‘‘platform,’’ ‘‘rim,’’ and ‘‘flank’’ were
originally derived from seismic features and are now
Visean A, Serpukhovian, and used for drilling purposes to denote geographic re-
Bashkirian Platform Cycles gions in the field. The platform region comprises
Visean A, Serpukhovian, and Bashkirian fourth- boreholes that penetrate the stack of backstepping
order central platform sequences generally adhere platforms, whereas boreholes that encounter any part
to the model shown in Figure 4. These sequences of the adjacent progradational wedge are described
feature depositional environments ranging from high- as rim and flank wells. As mentioned earlier, the
energy grainstone shoals to slightly deeper environ- wedge arises as a result of Serpukhovian prograda-
ments characterized by more poorly sorted, open- tion, but includes rocks of Serpukhovian and Bash-
marine grainstone to packstone, or in some cases, kirian age (see Figure 3). The rim area refers to the
mudstone and faunally restricted, fine-grained car- inner, structurally elevated rim crest or raised rim,
bonate sand. Outer platform cycles vary over the in- whereas the flank refers to the outer sloping surface.
terval but are generally characterized by an increase The raised rim is interpreted to have formed mainly
in algal grains and micritic intraclasts. Because of the by differential compaction caused by a mechanical
unique nature of microbial facies in the rim and contrast between grainy platform facies and more
flank, greater attention is given to Serpukhovian rigid upper-slope microbial boundstone facies in the
outer-platform cyclicity. These cycles are commonly rim and has resulted in as much as 50 ms of seismic
poorly expressed and range in thickness from less relief and as much as 200 –250 m (656 – 820 ft) of
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 59

FIGURE 3. Tengiz stratigraphic framework and the Serpukhovian rim and flank region. The seismic line (A) shows the
east-west asymmetry in the shape of the wedge and in the width of the structurally elevated portion (raised rim). The
seismic profile also shows the absence of Moscovian – Artinskian volcanic sediments on the upper flank of the buildup,
placing an anhydrite layer at the base of the salt in direct contact with the reservoir, as well as the source rock interval
associated with subsalt facies in the same interval in surrounding lows. These are two potentially important factors in rim
and flank diagenesis. The schematic framework (B) is derived from Weber et al. (2003), with details from this study
added in the rim and flank area. In particular, the diagram shows the rim and flank to consist of an early debris apron
that accumulated around the base of the buildup, followed by progradation of a boundstone-cored upper wedge (black
dashed line).
60 / Collins et al.

FIGURE 4. Idealized Tengiz fourth-order platform sequence, showing textural and compositional variations associated
with platform bathymetry for the Bashkirian, Serpukhovian, and late Visean. Outer-platform, rim, and flank facies
illustrated apply particularly to the Serpukhovian. For other intervals, outer-platform facies, in general, are characterized
by an increase in algal grains, micritic intraclasts, and coarse crinoid-brachiopod debris. Sea level lowstands are
associated with exposure of shallow-platform shoals, contemporaneous formation of an outer margin complex of
algal-intraclast grainstones and possible coral-algal patch reefs, and local development of restricted conditions on the
platform. Partly restricted-marine conditions that persisted during the early transgression were dominated by bra-
chiopods, including the formation of outer-platform brachiopod mud mounds. Widths of facies belts are not shown
to scale.

structural relief1 (see Figure 2). Evidence for platform into the rim, based on comparisons between Ten-
compaction includes: giz and other microbial boundstone-cored plat-
forms (Kenter et al., 2005).
1) a lack of volumetrically significant deep lagoonal
facies in the central areas of the Visean A, Ser- The break in slope separating the raised rim and
pukhovian, and Bashkirian platforms flank areas does not necessarily indicate the limit
2) upper Serpukhovian and Bashkirian platform of Serpukhovian progradation. In some regions, this
cycles in the raised rim that contain shallow fa- break appears to be controlled by faulting or mass
cies similar to age-equivalent cycles in the cen- wasting. In addition, differential compaction of a
tral platform basinward-thickening wedge of middle- to lower-
3) significant grain-to-grain compaction observed slope facies may be responsible for the slight basin-
in thin sections from the platform facies ward dip of the raised rim observed in some areas.
4) observation of high early cement volumes in the
rim and flank facies, indicating early develop- Rim and Flank Depositional Environments
ment of mechanical rigidity The rim and flank contain carbonate facies depos-
5) a lateral facies progression in Serpukhovian cy- ited in water depths ranging from a few meters to sev-
cles that deepens basinward from the platform eral hundred meters. Facies descriptions come mainly
from cores2 and from Formation MicroImager (FMI)
1
Weber et al. (2003) estimated that approximately two-thirds of the
2
raised rim relief was caused by differential compaction, and one-third was Approximately 1200 m [3937 ft] of core from 15 wells available as of
caused by a bioconstructed margin associated with upper Bashkirian cycles. year-end 2005 were examined for this study.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 61

logs calibrated to cores. Facies interpretation from FMI platform cycles, indicates that relief exceeded at
logs was hampered by use of oil-based mud during least some of the ambient low-amplitude eustatic
drilling, which tended to reduce conductivity con- variations, estimated to be between 25 and 50 m
trasts between porous and nonporous rock, and by (82 and 164 ft) (Ross and Ross, 1987, 1988).
frequent hole washouts and zones of lost circulation Whereas the succession of facies comprising cy-
that adversely affected image quality. Rim and flank cles in the central platform area is well documented
facies are associated with a depositional profile con- in cores (see Kenter et al., 2006), the descriptions of
sisting of the following environments, in the order Serpukhovian outer-platform cycles contained here
of increasing paleowater depth: are generalizations based on a limited number of
complete cycles present in the current suite of cores.
1) cyclic shallow-platform deposits The basal parts of some outer-platform cycles were
2) outer-platform skeletal-intraclast rudstone, grain- observed to contain breccias composed of colonial
stone, packstone, and minor boundstone coral detritus or, in other cases, rounded, platform-
3) upper-slope microbial boundstone derived grainstone clasts (Figure 6A, B). The breccias
4) middle slope welded to mosaic boundstone contain either microbial matrix textures and cements
breccia or a matrix of coarse skeletal debris. They may repre-
5) lower-slope detrital boundstone breccia and grainy sent lowstand deposits, indicating updip platform
platform-derived deposits exposure, or development of outer-margin coral patch
6) lower-slope and debris apron boundstone breccias reefs. In other sequences, brachiopod mud mounds a
and bedded, fine-grained periplatform deposits. few meters thick (Figure 6C) are encountered near
the base. The middle and upper parts of cycles are
These are partly based on environments and fa- characterized by wackestones, packstones, and rud-
cies recognized in outcrops of the Asturias plat- stones containing large crinoids, brachiopod shell
form margin in the Cantabrian region of northern fragments, and solitary rugose corals (Figure 6D), in-
Spain and from the Capitan margin in west Texas dicating an increase in water depth. This facies is
and New Mexico (Figure 5). The Tengiz outer plat- associated with dipping beds (Figure 6E), indicat-
form, slope, and debris apron facies are described ing deposition on a paleoslope, and may also ex-
in the ensuing paragraphs and are also summa- hibit microbial matrix textures. Cycle tops tend to be
rized in Appendix 1. The cyclic, shallow-platform characterized by a reduction in crinoid-brachiopod-
facies are described in Kenter et al. (2006). coral grains and an increase in coated-grain or algal-
intraclast grainstone, resembling some shallow-
Outer-platform Facies platform grainstones. Subaerial exposure effects like
The Asturias and Capitan margins were both formed those described from the central platform (Kenter
by deeper water microbially influenced boundstone et al., 2006) are commonly absent, but large disso-
margins (Kirkland et al., 1998; Della Porta et al., 2003) lution voids filled with multilayered geopetal sedi-
and support the interpretation of considerable origi- ment and multiple generations of carbonate cement
nal outer-platform relief above the Tengiz upper slope are found, as are brecciated grainstone intervals
during the Serpukhovian. The Asturias margin had with evidence of clast rotation and mixing. Iso-
outer-platform relief of between 20 and 50 m (66 and lated, sediment-filled dissolution voids are also scat-
164 ft) during times of progradation (Bahamonde tered throughout the section, but appear to increase
et al., 2000; Kenter et al., 2005). Estimates of the re- in frequency at or near the tops of cycles. These may
lief across the Capitan outer platform margin show be associated with longer term exposure at superse-
variations through time from 10 m to more than quence boundaries (Serp_SSB and Bash_SSB), or they
60 m (33 to more than 197 ft) (Kerans and Harris, may be unrelated dissolution features whose distri-
1993; Tinker, 1998). Total paleobathymetric relief bution was influenced by stratigraphic permeability
from the shallow Tengiz platform to the start of variations.
the upper-slope microbial boundstone at the edge
of the outer platform was at least tens of meters Upper- to Middle-slope Facies
(Weber et al., 2003) and probably higher. Apparent The upper slope consists mainly of in-situ microbial
thickening in the outer platform of the uppermost boundstone and boundstone breccia. Similar facies
Serpukhovian fourth-order cycles, plus the limited are recognized from the Asturias margin (Bahamonde
presence of subaerial exposure at the tops of outer- et al., 2000; Della Porta et al., 2003) as mentioned
62 / Collins et al.

FIGURE 5. Outcrop analogs with facies and depositional relationships similar to the Tengiz outer platform, rim, and
flank. Like Tengiz, the platform margins illustrated are characterized by dipping, basinward-deepening platform facies
that pass downward into upper-slope, microbially influenced boundstones and foreslope facies with reef-derived breccias.
Based on these profiles, the structurally raised rim visible on Tengiz seismic data (see Figure 3) is largely an artifact
of postdepositional differential compaction (see text for discussion). Microbial boundstone types A, B, and C noted in the
Asturias example closely match boundstone types recognized in Tengiz cores.

previously, from Triassic platforms in Italy (Blendin- micritic to peloidal fabrics (type A) or irregular lami-
ger, 2001) and from Famennian platforms in the nar fabrics and amalgamated semiconcentric laminar
Canning basin (Playford, 1984; Kerans et al., 1986; masses (type B). A third boundstone type (type C),
Stephens and Sumner, 2003). Tengiz cores and FMI characterized by massive to peloidal fabric with com-
logs suggest that the in-situ boundstone interval is mon skeletal grains (bryozoans, algae, forams, ostracods,
perhaps 150– 200 m (492–656 ft) thick. In-situ mi- pelecypods, crinoids, corals, and sponges), represents
crobial boundstones (Figure 7) are light colored in the transition between upper-slope and outer-platform
core, with textures that include relatively featureless facies. Small- to medium-size fenestrae filled with banded
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 63

FIGURE 6. Serpukhovian outer-platform facies. Breccias composed of platform-derived clasts (A) or colonial coral
fragments (B) occur toward cycle bases. The breccia in (A) contains cement and clotted fabric of possible microbial origin,
whereas the matrix in (B) consists of coarse crinoidal rudstone. Brachiopod mud-mounds (C) as much as a few meters
thick also sometimes occur near cycle bases. The middle and upper parts of typical outer-platform cycles contain
abundant packstone and wackestone with large crinoids, brachiopods, and solitary corals (D), as well as intervals with
dipping beds (E), indicating deposition on a paleoslope.

cements are commonly abundant, and larger cement- subangular boundstone clasts with stylolitic (welded
lined cavities filled with internal sediment or skeletal breccia) or fracture (mosaic breccia) contacts are found
accumulations are observed. Boundaries between mi- in the Tengiz middle to upper slope (Figure 8) and
crobial masses are sometimes formed by networks of may be analogous to the Asturias intact breccias. The
elongate cavities that are also frequently cement filled. matrix consists mainly of several generations of calcite
The boundstone interval at Tengiz could be ex- cement, secondary microbial encrustations, micro-
tended down to more than 300 m (984 ft) if certain bial cement, and minor amounts of platform-derived
types of breccias are included. The upper slope inter- skeletal sediment, thin-bedded marly volcanics, or
val in the Asturias margin is comparable in thickness argillaceous carbonate. Very little open primary pore
and includes both in-situ microbial boundstone, and space was observed in these breccias.
nearly in-situ boundstone breccia generated by rap- Formation MicroImager logs over upper-slope mi-
id marine cementation of incipient boundstone fail- crobial boundstone and breccia intervals show little
ure (Bahamonde et al., 2000; Della Porta et al., 2003). evidence of cyclicity, suggesting that sea level varia-
To date, similar breccias held together by thick ma- tions that affected the platform had little or no effect
rine cements have not been observed in abundance on production of microbial boundstone. As a result of
at Tengiz, although core coverage is still limited. deposition below the range of sea level fluctuations
However, breccias consisting of irregularly shaped to and productivity across a wide range of water depths,
64 / Collins et al.

FIGURE 7. Serpukhovian microbial boundstone facies, shown in order of increasing paleodepth. (A) Type C boundstone
(skeletal-micritic) with corals and crinoids, and numerous cement-filled, irregular-shaped fenestrae. The matrix is
skeletal rich with micritic to peloidal microbial texture. (B) Type C boundstone containing numerous bryozoans (arrows)
and thick-banded cements of probable marine origin. (C) Type B boundstone (laminar cemented), characterized by
irregular to rounded, concentrically laminated microbial masses. These boundstones contain as much as 75% cement.
(D) Type A boundstone (peloidal-micritic), featuring a dominantly peloidal to micritic matrix, scarce fauna, and numerous
fenestrae or tubular voids.

including subphotic depths, the boundstones were The matrix contains variable amounts of bioclastic
apparently able to prograde continuously seaward debris, fine-grained carbonate, argillaceous mudstone,
during all stages of sea level rise and fall. and volcanic ash. The most visible skeletal components
are large crinoids and brachiopod shell fragments
similar to outer-platform facies. Relationships be-
Middle- and Lower-slope Facies tween clast size, clast packing, matrix grain size, and
The middle and lower slope consists of mainly de- matrix composition have not been fully evaluated,
trital facies (Figure 9), including sedimentary bound- but they appear to be poorly correlated. Clasts are
stone breccias; coarse skeletal rudstones; fine-grained dominantly upper-slope microbial boundstone frag-
breccias; massive or poorly bedded allochthonous ments, although platform-derived clasts and reworked
grainstones; and thin-bedded to laminated grainstones, slope clasts are increasingly observed in breccias near
packstones, wackestones, mudstones, argillaceous car- the base of the section. Some detrital breccias are mas-
bonates, volcanic ashes, and cherts. Clast size and pack- sive and unbedded over very thick intervals, whereas
ing varies in the breccias, with textures ranging from others are layered and contain intervals of bedded,
matrix-supported floatbreccia to welded breccia with grainy periplatform facies. Formation MicroImager logs
stylolite-bounded clasts (Figure 9D – F). Clast sizes again do not indicate an ordered cyclicity that can be
range from less than 1 cm (0.4 in.) to several meters. related to the platform sequences.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 65

FIGURE 8. Upper- to middle-slope diagenetic and sedimentary breccias. (A) Fractured microbial boundstone healed
by banded, probably marine cements. (B) Partly cemented mosaic breccia containing earlier cemented voids dissected
by later fractures and voids containing bitumen and equant calcite spar (arrows). Larger open voids are mostly a product
of leached cement fills in the secondary breccia. Note also that the early, cemented fracture in (A) is also intersected by
a later, bitumen-filled fracture. (C) Mosaic to welded breccia containing microbial boundstone clasts with stylolitic contacts.
(D) Sedimentary floatbreccia with irregularly shaped boundstone clasts, and crinoid-rich, skeletal-lithoclast rudstone
matrix.

DEPOSITIONAL HISTORY OF THE breccia is the dominant facies present in rim and
TENGIZ RIM AND FLANK flank cores recovered to date; therefore, its distribu-
tion is a critical element of reservoir prediction. This
The additional facies control provided by FMI logs requires a deeper understanding of depositional and
(Figure 10) proved helpful in interpreting the rim failure processes at the Tengiz margin.
and flank regions. Attempts to extrapolate fourth- The analogous Asturias and Capitan platforms are
order platform correlations into the rim and flank on also associated with middle- and lower-slope breccia
ordinary wire-line logs proved difficult because of the deposits. The Asturias breccias occupy a basinward-
presence of apparently nonstratiform porosity vari- thickening middle- to lower-slope wedge (see Figure 5),
ations and diagenetic effects manifested as anoma- whereas the Capitan margin is associated with both
lous radiogenic responses on gamma-ray logs. Addi- middle- and lower-slope breccias and basinal debris
tionally, rim and flank stratigraphy is complicated by flows out in front of the shelf (Melim and Scholle,
large-scale failure and mass wasting because of insta- 1995). These dominantly progradational margins have
bility, both during and after deposition. Seismic and maximum upper-slope angles of 30 – 358 (Tinker,
biostratigraphic control indicates that the interval con- 1998; Della Porta et al., 2003; Kenter et al., 2005).
tains breccias of Serpukhovian age as well as Bashkirian- Slope angles for the Tengiz rim and flank are known
aged breccias with Serpukhovian clasts. Boundstone from structural and seismic profiles of the Bashkirian
66 / Collins et al.

FIGURE 9. Middle to lower slope and debris apron facies. Middle- to lower-slope facies include coarse skeletal (crinoid-
brachiopod) rudstone (A), skeletal wackestone to packstone with dipping beds (B), and thin-bedded volcanics,
mudstone, packstone, and fine grainstone (C). Debris apron facies include coarse breccia with meter-size boundstone
clasts (D), packbreccia with platform-derived, crinoid-rich matrix (E), and float- to packbreccia interbedded with
allochthonous grainstone and packstone (F). Note the relatively low dip angles in (F).

surface and from dipmeter data. In the steeper sloped seismic profiles, even considering geometries created
upper flank, the Bashkirian is thin, consists of mostly by depositional or compactional drape over under-
fine-grained deeper water facies, and, thus, forms a lying backstepping platforms. This shape varies sys-
prominent seismic reflector that closely approximates tematically around the perimeter of the buildup and,
the Serpukhovian profile. Measured angles vary be- combined with facies obtained from core and FMI
tween 20 and 258 for apparent accretionary slopes, logs, allows the rim to be divided into subregions gov-
whereas angles approaching 308 are measured in areas erned by different depositional processes.
of obvious detachment or failure. These measure-
ments represent angles determined from a compacted
body of rock and do not necessarily reflect original Allochthonous and Accretionary Rim Sectors
depositional dips. Because of the chaotic nature of The rim and flank area consists of two distinct
boundstone and breccia facies at Tengiz, FMI dip mea- regions, the allochthonous and accretionary sectors
surements commonly produce inconsistent results, (Figure 11). The allochthonous sector refers to the
but some intervals indicate that preserved bedding southwestern and northwestern parts of the rim. It
dips of 25 –308 in the upper slope may be typical, and is characterized by a lower, distally thickened apron
angles of 35 – 458 occur locally. Directionality is vari- that onlaps the Tournaisian and Visean platforms,
able, however, and dips do not always point directly which itself appears to be downlapped by a narrow
away from the platform. Multiple azimuth changes progradational wedge (Figure 12A). The raised rim is
have been recorded vertically in a single well or in reduced or, in some places, completely absent along
comparing two nearby wells. Formation MircoImager this trend (Figure 12B). The accretionary sector re-
logs also suggest that facies distribution is more com- fers to the northern, northeastern, and southeastern
plex than expected for simple progradation. Whereas regions of the rim, where it is characterized by a con-
the Asturias and Capitan margin profiles have regu- spicuous raised rim crest and a considerably wider
lar shapes, the Tengiz flank has an irregular shape on progradational wedge (Figure 13A).
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 67

FIGURE 10. Formation MicroImager log to core facies calibrations. FMI facies, indicated by different colors in the depth
columns (depths are in meters), include boundstone-breccia (BBr), boundstone-horizontal fabric (Bhz), breccia, rud-
stone, grainstone, and laminated fine-grained (lam-FG). BBr facies, characterized by a chaotic image lacking horizontal
layering (C – E), includes both boundstones and breccias from core, which, except occasionally, were not distinguish-
able on FMI logs. Bhz facies corresponds to poorly bedded outer platform skeletal rudstone and type C microbial
boundstone (F). FMI rudstone is characterized by nodular texture and some horizontal layering (A, D) and includes
lithoclast-skeletal rudstones, fine-grained breccias, and poorly bedded skeletal packstone, wackestone, and floatstone
from core. Formation MicroImager grainstones, identified by a low-contrast image with faint layering (B), correspond
to generally tightly cemented allochthonous grainstones. Lam-FG facies, marked by frequent high-contrast layers (A),
are equivalent to thin-bedded, deep-water volcanics, mudstones, and grainstones in core. In some intervals, image quality
is adversely affected by bitumen, which produces a snowy texture overprinted on the facies (C).

In the allochthonous sector, the lower apron con- parently bottomless faults, but also to a prominent mid-
sists of bedded to massive boundstone breccia in- slope bulge. The lateral continuity of this latter fea-
terlayered with bedded platform-derived sediments, ture (see Figure 11) could be accounted for by slight
suggesting relatively continuous failure of a bound- downslope movement of upper-slope material or al-
stone margin. In the northern half of the accretion- ternately by downlap of a late-stage progradational
ary sector, the flank slope is typically irregular. This is wedge onto a protruding debris apron (Figure 13B),
partly caused by a series of local slump scars and ap- analogous to the allochthonous sector. Seismic data
68 / Collins et al.

FIGURE 11. Dip magnitude on


the top of the reservoir (Bashkir-
ian surface) measured from 3-D
seismic data, showing the extent
of the accretionary and alloch-
thonous sectors. The map shows
a thin band of elevated dip
(dashed white line) that is di-
rected toward the platform inte-
rior caused by differential com-
paction between platform facies
and upper-slope microbial
boundstones and indicates the
position of the platform margin
prior to progradation. The zone
of low dip between this band
and the slope break is the raised
rim and indicates the limit of
progradation (possibly modified
by failure processes). The bound-
ary between the accretionary
and allochthonous sectors occurs
where the band of compactional
dip intersects the slope break
(white arrows). In the accre-
tionary sector, the slope break is
marked by several local scarps
indicative of localized failure
(yellow lines). In the allochtho-
nous sector, the raised rim is thin
or absent, indicating either nearly
complete rim failure or a poorly
developed progradational wedge.
The locations of seismic profiles
in Figures 12–15 are also shown.

suggest the existence of an early debris apron in the primary microbial production may have varied around
accretionary sector in another way as well. The north- the platform. In the accretionary sector, elevated pro-
ern and southern margins of the Tengiz platform rep- ductivity could have resulted in faster accumulation
resent transitional regions between the allochthonous of a debris apron and, thus, earlier initiation of upper
and accretionary sectors. Both areas are characterized wedge progradation. In a lower productivity alloch-
by an upper wedge that gradually widens eastward thonous sector, the debris apron required a longer
while maintaining a downlapping relationship to lat- time to accumulate and was perhaps still in progress
erally continuous debris apron deposits (Figure 14). at the end of the Serpukhovian, accounting for the
The accretionary and allochthonous sectors dem- more limited upper wedge development.
onstrate internal and comparative differences with Tectonic activity is a possible external mechanism
respect to the total volume of rock in the rim and flank, that could also account for rim asymmetry. Rim fail-
as well as the relative proportions in the upper wedge ure could have been restricted to, or enhanced in the
and lower apron. This asymmetry could result from allochthonous sector because of alignment with re-
both extrinsic and intrinsic controls. For example, gional stress regimes or as a result of directly facing
one intrinsic control might be the pre-Serpukhovian pressure waves from the tectonically active southern
profile created by the stacking pattern of Tournaisian margin of the Precaspian Basin (see Figure 1). Whether
and Visean platforms. In areas of significant backstep- the apron is mostly an early or a late feature is there-
ping, a bench or gentle slope might have provided fore still somewhat uncertain because of these pos-
stability for progradation (see Figure 13A), whereas in sibilities. The significant role of rim failure and the
areas where platform stacking is more vertical, insta- timing uncertainty precluded erecting a stratigraphic
bility and rim failure might dominate. Alternatively, framework in the rim and flank that can be specifically
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 69

FIGURE 12. Seismic profiles


from the allochthonous rim
sector. In (A), a slightly raised
rim indicates a narrow pro-
gradational wedge of upper-
slope microbial boundstone
that formed after deposition
of an extensive debris apron.
The figure in (B) shows a re-
gion where the raised rim is
absent, indicating failed pro-
gradation. Formation Micro-
Imager logs from T-7040
indicate the upper wedge de-
posits at this location most
likely consist of coarse brec-
cia. See Figure 11 for profile
locations.

forms a mound-shaped pile


that is almost continuous with
the distal apron (Figure 15).
To date, the upper wedge is
penetrated by 14 wells with
core slabs and/or FMI images
(T-463, T4346, T-4635, T-70,
T-7040, T-5056, T-5857, T-7252,
T-6457,T-4556,T-5059, T-5454,
T-7450, and T-7052). These
wells indicate facies consistent
with a prograding boundstone-
cored wedge in areas of accre-
tion in both sectors and facies
dominated by breccia in areas
of detachment in the alloch-
thonous sector (Figure 16).
Specific cored examples in-
clude cyclic shallow platform
tied to Serpukhovian and Bashkirian fourth-order overlying deeper water outer-platform facies (T-4556
platform cyclicity; however, core control, FMI facies, and T-7052); upper-slope in-situ boundstone facies
biostratigraphic data, and, to a lesser extent, wire-line (T-7450 and T-5056); and upper-slope boundstone
logs were still used to separately constrain the dis- and welded-to-mosaic breccia from a detached but
tribution of facies in the lower apron and the upper relatively intact upper wedge (T-4635, Figure 15A).
wedge as a result of their relative spatial separation. Other cored examples include middle-slope breccias
(T-5056, T5857, and T-7252), and middle- to lower-
Upper Wedge Facies Distribution slope clast- and matrix-supported boundstone breccia
From previous discussions, the raised rim effect interbedded with platform-derived crinoid-brachiopod
is interpreted to indicate the presence of mechani- rudstone to packstone (T-463). The Bashkirian in the
cally rigid upper-slope microbial boundstone facies upper wedge consists of shallow-platform facies in
in the upper wedge. In places along the allochtho- the raised rim area and, on the basis of mainly short
nous sector, the upper wedge is partly or completely cores and core chips from older wells, deeper water
detached from the platform, and the raised rim effect facies composed of laminated to thin-bedded mud-
is absent. In some cases, the wedge appears to have stones, packstones, fine grainstones, and volcanics in
detached more or less intact, whereas in others, it the flank.
70 / Collins et al.

FIGURE 13. Seismic profiles


from the accretionary rim
sector. The flank profile in
(A) is smooth and lacks high-
amplitude internal reflections.
Here, a bench formed by a
backstepping of the Visean A
and B platforms may have
provided stability for early
upper wedge progradation.
The flank profile in (B) shows a
distinct midslope bulge. For-
mation MicroImager data
from T-5059 and T-5660 dem-
onstrate that the upper part of
the Serpukhovian interval in
the bulge contains a prograd-
ing boundstone sequence,
whereas the lower part con-
tains bedded boundstone
breccias (see Figure 17). The
change occurs near a zone of
high-amplitude, subparallel
reflectors on the seismic line,
which have been interpreted
to indicate areas in the accre-
tionary sector where deposi-
tion of early debris aprons
filled in some of the accom-
modation space as a prerequi-
site to upper wedge pro-
gradation. See Figure 11 for
profile locations.

boundstone breccia and


platform-derived grainy fa-
cies, but which also indicate
significant lateral variations
in bed thickness, clast size,
and amount of platform-
derived grainy material ver-
Lower Apron Facies Distribution sus breccia. In two separate cored intervals from
The lower apron is penetrated by four wells in the T-6337, for example, both clast- and matrix-supported
allochthonous sector with cores and/or FMI data breccias are frequently interbedded with fine-grained
(T-16, T-6337, T-3938, and T-3948). Assuming that an periplatform mudstone, packstone, and grainstone.
early apron is also present in the accretionary sector, Overall clast size is larger compared to the T-463 core
five other wells potentially have representative core from the upper wedge, but the platform-derived frac-
and/or FMI data (T-47, T-5963, T-5059, T-5660, and tion is finer grained. By contrast, a long, cored inter-
T-6261). In particular, the FMI facies from T-5660 and val at the T-3948 core consists almost entirely of
T-5059, combined with three separate cored inter- poorly bedded boundstone breccia made up of very
vals from nearby T-5056 in the raised rim, indicate a large clasts, with very few interbeds of fine-grained
possible distinction between apron breccias and upper platform-derived material. The T-3938 FMI log sug-
wedge facies in the accretionary sector (Figure 17). gests that the apron interval there also consists of
Apron facies are best known from cores in the al- massive boundstone breccia with few interbeds of
lochthonous sector, which reveal that, in general, platform-derived facies. The lateral extent of this mas-
the apron consists of mixtures of upper-slope-derived sive, coarser apron breccia facies appears to coincide
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 71

FIGURE 14. Seismic profiles


from the northern (A) and
southern (B) rim and flank
regions. These areas are tran-
sitional between the alloch-
thonous and accretionary sec-
tors and show both a lower
debris apron and significant
upper wedge progradation,
suggesting that variations in
wedge development did not
necessarily occur at the expense
of the apron. See Figure 11 for
profile locations.

sent there. Cores across the


base of rim and flank brec-
cias have been recovered at
T-6337, T-3938, T-3948, and
T-7252. Formation Micro-
Imager logs from these and
other wells demonstrate that
the breccias were deposited
abruptly on top of a con-
tinuous interval containing
mainly thin, dipping beds
of deeper water facies. Bio-
stratigraphic analyses from
the cores indicate that in both
sectors, the onset of breccia
deposition occurred close
to the Visean–Serpukhovian
boundary.
Biostratigraphic data also
demonstrate that Bashkirian-
aged sediments containing
Serpukhovian boundstone
clasts occupy as much as
300 m (984 ft) at the top of
with chaotic seismic character and, therefore, may be the apron breccias in the allochthonous sector, indi-
regionally mappable within limits imposed by image cating that significant postdepositional rim failure
quality (see Figure 16). occurred in that region. The Bashkirian interval ap-
Apron breccias in the allochthonous sector con- pears to be associated with a distinctive pattern on
tain increasing numbers of transported slope and gamma-ray and resistivity logs (Figure 18), and well
platform clasts toward the base of the interval. Cores correlations suggest that Bashkirian breccias are vari-
from T-3938 and T-3948 demonstrate that some- able in thickness but laterally extensive (see Figure 16).
what finer grained breccias with abundant platform The apparent absence of equally thick Bashkirian de-
and slope clasts are associated with a resistivity de- posits or Bashkirian-aged breccias in the accretionary
crease on petrophysical logs. This low-resistivity zone sector is an important distinction that perhaps indi-
is not observed on logs from the accretionary sector, cates that relative depositional stability was achieved
but because this could be caused by a change in po- in this region prior to the end of the Serpukhovian.
rosity or other diagenetic effect, its absence does not Cross sections illustrating the depositional concepts,
necessarily indicate that similar breccias are not pre- core and FMI facies control, and typical porosity and
72 / Collins et al.

FIGURE 15. Seismic profiles illustrating


different modes of upper wedge fail-
ure. The figure in (A) shows a com-
pletely detached but largely intact upper
wedge. Core from T-4635 contains mainly
type C boundstones and breccias, indi-
cating original deposition in shallower
water. The figure in (B) shows a narrow
upper wedge with a downslope bulge,
indicating possible partial failure. The
debris apron in this area consists of
poorly bedded boundstone breccia, as
indicated on FMI logs from T-3938.
Biostratigraphy from drill cuttings re-
vealed a thick Bashkirian-aged section
at the top of the apron breccias. The
figure in (C) shows local fault detach-
ment in the upper wedge. In addition,
chaotic seismic character in the debris
apron is separated from the upper
wedge slope by a zone of relatively
continuous dipping reflections from a
distinct lower slope facies. T-3948 cores
reveal that the apron at this location
consists of massive boundstone breccia
with very large clasts. See Figure 11
for profile locations.

DIAGENESIS

As a long-lived shallow platform,


Tengiz was expected to have a com-
plex distribution of pore types be-
cause of its multifaceted stratigraphic
hierarchy and associated facies vari-
ations, repeated exposures to ma-
rine, meteoric, and shallow burial
diagenesis, and, because of the age
range involved, the influences of
both greenhouse and icehouse glob-
al themes. Core and thin-section ob-
servations, combined with prelimi-
nary geochemical analyses, provide
insights into processes of porosity
creation, destruction, and preserva-
tion in outer platform, rim, and flank
reservoirs and also illuminate rela-
tionships that determine reservoir-
quality distribution in the rim and
flank. These relationships suggest
that present-day reservoir-quality var-
gamma-ray expression of the lower apron and upper iations are strongly influenced by a late diagenetic
wedge in the Tengiz allochthonous and accretionary overprint involving fracturing, dissolution, and bi-
rim sectors are provided in Figures 18 and 19. tumen cementation.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 73

FIGURE 16. Facies distribu-


tion in the rim and flank. The
Visean A outline indicates
the platform configuration
just prior to Serpukhovian
progradation. The raised rim
area (upper wedge) repre-
sents the preserved extent of
Serpukhovian platform pro-
gradation. The outer limit of
the upper-slope boundstone
was determined by an as-
sumed vertical thickness of
about 200 m (656 ft) below
the slope break. The lower
apron contains Serpukhovian
and Bashkirian coarse brec-
cias, mapped from seismic
character and well-log corre-
lations. Serpukhovian brec-
cias occur on both sides of the
platform, whereas Bashkirian
apron deposits are generally
restricted to the allochtho-
nous sector, indicating depo-
sitional instability in that area
throughout the Serpukhovian
and Bashkirian. Facies control is
indicated for rim and flank
wells only.

tant for platforms like Ten-


giz with considerable verti-
cal height.
Early voids commonly con-
tain several cement genera-
tions. These cements, starting
with inclusion-rich calcites
of probable marine origin and
culminating with medium-
Early Dissolution, Cementation, and to coarse-grained clear equant calcite of burial origin,
Bitumen Formation form a sequence that destroyed much of the early po-
Early pore types consist of primary and early dia- rosity (Figure 20). The residual pore space commonly
genetic matrix pores and larger voids from at least contains bitumen, either as a final fill phase or as beads
one significant dissolution event. Existing cores sug- and partial linings on the equant calcite. The early
gest that this early dissolution was more prevalent in cement sequence is therefore referred to collectively as
the outer-platform and upper-slope facies compared the prebitumen sequence. Petrographic relationships
to the middle and lower slope. Early dissolution may between bitumen and equant calcite (Figure 20C)
have occurred during platform margin karst events suggest that they were emplaced approximately con-
associated with major sequence boundaries or as a temporaneously. Equant calcite and bitumen occur
result of Kohout-type convection driven by large fluid- in some dissolution features as the only phases, in-
density contrasts between platform-top groundwa- dicating a second fracturing and dissolution event
ters and seawater (Kohout et al., 1977; Saller, 1984; prior to bitumen emplacement. Fluid-inclusion ho-
Sanford et al., 1998). Such contrasts, controlled by mogenization temperatures indicate that the equant
temperature and salinity, could be especially impor- calcite was precipitated at temperatures in the 80–958C
74 / Collins et al.

FIGURE 17. Selected core and FMI images from T-5660 (top), showing facies representing the upper wedge sequence
overlying a possible debris apron sequence in the accretionary sector. The upper wedge sequence consists of distal bedded
rudstone and fine breccia toward the base (E, F), which are overlain by successively coarser breccias (B – D). Above the
cored interval, the section consists of massive boundstone or breccia (A). The apron (G – J) consists of interbedded lam-FG,
rudstone, and breccia FMI facies (the latter presumably containing microbial boundstone clasts). For definitions of FMI
facies, see Figure 10. The cross section (bottom, see Figure 16 for well locations) shows that the inferred apron section
in T-5660 is associated with reduced radioactivity on the spectral gamma-ray (SGR) log compared to the wedge interval
and may be correlated to nearby wells on this basis. Other log curves shown include computed gamma-ray (CGR,
calculated with uranium subtracted) and deep resistivity (DRES).
FIGURE 18. Facies and stratigraphic relationships in the allochthonous rim sector. Wells are shown in approximate order of increasing distance from the edge
of the platform (see Figure 16 for well locations). The thick, allochthonous Bashkirian section at T-6337 is supported by biostratigraphic analysis of the cored
interval. Comparison of spectral gamma-ray (SGR) and computed gamma-ray (CGR, calculated with uranium subtracted) logs indicates uranium enrichment
throughout the rim and flank interval. Note the low-resistivity zone on the deep resistivity (DRES) curves in T-3938 and T-3948 at the base of the apron. Also note the
similarities in the FMI facies and gamma-ray logs in T-6337 and the interpreted apron sections in T-5660 and T-5059 from the accretionary sector (Figures 17, 19).
For FMI facies definitions, see Figure 10.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 75
76 / Collins et al.

FIGURE 19. Stratigraphic and facies relationships in the accretionary rim sector. Wells are shown in order of increasing distance from the platform (see Figure 16
for well locations). Raised rim wells contain Serpukhovian middle-slope breccia and upper-slope boundstone overlain by upper Serpukhovian to Bashkirian outer-
and shallow-platform facies. Wells closer to the platform (T-4556 and T-7450) are associated with reduced uranium enrichment, as shown by comparing the
spectral gamma ray (SGR) with the computed gamma ray (CGR, calculated with uranium subtracted). Uranium enrichment also tends to be higher in upper
wedge boundstone breccia and associated facies in flank wells, above the apron facies interpreted in T-5059 and T-5963. FMI facies and deep-resistivity (DRES)
logs are shown for some wells. For FMI facies definitions, see Figure 10.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 77

range (Appendix 2); thus, prebitumen cementation tures that are mainly open and lack cement are present,
was completed shortly before hydrothermal tempera- representing an even later generation following bitu-
tures were achieved in the reservoir. Equant calcite is men deposition, and at least one set of fractures ap-
a volumetrically important cement in the rim and pears to be aligned with the present-day stress regime
flank and is referred to as 908 calcite on the basis of the (D. G. Carpenter, 2005, personal communication). Partly
fluid-inclusion data. open and open fractures are commonly associated
with local solution enlargement along fracture walls,
Fractures and Later Dissolution (Corrosion) and with bitumen cementation and corrosion in the
Larger fractures observed in core can be separated surrounding rock matrix. Matrix corrosion occurs
into sets according to their fill phases. Earlier frac- as microporosity in the micritic components of the
tures, like early dissolution voids, are extensively filled rock, for example, skeletal grains and microbial fa-
with prebitumen cement phases. Later fractures con- brics (Figure 21A). In facies containing micritic ma-
tain only bitumen, or equant calcite accompanied by trix, microporosity accompanied by small vugs can
bitumen, and are filled to partly open. Systematic frac- sometimes be extensively developed. In addition to

FIGURE 20. Earlier stages of diagenesis in the Tengiz rim and flank. Voids are filled by a prebitumen cement sequence
(A and B) that includes inclusion-rich, probably marine calcite (1), prismatic to equant zoned calcite and inclusion-rich
dolomite (2), and clear blocky equant calcite (3). This sequence is commonly followed by bitumen (4). Prebitumen cements
filled much of the primary pore space, as well as porosity created during early diagenetic dissolution and fracturing.
Some voids are filled only with equant calcite and bitumen (C and D), indicating a separate, later fracturing and dissolution
event. Even later dissolution is suggested by corrosion along crystal growth planes (5) in the equant calcite. Fluid-
inclusion homogenization temperatures of 80 – 958C for the equant calcite indicate the existence of prehydrothermal
conditions in the reservoir prior to bitumen formation. Scale bars = 1 mm.
78 / Collins et al.

matrix corrosion adjacent to solution-enlarged fractures, microporosity and bitumen-free microporosity occur in
bitumen and corrosion are noted between cements the same thin section are common (Figure 21B), and in
and walls of partly filled fractures. Bitumen and cor- some cases, dissolution is observed that appears to post-
rosion are also sometimes observed within inclusion- date bitumen formation (Figure 21C). This suggests
rich prebitumen cements or along cleavage and growth that some matrix corrosion probably accompanied
planes of individual cement crystals. bitumen emplacement, and that some dissolution may
Bitumen and matrix corrosion are both observed also have occurred later along the same pathways.
to decrease away from vugs and enlarged fractures, However, many other instances indicate mainly pas-
suggesting that bitumen, corrosion, and large-scale sive fill of open void space by bitumen, accompanied
dissolution may be genetically linked. Petrographic by only minor micrite corrosion.
observations indicate that significant corrosion and Other postbitumen diagenesis includes fractures,
solution enlargement postdates the early cement se- noncarbonate cement such as fluorite and quartz (based
quence. A key question is whether most of this occurred on fluid-inclusion temperatures; see Appendix 2), and
before, during, or after bitumen formation or a combina- possibly some calcite spar (Figure 21D). The existence
tion of the three. Examples where bitumen-associated of syn- and postbitumen cementation and dissolution

FIGURE 21. Bitumen-related and postbitumen diagenesis in the Tengiz rim and flank. Bitumen is dominant in some
voids and fractures (A, 4) and is commonly associated with corrosion of microbial fabrics (A, 5). Some micrite corrosion
appears to be free of bitumen (B, 6) and possibly postdates bitumen formation. Postbitumen diagenesis also includes
dissolution (selective dissolution of dolomite rhombs in a partial dedolomite, C) and late fracturing (D). The example
in (D) also shows postbitumen calcite cement. A key question in the Tengiz rim and flank is how much postbitumen
dissolution and fracturing contributes to present-day reservoir quality and distribution. Scale bars = 1 mm, except
where indicated.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 79

implies that an active hydrothermal system was prob- saturations are mixed (Hill, 1995), or when H2S and
ably a factor in the development of the rim and flank CO2 generated by TSR mix with oxygenated forma-
reservoir. Two scenarios are proposed: hydrothermal tion water in a carbonate reservoir (for example, see
fluid mixing and geothermal convection. Mechanisms Hill, 1990). In the absence of oxygenated formation
for hydrothermal mixing includes fluids with differ- waters, the amount of carbonate dissolution that
ent H2S concentrations or thermochemical sulfate occurs in the presence of H2S and CO2 is controlled
reduction (TSR) reactions between hydrocarbons and by pH of the formation waters and the presence of
sulfate minerals. These possibilities are suggested by: iron oxides or aqueous sulfides (Stoessell, 1992). Car-
bonate dissolution commonly occurs because of car-
1) the elevated H2S content (13%) of hydrocarbons bonic and/or sulfuric acid generated during these
in the reservoir reactions (Hill, 1987; Machel, 1987, 1989). Once H2S
2) the presence of known TSR byproducts in the saturation became established in the reservoir, either
reservoir by migration of H2S and CO2 gases or after migration
3) the thermal regime, estimated from fluid-inclu- of hydrocarbons, additional dissolution may have
sion studies, under which these products were occurred as a result of periodic mixing with external
precipitated hydrothermal fluids during even later burial.
4) the presence near or within the reservoir of The high burial temperatures, elevated H2S levels,
geological conditions and ingredients favorable and the presence of bitumen, dedolomite, and anhy-
for TSR reactions. drite suggest the possible involvement of TSR in dia-
genesis of the rim reservoir. In the Tengiz region, po-
Geothermal convection is supported by two- tential reactants include hydrocarbons generated from
dimensional numerical flow models of Tengiz using Carboniferous to Permian subsalt source rocks adjacent
temperatures derived from basin modeling, known to the buildups (Lisovsky et al., 1992) and an anhy-
thermal rock and fluid properties, and a simplified drite layer at the base of the thick Permian halite for-
layering scheme based on bulk reservoir permeabil- mation that encases the buildups. Figure 23 shows two
ities. It is also possible that both convective and H2S- possible regions where reactions might have occurred:
related processes contributed to rim diagenesis and
reservoir quality at different times or at the same time. 1) basinal areas between platforms where hydrocar-
bons were being generated; in this scenario, car-
Hydrothermal Fluid Mixing and bonate dissolution occurred after migration of
H2S-related Diagenesis H2S and CO2 from reaction sites into the reservoir
According to basin models, the Tengiz reservoir 2) in the reservoir itself, with carbonate dissolution
attained hydrothermal temperatures during the early resulting from reactions that occurred when
stages of post-Triassic rapid subsidence (Figure 22). liquid hydrocarbons entered the reservoir
During and after this time, basin-generated hydro-
thermal fluids may have periodically charged the res- The latter mechanism is possible because the ab-
ervoir, mixing with formation fluids. Thin sections sence of Moscovian–Artinskian volcanics on the up-
and quantitative x-ray diffraction data from core plugs per slope of the Tengiz flank places the anhydrite layer
reveal the presence of anhydrite, dolomite, chalcedony, in direct contact with the reservoir. It has the advan-
quartz, barite, fluorite, fluorapatite, and pyrite as po- tage of also accounting for, as a common by-product
tential hydrothermal species in the rim and flank. of TSR, at least some of the ubiquitous bitumen in the
Most of these minerals exist in quantities less than rim and flank. Reservoir temperatures of 80 –958C
4% by weight. Species that may have a hydrothermal achieved prior to bitumen emplacement (based on
origin and have been measured as much as 50% and prebitumen equant calcite) indicate that bitumen
higher by weight in individual samples include fluo- formation probably occurred as the reservoir entered
rapatite, anhydrite, and dolomite (anhydrite and do- the hydrothermal regime.
lomite may also have other origins). Chalcedony and Carbonate reservoirs exposed to TSR commonly
quartz are also locally abundant. Homogenization contain anhydrite, dedolomite, sulfide minerals, ele-
temperatures of 110 – 1258C have been obtained from mental sulfur, and uranium and iron enrichment (Hill,
examples of quartz and fluorite (Appendix 2). 1990, 1995). In particular, anhydrite and dedolomite
Significant calcium carbonate dissolution can oc- of apparent nonevaporative or nonsubaerial origin
cur in reservoirs where fluids with different H2S are commonly observed in cores and thin sections
80 / Collins et al.

FIGURE 22. Temperature and burial-


history plots for the Tengiz rim and
flank reservoir from basin modeling.
The vertical depth range indicates the
suite of samples from different parts of
the reservoir interval used in fluid-
inclusion analyses (Appendix 2). Bitu-
men formation in the reservoir began,
probably under hydrothermal condi-
tions, after cementation by equant
calcite (top) during Mesozoic subsi-
dence. Potential times of burial fracture
generation include stresses generated
during salt diapirism, uplift (approxi-
mately 220 – 175 Ma), rapid Mesozoic
subsidence, and the present-day stress
regime (bottom).

rim and flank (G. D. Jones and Y.


Xiao, 2005, personal communica-
tion), mostly because of the high ver-
tical permeability possible in frac-
tured reservoirs and because of a
relatively large center-to-edge tem-
perature gradient in the buildup
caused by the thick, conductive en-
velope of halite around the exterior.
Once convection cells are estab-
lished in a closed system, areas of net
dissolution and net cementation can
occur while maintaining average
saturation equilibrium in the cell
with respect to the reservoir. Be-
cause of the inverse relationship at
from the Tengiz rim, and GR logs from rim and flank constant pH and pCO2 between calcium carbonate
penetrations typically show uranium enrichment solubility and temperature, dissolution occurs where
(see Figures 18, 19). The presence of dedolomite sug- saturated or nearly saturated fluids are cooling (rising
gests the involvement of fluids with excess Ca2+, pos- cell limbs), and cementation occurs where such fluids
sibly from TSR-related anhydrite dissolution. are warming (descending limbs). However, this pro-
cess results in mass transfer of carbonate from one
Geothermal Convection area to another, and closed-system cells may, over
Groundwater circulation can be important in the time, be forced to readjust in shape or shut down in
diagenetic history of carbonate reservoirs, particu- response to changing permeability patterns.
larly for isolated platforms (Jones et al., 2004). Platform Numerical models further indicate that convec-
margins are favorable sites for convective circulation tion cells, if they remain active long enough, can
during burial, as well as during deposition. Deposi- transfer sufficient calcium carbonate to significantly
tional or early postdepositional convective processes affect porosity in the reservoir (G. D. Jones, 2005, per-
were previously briefly cited with respect to early sonal communication). Several factors may have facil-
dissolution at the Tengiz margin. Geothermal con- itated long-term active convection in the rim and flank
vection can occur during burial under certain condi- of Tengiz. The absence of Moscovian to Artinskian
tions (Machel and Anderson, 1989; Morrow, 1998; sediments on the upper slope of Tengiz (Figure 23) may
Wendte et al., 1998; Heydari, 2000). Two-dimensional have allowed continuous seawater exchange for rim
numerical fluid-flow simulations indicate that large- and flank convection throughout an extended period
scale geothermal convection is feasible in the Tengiz following deposition. Second, Tengiz appears to have
FIGURE 23. Evolution of the Tengiz rim and flank high-permeability reservoir, based on data and analyses contained in the text. Prior to burial (A), the reservoir
consisted of fractures and large vugs caused by syndepositional instability and platform margin karsts. Initial large-scale dissolution and cementation patterns
may have been controlled by Kohout convection and by meteoric and mixing processes. During early burial (B), they may have been controlled by geothermal
convection, promoted by still-open fractures and by deposition of thermally conductive halite. Conductivity loss and differential compaction caused by salt
diapirism may have reduced or eliminated geothermal convection (C), although compactional stresses may have rejuvenated the fracture system. Prebitumen
cements had significantly reduced nonmatrix porosity and permeability by the end of this stage. New tectonic stresses again reactivated the fracture system,
possibly facilitating movement and distribution of hydrothermal fluids in the reservoir (D). Diagenesis related to thermochemical sulfate reduction (TSR) is
believed mainly responsible for solution enlargement of open fractures and lost circulation zones (LCZ), either from migration of CO2 and H2S gas, or as a direct
result of the liquid hydrocarbon charge. These fluid charges may also be responsible for enhanced dissolution in adjacent platform facies, forming the
transitional reservoir zone (TRZ). Note that overall fracture density, but not necessarily openness, increased in reservoir over time.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 81
82 / Collins et al.

porosity (fractures, large vugs,


and lost circulation zones) in
some facies, the rim and
flank are considered a dual-
porosity reservoir. Differ-
ent reservoir facies in the
rim and flank can be rec-
ognized based on both ma-
trix and nonmatrix pore
types.

Matrix Reservoir Types


Air permeability data
combined with thin sec-
tions from core plugs were
used to guide the recogni-
tion of matrix reservoir rock
FIGURE 24. Tengiz rim and flank paragenesis. Cementation events are indicated in types in the Tengiz rim and
black, dissolution events are indicated in white, and events that involve mainly re- flank. This data, while not
placement or recrystallization are shown in gray. Approximate absolute timing has quantitatively accurate for
been established for zoned prismatic spar, dolomite cement, and subhedral equant characterization at reservoir
calcite (150 ± 25 Ma; see Figure 22 ) from fluid-inclusion results. Based on petrographic conditions, nevertheless rep-
relationships, these cements were emplaced just prior to bitumen deposition (see
Figure 20) and the time when hydrothermal temperatures were attained in the res-
resent a relatively consistent
ervoir. All other diagenetic products are shown in relative order only, as part of either laboratory test that is useful
the prebitumen or postbitumen sequences. Other phases might be constrained by for distinguishing reservoir
processes related to burial history, which has an implied time scale (see Figures 22, 23). rock types based solely on
rock properties. The classi-
had a complex fracture history (Figure 24). Multiple fication scheme used for analysis of Tengiz plugs was
fracture events and/or repeated reactivation of pre- modified from that of Lucia (1999) and recognizes
vious fractures in the rim and flank may have main- five matrix pore-type classes (interparticle porosity,
tained or increased vertical permeability over time, microporosity, touching vugs, separate vugs, and frac-
allowing geothermal convection to remain active dur- tures) based on independent permeability behavior
ing burial. as a function of porosity. Applied to Tengiz, these
data permit the distinction of three main reservoirs
RESERVOIR QUALITY AND DISTRIBUTION based on matrix properties:

The distribution of reservoir rock types in carbon- 1) The rim and flank reservoir (Figure 25) encom-
ate reservoirs is a product of interactions between passes upper-, middle-, and lower-slope facies
depositional facies and diagenetic processes. Reser- containing mainly microbial boundstone or
voir quality should tend to average out somewhat breccia (but also including some interbedded
when comparing deeper well penetrations or wells platform-derived facies).
from the same environment in the field because these 2) The transitional reservoir zone (Figure 26) con-
wells encounter a wider range of stratigraphic ages, sists of shallow-platform or outer-platform facies
facies, and potential diagenetic themes in the Tengiz with significant bitumen cement and matrix
spectrum. Reservoir quality at Tengiz has so far shown dissolution. This term is used instead of outer
a stronger relationship to specific well locations than platform because the reservoir type is not re-
to facies or stratigraphy. A partial explanation for stricted to that facies. It should not be confused
these geography-based reservoir-quality variations lies with the term ‘‘transition zone,’’ commonly used
in understanding the relative contributions of disso- to describe certain conditions in hydrocarbon
lution, cementation, and fracturing to matrix and non- columns.
matrix porosity in the outer platform, rim, and flank. 3) The central platform reservoir includes the
Because of the significant contribution of nonmatrix central regions of the Tengiz platform that are
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 83

FIGURE 25. General characteristics of the rim and flank reservoir. (A) Local solution enlargement of a partly cemented
fracture-breccia in microbial boundstone. Dissolution is found along the fracture walls and between cement crystals.
Bitumen-filled microporosity (darker color) extends a few centimeters from the fractures, but matrix porosity is otherwise
low. A lost circulation zone is associated with this feature. (B) Microbial boundstone with low matrix porosity and
systematic bitumen-filled hairline fractures. (C) Probable fractured zone in skeletal-lithoclast rudstone with extensive
matrix corrosion, resulting in core rubble. The dark color is caused by pervasive bitumen cementation. (D) Microbial
boundstone (clast in breccia) with bitumen and miroporosity developed along microbial filaments. (E) Microbial
boundstone (clast in breccia) with vugs formed by localized solution enlargement between microbial masses.

relatively unaffected by bitumen plugging and of deeper, outer-platform facies or by the presence of
associated effects (see Kenter et al., 2006). microbial or marine cements (for example, see Figure 6).
Figure 27 shows a more subtle contrast between the
It can be demonstrated that much of the Visean A, transitional reservoir zone and the rim and flank
Serpukhovian, and Bashkirian platform reservoir is reservoir. Because bitumen and dissolution affected
in pressure communication with the rim and flank. both of these reservoirs, the difference is more likely
As shown in Figure 23, the transitional reservoir zone caused by the grainy nature of transitional reservoir
refers to parts of the outer regions of these platforms zone facies, which probably had higher matrix
that may have been altered by the incursion of vari- porosity prior to bitumen and dissolution, in con-
ous diagenetic fluids active in the rim. For example, trast with rim and flank facies that are dominated by
the extent of this region may represent the equili- early-cemented microbial boundstone and bound-
bration limit of corrosive hydrothermal fluids that stone breccia.
charged the reservoir or the innermost edges of active
convection cells. The greatest distinction in matrix Transitional Reservoir Zone
properties is between the transitional reservoir zone, The porosity range for the transitional reservoir
affected by bitumen plugging and matrix dissolu- zone determined from plug data is intermediate be-
tion, and the central platform reservoir (Figure 27). tween the central platform range and that of the rim
Because similar depositional facies are present in both and flank reservoirs (Figure 27A). The data also show
the outer and central regions for shallow-platform cy- a large amount of permeability scatter around the
cles, bitumen and dissolution are mostly responsible central platform trend (Figure 27B). The transitional
for the observed differences. In some cycles, early po- reservoir zone matrix is characterized by larger pore
rosity and permeability in the transitional reservoir sizes because of solution enlargement of early in-
zone may have been adversely affected by the presence terparticle and moldic porosity and because smaller
84 / Collins et al.

FIGURE 26. Characteristics of the transitional reservoir zone, which affects the outer regions of the Visean A, Serpukhovian,
and Bashkirian platforms. Facies include skeletal packstone-rudstone with brachiopods and large crinoids (D, F, J, and K),
sponge or coral boundstone (E and H), and occasional high-energy ooid grainstone (G). These examples have in common
dissolution enlargement and bitumen cementation in the matrix. Dissolution enhancement includes matrix vugs
(D, arrows) and enlarged fractures or stylolites (D, J, and K). Bitumen (darker colors in core examples) is somewhat
fabric selective in (E – H). The example in (F), interpreted as a karst feature, shows differences in relative bitumen
abundance relative to the fracture and to the clasts in the cavity. In thin section, bitumen is commonly associated with
minor corrosion (cement corrosion in B and C, arrows). Bitumen and solution enlargement have offsetting effects
on porosity and permeability, as shown in (A), in which both bitumen-cemented pores (arrows) and solution-enlarged
pores are visible.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 85

pores tend to be bitumen filled (Figure 26A, B). on porosity logs. In particular, dissolution some-
Larger pores exhibit only partial fills or incomplete what disguises the loss of porosity observed at se-
linings. The reduced porosity range in the transi- quence boundaries in the central platform (Kenter
tional reservoir zone, thus, appears to be caused by et al., 2006) and adds high-frequency vertical po-
bitumen cementation of smaller pores, whereas the rosity variations in the cycles. Log character thus
permeability scatter appears to be caused by a com- provides a means of constraining the width of the
bination of bitumen plugging and solution enhance- transitional reservoir zone along the outer edge of
ment that resulted in a fundamental change in matrix the platform. Based on current well and core control,
pore size, shape, and connectivity. For the same po- the width appears to vary vertically and may range
rosity value or porosity range, the transitional reser- from less than 500 m (1640 ft) to more than 2 km
voir zone has slightly better average matrix perme- (1.2 mi).
ability compared to the central platform reservoir.
Thus, whereas bitumen plugging appears to have re- Rim and Flank Reservoir
duced the porosity range in the transitional reservoir The data from the rim and flank reservoir in
zone, an expected reduction in permeability over that Figure 27 indicate a further reduction in matrix po-
range has been offset by dissolution. rosity range but permeability scatter similar to the
Because significant matrix porosity likely remained transitional reservoir zone. Although it was originally
in the transitional reservoir zone after depositional anticipated that matrix reservoir quality might depend
and early diagenetic processes, cyclic controls on on the amount of grainy matrix in different facies,
platform porosity probably influenced patterns of comparisons of matrix-rich breccias, more tightly
later dissolution and cementation. The net effect of packed breccias containing little or no matrix, and
porosity changes caused by bitumen plugging and in-situ microbial boundstones do not reveal greatly
matrix dissolution overprinted across facies results different porosity ranges (Figure 28). It is also sur-
in a noticeable change in the expression of cyclicity prising, given that matrix dissolution features are less

FIGURE 27. Plug porosity and permeability data comparing the matrix reservoir from the central platform, transitional
reservoir zone (TRZ), and rim and flank areas. The central platform trend includes almost 5000 plugs (only a partial set
is plotted) and demonstrates a single trend despite a variety of facies, textures, and reservoir rock types represented. The
rest of the data consists of approximately 1700 plugs from the rim and flank and 700 from Bashkirian and Serpukhovian
outer-platform facies representing the TRZ. Increased permeability scatter and reduced porosity ranges characterize
both the TRZ and the rim and flank trends compared to the central platform (A). In the TRZ, reduced porosity is caused
mainly by bitumen, whereas the permeability scatter is caused by a change in pore connectivity from the combination
of bitumen plugging and solution enlargement. In the rim and flank, the porosity range is reduced further (B), because
of the presence of tightly cemented microbial boundstone facies. The permeability scatter is partly a function of solution-
enlarged pores, but also because of the increased importance of vuggy-moldic pore types and microfracture porosity.
86 / Collins et al.

FIGURE 28. Histograms from plug data comparing matrix porosity ranges in the rim and flank by facies (A) and by
well (B). Boundstones and breccias have similar distributions, rudstones have slightly better porosity (based on higher
frequency in the 3 – 5% bins), whereas fine-grained platform-derived facies have a lower porosity range. Overall, however,
porosity variability between facies is instead weak compared to the variations by well, suggesting a strong diagenetic
overprint.

abundant overall and appear to be more poorly con- upper-slope facies are less noticeable. Burial cemen-
nected, that rim and flank matrix permeability is tation and dissolution processes may be part of the
comparable to that of the transitional reservoir zone. reason why there is no good evidence that breccias,
This may be partly an artifact of the increased abun- in general, have better reservoir quality compared to
dance of small fractures, producing a greater amount other facies, or that better reservoir quality is asso-
of plug damage. However, dissolution is observed ciated with a particular type of breccia.
along some small fractures in plugs, so whereas the
number of plugs with high permeability may be in- Matrix and Nonmatrix Reservoir Distribution
flated, the range is probably real. It should also be Nonmatrix features consisting of fractures, large
remembered that these are comparative statistical vugs, and lost circulation zones are an important
relationships that only apply where matrix porosity component of the rim and flank reservoirs. These
exists and cannot be extrapolated to generalizations elements are absent or greatly reduced by compar-
about the overall distribution of porosity. ison in the central platform area. The distribution of
Porosity in upper- and middle-slope microbial these features and their relationships to matrix prop-
boundstones and breccias was substantially reduced erties are the basis of rim and flank reservoir char-
by sediment fill and prebitumen cementation of acterization. As described previously, open fractures,
early pore types. These include fenestrae, microbial vugs, and lost circulation zones appear to be linked
growth and framework cavities, and other primary to areas of bitumen cementation and solution en-
voids, as well as early fractures and dissolution voids. largement (see Figure 25). Whereas bitumen is asso-
In middle- to lower-slope and apron facies, it was ciated with dissolution along fractures and in the
expected that the presence of skeletal rudstones or surrounding matrix, it is also responsible for plug-
floatbreccias with grainy matrix might improve ging up some fractures and vugs. Furthermore, bitu-
reservoir quality in some areas because of possible men and dissolution affect matrix reservoir quality to
preservation of primary porosity, compared to areas varying degrees in fractured intervals, and matrix dis-
containing tightly packed breccias or breccias with solution and bitumen can sometimes be more exten-
finer grained or micritic matrix. However, like the sively developed in areas of lower fracture density.
upper slope, matrix reservoir quality in most middle- Understanding these variations in spatial and genetic
to lower-slope facies is affected by bitumen, corro- relationships is a key to improved reservoir-quality
sion, and prebitumen cement, in particular syntaxial prediction.
calcite (of probable early to moderate burial origin) Bitumen-associated matrix dissolution includes
and clear equant calcite (including 908 calcite of microporosity, micrite corrosion, vugs, and enlarged
burial origin). Early marine cements prevalent in the primary voids. In detrital middle- to lower-slope and
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 87

apron breccias, residual matrix porosity controlled cavities (Figure 29). Almost all of these features are
by primary texture and prebitumen cementation completely filled with sediment and, to a lesser ex-
appears to have regulated the extent to which matrix tent, cement. Later bitumen and corrosive diagenesis
corrosion is developed locally. Incompletely cemented are present in these intervals and, in some cases, are
grainy facies lend themselves to more extensive cor- even concentrated within or around the karst or dis-
rosion, and small-scale variations in these parameters solution features (Figure 29A). However, when com-
sometimes result in apparent bedding or facies control, pared to randomly selected cored intervals or to cored
obscuring a relationship to larger scale fracture den- intervals of equal thickness where similar features are
sity. In upper-slope boundstones and breccias, mi- absent, the intervals containing the karst features
croporosity occurs mainly in primary microbial fab- tend to have, on average, diminished amounts of bi-
rics and entrained micritic skeletal grains. Because tumen and dissolution, and average reservoir quality
little or no matrix is present in these rocks, a general in these zones does not appear to be consistently
relationship to open and partly open fractures is more better or worse than elsewhere.
clearly evident. The amount of matrix corrosion and
dissolution that develops depends on the type of Lost Circulation Zones
boundstone (which, in turn, determines the density Lost circulation during drilling in the rim and flank
of entrained grains and microbial fabrics) and on the can be shown to occur in association with increased
amount of prebitumen cement (commonly abundant open-fracture density or with solution enlargement
because of high volumes of marine cement). Whereas of fractures in instances where FMI control, or in
this combination can produce a complex-looking dis- some cases, core control, exists. Recent drilling with-
tribution of corrosion and bitumen relative to primary out catastrophic losses and with more complete core
textures in boundstones and breccias, a clear spatial recoveries in intervals dominated by vuggy porosity
connection exists between the matrix and nonmatrix further suggests that perhaps the distribution of frac-
components of bitumen and associated dissolution, tures or solution-enlarged fractures are the most sig-
and variations in matrix components are determined nificant controls on lost circulation in general. Because
more by prebitumen cementation patterns than by solution-enlarged fractures are commonly associated
fracture density or facies. with areas of dissolution, corrosion, and bitumen, an
important question is whether fracture distribution
Early Karst Features was the primary control on where solution enlarge-
Early karst features are commonly cited as impor- ment and lost circulation occurred, or whether the
tant influences on porosity and permeability distri- distribution of corrosion and dissolution, constrained
bution in carbonate reservoirs. The overall effect may by other factors, determined where fracture enlarge-
be positive or negative (Esteban and Wilson, 1993; ment and lost circulation occurred. Ultimately, the
Wagner et al., 1995). Early karst features can vary distribution of lost circulation zones in the rim and
considerably in spatial distribution and relationship flank may indicate specific pathways through the
to stratigraphy (for example, Choquette and James, reservoir taken by corrosive fluids or map out areas of
1987; Lucia, 1995; Mylroie and Carew, 1995), al- elevated fracture density.
though for isolated platforms, dissolution is com- As previously outlined, different fracture sets in
monly concentrated near the platform margin (for the Tengiz rim and flank can be distinguished by
example, Cander, 1995; Tinker et al., 1995). Early orientation relative to core, by types of fill, and by
karsts may also influence later diagenesis, serving as the amount of dissolution. Possible origins are varied,
targets for selective re-excavation or forming signifi- including syndepositional instability, burial stresses
cant impediments to later fluid flow (Ford, 1995). created by differential compaction of the rim and
Large-scale features interpreted as early karst are platform areas (and, to a lesser extent, the rim and
present at Tengiz; however, observed relationships flank areas), differential stresses resulting from over-
between these features, the distribution of later dia- burden loading during salt diapirism, and stresses
genetic features (bitumen cement and corrosion), associated with various tectonic events affecting the
and overall porosity and permeability suggest that Precaspian Basin (D. G. Carpenter, 2005, personal com-
early karsts are largely incidental to reservoir quality. munication). Fracture density controlled by the me-
Potential early karst features include local subaer- chanical properties of different facies is one possible
ial exposure effects associated with sequence bound- influence on the current distribution of lost circulation
aries, brecciated intervals, and large, sediment-filled zones. Both core and FMI data show that microbial
88 / Collins et al.

FIGURE 29. Core examples with sediment-filled features of possible karst origin. (A) Cavities filled with laminated,
greenish-colored shale and carbonate formed in outer-platform crinoid skeletal packstone-rudstone. Cavity walls (yellow
dashed lines) are surrounded by a zone of matrix corrosion and bitumen (white lines), but matrix reservoir quality overall
is marginal. (B) Cavity developed in outer-platform to upper-slope boundstone filled with laminated green-gray marl and
boundstone breccia. Because of the unique setting of the T-4635 well (see Figure 15A), this feature could be either a karst
feature or a cavity that opened during rim detachment and subsequently filled in a deeper water setting. (C – E) Examples
of smaller, geopetally filled fenestral cavities. The sediment fill sequence, which includes green and red clay (C and D),
banded (marine) cement and cream-colored micrite (D), and dark-colored sediment (D and E), suggests a complex history
of cavity formation and fill. All examples are associated with relatively poor matrix reservoir quality, however.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 89

FIGURE 30. Distribution of lost circu-


lation zones (LCZ) in the rim and flank.
Most wells in the rim and flank have
lost circulation somewhere in the
reservoir section. The charts represent
48 LCZ in 13 wells (T-463, T-4346,
T-4635, T-4556, T-5056, T-5059, T-5435,
T-5454, T-5857, T-6261, T-6457,
T-7252, and T-7453) that have been
quantified by sustained loss rate,
depth, and thickness. The distribution
by distance in meters from the top
of the section does not show a clear
trend (left), but when the distribution
is normalized to indicate relative
proportion in the rim interval regard-
less of well position on the rim
profile (right), the pattern supports
anecdotal evidence that LCZ are
more common in the upper part of
the section, regardless of facies pres-
ent there.

grainy facies and loosely packed


breccias) may have allowed corro-
sive fluids to spread away from
fractures and, thus, may ultimately
prove less prone to lost circulation.
This may be a reason why, in the
transitional reservoir zone where per-
meable grainy facies are most abun-
boundstones and tightly packed breccias have a dant, a relationship between lost circulation zones,
greater fracture density compared to rudstones, pack- bitumen, and corrosion is less obvious. Combined
stones, and matrix-supported breccias, suggesting with reduced fracture density in the transitional res-
that lost circulation zones might be more common ervoir zone, this factor may also be why lost circu-
in upper- to middle-slope facies. However, based on lation zones are less frequent in the outer platform
well occurrences, it can be demonstrated that lost region. Whatever diagenetic processes prove ultimately
circulation zones are more common in the upper half responsible for dissolution and lost circulation, their
to one-third and slightly more common near the base effects have produced a reservoir with large-scale in-
of the rim and flank section, regardless of position in terwell connectivity across the transitional reservoir
the field (Figure 30). Such a distribution is inconsis- zone, rim, and flank, despite a major stratal surface
tent with a primary facies control on lost circulation separating them and a complex facies distribution
and, like many other observations, points to a dia- within each.
genetic overprint.
Current well spacing is still insufficient to clearly SUMMARY AND CONCLUSIONS
demonstrate geographical distribution patterns for
lost circulation zones, a situation made even worse Based on currently available data, the Tengiz rim
because most rim and flank wells drilled to date and flank are interpreted to have accumulated in
have encountered lost circulation somewhere in the two stages. The early stage was dominated by insta-
reservoir section. More generally, facies in which bility and relatively continuous large-scale failure and
fracture permeability was dominant compared to mass wasting, resulting in deposition of debris aprons
residual matrix permeability may be more prone to around the base of the buildup consisting of coarse
lost circulation because later dissolution was con- boundstone breccias and grainy platform-derived
centrated along fractures. Facies with higher residual sediments. In the second stage, a wedge consisting
matrix permeability (for example, poorly cemented of shallow- and outer-platform facies, upper-slope
90 / Collins et al.

microbial boundstone, middle-slope boundstone brec- controlled by either fracture distribution or the
cia, and lower-slope breccia and platform-derived distribution of late dissolution fairways.
sediment prograded out over the early debris apron. 6) Enhanced matrix porosity caused by late disso-
This stage was more successful along the eastern lution and bitumen cement is commonly better
margin of the buildup because of favorable platform developed where solution-enlarged fractures are
profile geometry, or elevated microbial productivity, present; however, extensive matrix corrosion
or reduced tectonic influence in those areas, or some may be associated with intervals that are less
combination of those factors. Along the western mar- likely to produce severe lost circulation.
gin of the buildup, progradation was only locally or 7) Reservoir-quality variability exists at large and
partially successful, and rim failure predominated small scales, resulting in considerable well-to-well
throughout the Serpukhovian and Bashkirian. differences in regions that are similar in terms of
This study does not provide a process-based, fully facies, stratigraphy, and depositional history. This
predictive model of the distribution of reservoir prop- suggests that facies is a weak secondary control on
erties in the Tengiz outer platform, rim, and flank. It reservoir quality.
raises as many questions as it answers, and some
critical uncertainties remain. Some conclusions have
alternatives and will likely require modification or ACKNOWLEDGMENTS
change as drilling continues. Future surprises are
probable, and subsequent generations of models will This study involved the work of a large group of
no doubt be different from the present one in some geoscientists and specialists. The authors recognize
manner. However, the following observations indi- the following individuals for insightful contributions
cate factors that are likely to have a significant impact to this manuscript: Kevin Nahm (Chevron), who
on reservoir quality and prediction in the rim and believed early on that mechanically isotropic micro-
flank: bial boundstones could be mapped with seismic
analysis; Brent Francis (ExxonMobil), who progressed
1) Reservoir quality is dominantly controlled by frac- this idea through development of the compaction
tures and large-scale dissolution and is a function model and who provided seismic interpretations for
of earlier porosity destruction because of prebitu- the debris apron concept; and Tom Heidrick (retired
men cementation versus later dissolution associ- from Chevron), who engaged in valuable discussions
ated with bitumen formation. regarding the minimal effects of early karst and the
2) Porosity destruction caused by sediment infill potential for later burial fractures. Wayne Narr
and prebitumen cementation was significant and (Chevron), Jim DeGraff (ExxonMobil), and Dan
affected primary voids, early dissolution voids, Carpenter (ExxonMobil) are acknowledged for build-
early fractures, and early karst features. In par- ing on this perception through rigorous description
ticular, early fractures and large-scale dissolution and analyses of fracture characteristics and distribu-
features or karsts are largely filled and do not tion. Perhaps most importantly, Michael Clark (Chev-
have a major impact on reservoir quality. ron) is thanked for initiating an intense coring and
3) Corrosive dissolution and bitumen are a late- core analysis program that formed the basis for most
diagenetic overprint and are responsible for sig- of our geologic studies and for his strong support of the
nificant porosity modification in a wide variety sequence-stratigraphic framework.
of features including later fractures, microbial We also gratefully thank the following individuals
fabrics, micritic skeletal grains, breccia clasts, cer- from ExxonMobil who made significant contributions
tain cements, and stylolites. of supporting data or data analysis: Gareth Jones and
4) Facies control of fracture density results in both Yitian Xiao (numerical simulation of Tengiz diagenet-
open and closed types being more common in ic processes); Sean Guidry (fluid-inclusion analyses);
mechanically rigid facies, such as microbial bound- Paul Hicks (temperature histories from basin model-
stone, cemented grainstone, and clast-supported ing); Shin-Ju Ye and Peter Hillock (facies analysis of
breccia, and less abundant in facies such as rud- FMI logs). Early versions of the manuscript were
stone, packstone, and matrix-supported breccia. improved by helpful reviews from Niall Toomey,
5) Lost circulation zones appear to be more com- Kelly Bergman, and Gareth Jones. Finally, we thank
monly associated with solution-enlarged fractures TengizChevroil and its shareholder companies for
than with vuggy intervals and are therefore likely permission to publish this study.
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 91

APPENDIX 1: RIM AND FLANK FACIES (MODIFIED FROM WEBER ET AL., 2003)

Lithofacies and Description Environment


Microfacies
Crinoid-brachiopod Medium-dark gray; large crinoids and/or Deep subtidal outer platform, possibly
skeletal rudstone fragmented thick-shelled brachiopods below wave base; abundant in central
to packstone dominate; sometimes with microbial to upper parts of late Visean and
fabrics and cements Serpukhovian outer-platform cycles
Outer platform breccias Grainstone floatbreccia to packbreccia: Deep outer platform, possibly lowstand;
and conglomerates light-medium brown; subangular to basal parts of Serpukhovian outer-platform
subrounded clasts of platform-derived cycles
grainstone and packstone; skeletal
wackestone to grainstone matrix,
sometimes with microbial fabrics and
cements
Grainstone packbreccia: light brown, gray, Shallow outer platform; lag or possibly
and black; angular clasts of grainstone and karst? Tops of Serpukhovian outer-platform
rudstone with mixed lithofacies types cycles
Coral rudstone to conglomerate: medium Shallow outer-platform; basal parts of
brown to gray; colonial coral fragments; Serpukhovian outer-platform cycles;
coarse-grained skeletal matrix with possible patch-reef detritus
abundant crinoids
Brachiopod rudstone Light gray-brown to medium brown-gray; Shallow outer-platform; basal parts of
(coquina) to boundstone brachiopod shell fragments and whole Serpukhovian outer-platform cycles;
shells; some articulated, in feeding or mud-mounds, mound detritus, and
growth position; abundant micritic matrix, intermound deposits
some microbial fabrics and calcite cement;
minor coarse skeletal debris, crinoids
Microbial boundstone Type C (skeletal-micritic boundstone): Serpukhovian deep, subtidal, outer-platform
light-medium brown, massive to peloidal to upper-slope platform break
with abundant fenestrae; crinoids,
brachiopods, fenestellid bryozoans, algae,
forams, and other skeletal grains bound
together by clotted microbial fabric and
multiple cement generations (as much as
50% cement)
Type B (laminar-cemented boundstone): Serpukhovian upper slope
light brown; irregular concentric or wavy-
aminated fabric consisting of microbial
filaments and early cement (botryoidal,
fibrous and radiaxial; inclusion rich; as
much as 75%); minor skeletal grains,
including fenestellid bryozoa, crinoids,
ostracods, thin-walled pelecypods and
forams
Type A (peloidal-micritic boundstone): Serpukhovian upper to middle slope
light gray-light brown; generally massive
or featureless, with minor growth structures
and abundant irregular fenestrae or tubules;
clotted peloidal fabric and early cement
(as much as 25%); sparse fauna, including
fenestellid bryozoans, ostracods, and thin-
walled pelecypods
92 / Collins et al.

APPENDIX 1: (CONT.)

Lithofacies and Description Environment


Microfacies
Boundstone breccia Welded to mosaic breccia: light-medium Serpukhovian upper to middle slope and
brown and gray; subangular centimeter- to debris apron
meter-size clasts of upper-slope boundstone
types; tight, clast-supported fabric, sometimes
with stylolitic clast boundaries (welded
breccia), sometimes with blocky calcite
cement (mosaic breccia); minor platform-
derived matrix (skeletal grains, large crinoids);
occasional platform-derived grainstone or
packstone lithoclasts
Float- to packbreccia: light-dark brown and Serpukhovian middle to lower slope and
gray; subangular to subrounded, millimeter- debris apron
to meter-size clasts of upper slope-derived
boundstone types; grainstone-packstone
matrix with platform-derived skeletal grains;
large crinoids and fragmented brachiopod
shells; occasional platform-derived grainstone
or packstone lithoclasts
Crinoid-lithoclast Brown-light gray; poorly to moderately Serpukhovian middle to lower slope and
rudstone-packstone sorted, crinoid-rich rudstone (platform- debris apron
derived); fragmented brachiopod shells;
sand-size skeletal grains and millimeter- to
centimeter-size lithoclasts (upper slope and
platform derived)
Skeletal-lithoclastic Medium-dark gray; thin to thick bedded, Serpukhovian lower slope and debris apron
rudstone-grainstone- wavy-bedded to laminated, graded; locally
floatstone siliceous to slightly dolomitic; platform-
derived grains (forams, crinoids, algae,
brachiopods, crinoids); occasional
millimeter- to centimeter-size boundstone
or slope-derived clasts, volcanic siltstone
or sandstone clasts
Allochthonous grainstone Light brown-medium gray; moderately Lower slope and debris apron; Serpukhovian
sorted ooid or skeletal grainstone; massive and Bashkirian
to well bedded
Lime wackestone- Brown, gray, and black; argillaceous, Lower slope, debris apron and basinal;
mudstone siliceous, and commonly dolomitic; thin Visean, Serpukhovian, and Bashkirian
laminated to thin bedded; sponge spicules,
minor skeletal grains, and sparse
radiolarians; commonly interbedded with
volcanic silt and sand; potential source rock
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 93

APPENDIX 2: FLUID-INCLUSION RESULTS

Well Depth (m) Platform Sequence Phase Type # incl. Thomog (Water) (8C)
T-4635 4466.25 Serpukhovian rim Equant CC Primary 2 85 – 90
T-4635 4466.25 Serpukhovian rim Equant CC Primary 2 90 – 95
T-4635 4482.63 Serpukhovian rim Equant CC Primary 1 88
T-4635 4482.63 Serpukhovian rim Equant CC Primary 3 80 – 85
T-4635 4482.63 Serpukhovian rim Equant CC Primary 2 85 – 90
T-4635 5055.25 Famennian Calcite Unspecified 5 105 – 110
T-4635 5055.25 Famennian Calcite Unspecified 6 105 – 115
T-4635 5055.25 Famennian Calcite Unspecified >10 110 – 120
T-4635 5055.25 Famennian Calcite Unspecified 4 90 – 95
T-47 5392.5 Famennian Quartz vein Unspecified No assemblage 115 – 125
T-5056 4194.2 Serpukhovian rim Equant CC Primary 1 80 – 85
T-5056 4194.2 Serpukhovian rim Equant CC Primary 5 85 – 90
T-5056 4200.47 Serpukhovian rim Fluorite Primary 4 110 – 125
T-5056 4200.47 Serpukhovian rim Fluorite Primary 4 110 – 125
T-5056 5453.55 Famennian Zoned CC Primary 2 100
T-5056 5453.55 Famennian Zoned CC Primary 3 95 – 100
T-5056 5453.55 Famennian Equant CC Unspecified 5 85 – 95
T-5056 5476.22 Famennian Calcite Primary 4 95 – 105
T-5857 4816.93 Serpukhovian rim Prismatic CC Primary 2 80 – 85
T-5857 4816.93 Serpukhovian rim Prismatic CC Primary 3 85 – 90
T-6337 4894.37 Visean D Calcite Primary 2 85 – 90
T-6337 4945.64 Visean D Equant CC Primary 4 85 – 90
T-6337 4945.64 Visean D Equant CC Primary 4 85 – 90
T-6337 4945.64 Visean D Equant CC Primary 3 85 – 90
T-8 3972 Bashkirian Coarse CC Unspecified 4 100 – 107
T-8 4044.5 Serpukhovian Zoned CC Primary 4 80 – 95
T-8 4044.5 Serpukhovian Equant CC Primary 2 85 – 95
T-8 4056.8 Serpukhovian Equant CC Primary 2 85 – 90

Data for hydrocarbon-water or vapor-filled inclusions:


Phase refers to cement types described in the text (see Figure 24). ‘‘Calcite’’ indicates the cement phase preceding bitumen formation and, in most
cases, refers to equant calcite. Prismatic calcite represents the cores of zoned cement crystals (see Figure 20), and zoned calcite represents subsequent
inclusion-rich growth zones (either calcite or dolomite). Coarse calcite is a postbitumen cement phase. An inclusion type of ‘‘unspecified’’ indicates an
assemblage or train whose orientation was not specified during analysis, but was not designated as cutting across growth planes and may be
primary. The column ‘‘# incl.’’ indicates the number of individual inclusions in a train or assemblage.

REFERENCES CITED platform, Kazakhstan: Revista Italiana di Paleontologia e


Stratigrafia, v. 109, no. 2, p. 131– 158.
Bahamonde, J. R., C. Vera, and J. R. Colmenero, 2000, A Cander, H., 1995, Interplay of water-rock interaction ef-
steep-fronted Carboniferous carbonate platform: ficiency, unconformities, and fluid flow in a carbon-
Clinoformal geometry and lithofacies (Picos de Euro- ate aquifer: Floridan aquifer system, in D. A. Budd,
pa, NW Spain): Sedimentology, v. 47, p. 645 – 664. A. H. Saller, and P. M. Harris, eds., Unconformities
Blendinger, W., 2001, Triassic carbonate buildup flanks in and porosity in carbonate strata: AAPG Memoir 63,
the Dolomites, northern Italy: Breccias, boulder fabric p. 103 – 124.
and the importance of early diagenesis: Sedimentol- Choquette, P. W., and N. P. James, 1987, Introduction, in
ogy, v. 48, no. 5, p. 919 – 933. N. P. James and P. W. Choquette, eds., Paleokarst:
Brenckle, P. L., and N. V. Milkina, 2003, Foraminiferal New York, Springer-Verlag, p. 1 – 21.
timing of carbonate deposition on the Late Devonian Della Porta, G., J. A. M. Kenter, J. R. Bahamonde, A.
(Famennian)–middle Pennsylvanian (Bashkirian) Tengiz Immenhauser, and E. Villa, 2003, Microbial boundstone
94 / Collins et al.

dominated carbonate slope (Upper Carboniferous, N Guadaupe Mountains, Texas and New Mexico, U.S.A.:
Spain): Microfacies, lithofacies distribution and stratal Journal of Sedimentary Research, v. 68, p. 956 – 969.
geometry: Facies, v. 49, p. 175–208. Kohout, F. A., H. R. Henry, and J. E. Banks, 1977,
Esteban, M., and J. L. Wilson, 1993, Introduction to karst Hydrology related to geothermal conditions of the
systems and paleokarst reservoirs, in R. D. Fritz, J. L. Floridan Plateau, in K. L. Smith and G. M. Griffin,
Wilson, and D. A. Yurewicz, eds., Paleokarst related eds., The geothermal nature of the Floridan Plateau:
hydrocarbon reservoirs: SEPM Core Worshop 18, p. 1 – Florida Department of Natural Resources Bureau
9. Geology Special Publication 21, p. 1 – 34.
Ford, D. C., 1995, Paleokarsts as a target for modern Lisovsky, N. N., G. N. Gogonenkov, and Y. A. Petzoukha,
karstification: Carbonates and Evaporites, v. 10, no. 2, 1992, The Tengiz oil field in the Pre-Caspian Basin of
p. 138 – 147. Kazakhstan (former USSR) — Supergiant of the 1980s,
Heydari, E., 2000, Porosity loss, fluid flow and mass in M. T. Halbouty, ed., Giant oil and gas fields of the
transfer in limestone reservoirs: Application to the decade 1978 – 1988: AAPG Memoir 54, p. 101 – 122.
Upper Jurassic Smackover Formation, Mississippi: Lucia, F. J., 1995, Lower Paleozoic cavern development,
AAPG Bulletin, v. 84, no. 1, p. 100 – 118. collapse, and dolomitization, Franklin Mountains, El
Hill, C. A., 1987, Geology of Carlsbad Cavern and other Paso, Texas, in D. A. Budd, A. H. Saller, and P. M.
caves in the Guadalupe Mountains: New Mexico Harris, eds., Unconformities and porosity in carbon-
Bureau of Mines and Mineral Resources, New Mexico ate strata: AAPG Memoir 63, p. 279 – 300.
and Texas Bulletin 117, 150 p. Lucia, F. J., 1999, Carbonate reservoir characterization:
Hill, C. A., 1990, Sulfuric acid speleogenesis of Carlsbad New York, Springer-Verlag, 222 p.
Cavern and its relationship to hydrocarbons, Dela- Machel, H. G., 1987, Some aspects of diagenetic sulphate-
ware basin, New Mexico and Texas: AAPG Bulletin, hydrocarbon redox reactions, in J. D. Marshall, ed.,
v. 74, no. 11, p. 1685 – 1694. Diagenesis of sedimentary sequences: Geological
Hill, C. A., 1995, H2S-related porosity and sulfuric acid oil- Society (London) Special Publication 36, p. 15 – 38.
field karst, in D. A. Budd, A. H. Saller, and P. M. Harris, Machel, H. G., 1989, Relationships between sulphate
eds., Unconformities and porosity in carbonate strata: reduction and oxidation of organic compounds to
AAPG Memoir 63, p. 301 – 313. carbonate diagenesis, hydrocarbon accumulations,
Jones, G. D., F. F. Whitaker, P. L. Smart, and W. E. Sanford, salt domes, and metal sulphide deposits: Carbonates
2004, Numerical analysis of seawater circulation in and Evaporites, v. 4, no. 2, p. 137 – 151.
carbonate platforms: II. The dynamic interaction Machel, H. G., and J. H. Anderson, 1989, Pervasive
between geothermal and brine reflux circulation: subsurface dolomitization of the Nisku Formation in
American Journal of Science, v. 304, p. 250 – 284. central Alberta: Journal of Sedimentary Petrology,
Kenter, J. A. M., P. M. Harris, and G. Della Porta, 2005, v. 59, no. 6, p. 891 – 911.
Steep microbial boundstone-dominated platform Melim, L. A., and P. A. Scholle, 1995, The forereef facies of
margins — Examples and implications: Sedimentary the Permian Capitan Formation: The role of supply
Geology, v. 178, p. 5 – 30. versus sea-level changes: Journal of Sedimentary Re-
Kenter, J. A. M., P. M. Harris, J. F. Collins, L. J. Weber, G. search, v. B65, no. 1, p. 107 – 118.
Kuanysheva, and D. J. Fischer, 2006, Late Visean to Bash- Morrow, D. W., 1998, Regional subsurface dolomitization:
kirian platform cyclicity in the central Tengiz buildup, Models and constraints: Geoscience Canada, v. 25,
Precaspian Basin, Kazakhstan: Depositional evolution p. 57 – 70.
and reservoir development, in P. M. Harris and L. J. Mylroie, J. E., and J. L. Carew, 1995, Karst development on
Weber, eds., Giant hydrocarbon reservoirs of the carbonate islands, in D. A. Budd, A. H. Saller, and P. M.
world: From rocks to reservoir chacracterization and Harris, eds., Unconformities and porosity in carbon-
modeling: AAPG Memoir 88/SEPM Special Publica- ate strata: AAPG Memoir 63, p. 55 – 76.
tion, p. 7 – 54. Playford, P. E., 1984, Platform-margin and marginal-slope
Kerans, C., and P. M. Harris, 1993, Outer shelf and shelf relationships in Devonian reef complexes of the
crest, in D. G. Bebout and C. Kerans, eds., Guide to the Canning basin, in P. G. Purcell, ed., The Canning
Permian reef geology trail, McKittrick Canyon, Guadalupe basin, Western Australia: Geological Society of Aus-
Mountains National Park, west Texas, Guidebook 26: tralia and Petroleum Exploration Society of Australia,
Austin, Bureau of Economic Geology, p. 32 – 43. Canning basin Symposium, Perth 1984, p. 189 – 234.
Kerans, C., N. F. Hurley, and P. E. Playford, 1986, Marine Ross, C. A., and J. R. P. Ross, 1987, Late Paleozoic sea levels
diagenesis in Devonian reef complexes of the Can- and depositional sequences: Cushman Foundation for
ning basin, Western Australia, in J. H. Schroeder and Foraminiferal Research Special Publication 24, p. 137 –
B. H. Purser, eds., Reef diagenesis: Berlin, Springer- 149.
Verlag, p. 357 – 380. Ross, C. A., and J. R. P. Ross, 1988, Late Paleozoic
Kirkland, B. L., J. A. D. Dickson, R. A. Wood, and L. S. Land, transgressive-regressive deposition, in C. K. Wilgus,
1998, Microbialite and microstratigraphy: Encrusta- B. S. Hastings, C. Kendall, H. W. Posamentier, C. A.
tions in the middle and upper Capitan Formation, Ross, and J. C. Van Wagoner, eds., Sea-level changes:
The Outer Platform, Rim, and Flank of the Tengiz Buildup / 95

An integrated approach: SEPM Special Publication 42, karst events related to stratigraphic cyclicity: San Andres
p. 227 – 247. Formation, Yates field, West Texas, in D. A. Budd, A. H.
Saller, A. H., 1984, Petrologic and geochemical constraints Saller, and P. M. Harris, eds., Unconformities and poros-
on the origin of subsurface dolomite, Enewetak Atoll: ity in carbonate strata: AAPG Memoir 63, p. 213– 237.
An example of dolomitization by normal sea water: Wagner, P. D., D. R. Tasker, and G. P. Wahlman, 1995,
Geology, v. 12, p. 217 – 220. Reservoir degradation and compartmentalization be-
Sanford, W. E., F. F. Whitaker, P. L. Smart, and G. D. Jones, low subaerial unconformities: Limestone examples
1998, Numercial analysis of seawater circulation in from west Texas, China, and Oman, in D. A. Budd,
carbonate platforms: I. Geothermal circulation: Amer- A. H. Saller, and P. M. Harris, eds., Unconformities
ican Journal of Science, v. 298, p. 801 – 828. and porosity in carbonate strata: AAPG Memoir 63,
Stephens, N. P., and D. Y. Sumner, 2003, Famennian mi- p. 301 – 313.
crobial reef facies, Napier and Oscar Ranges, Canning Weber, L. J., B. P. Francis, P. M. Harris, and M. Clark, 2003,
basin, Western Australia: Sedimentology, v. 50, p. 1283– Stratigraphy, lithofacies, and reservoir distribution,
1302. Tengiz field, Kazakhstan, in W. M. Ahr, P. M. Harris,
Stoessell, R. K., 1992, Effects of sulfate reduction on W. A. Morgan, and I. D. Somerville, eds., Permo-
CaCO3 dissolution and precipitation in mixing-zone Carboniferous carbonate platforms and reefs: SEPM
fluids: Journal of Sedimentary Petrology, v. 62, no. 5, Special Publication 78 and AAPG Memoir 83, p. 351 –
p. 873 – 880. 394.
Tinker, S., 1998, Shelf-to-basin facies distributions and Wendte, J., H. Qing, J. Dravis, S. L. O. Moore, L. D. Stasiuk,
sequence stratigraphy of a steep-rimmed carbonate and G. Ward, 1998, High temperature saline (thermo-
margin: Capitan depositional system, McKittrick flux) dolomitization of Swan Hills platform and bank
Canyon, New Mexico and Texas: Journal of Sedimen- carbonates, Wild River area, west-central Alberta: Bul-
tary Research, v. 68, p. 1146 – 1174. letin of Canadian Petroleum Geology, v. 46, p. 210 –
Tinker, S. W., J. R. Ehrets, and M. D. Brondos, 1995, Multiple 266.
3
Lindsay, R. F., D. L. Cantrell, G. W. Hughes, T. H. Keith, H. W. Mueller III,
and S. D. Russell, 2006, Ghawar Arab-D reservoir: Widespread porosity
in shoaling-upward carbonate cycles, Saudi Arabia, in P. M. Harris
and L. J. Weber, eds., Giant hydrocarbon reservoirs of the world: From
rocks to reservoir characterization and modeling: AAPG Memoir 88/
SEPM Special Publication, p. 97 – 137.

Ghawar Arab-D Reservoir:


Widespread Porosity in
Shoaling-upward Carbonate
Cycles, Saudi Arabia
Robert F. Lindsay, Dave L. Cantrell, Geraint W. Hughes, Thomas H. Keith,
Harry W. Mueller III, and S. Duffy Russell
Saudi Aramco, Dhahran, Saudi Arabia

ABSTRACT

G
hawar field is the world’s largest, most prolific field, producing 30–318 API oil
from the Arab-D carbonate reservoir. The field is more than 250 km (155 mi)
long, as much as 30 km (18.5 mi) wide, and has more than 300 m (1000 ft) of
structural closure. The Arab-D reservoir, limestone with some dolostone horizons,
stratigraphically comprises the D member of the Arab Formation and the upper part of
the Jubaila Formation. Based on ammonite and benthonic foraminiferal evidence, the
reservoir formations are Upper Jurassic, Kimmeridgian, in age. The reservoir has an
average thickness of more than 60 m (200 ft), an average porosity of more than 15%,
and a permeability up to several darcys. The upper half of the reservoir is dominated by
exceptionally high reservoir quality; the lower half contains interbeds of high and
relatively lower reservoir quality. Early correlation of well logs and cores, before the
advent of sequence stratigraphy, subdivided the reservoir from top to base into zones 1,
2, 3, and 4. Zones 2 and 3 have been subsequently subdivided into zones 2a and 2b and
zones 3a and 3b, with a more detailed zonation scheme for reservoir management.
The reservoir is composed of two composite sequences. One composite se-
quence is the Arab-D Member of the Arab Formation, with the upper boundary at
the top of Arab-D carbonate and below the C-D evaporite, with the sequence
boundary locally marked by pods of collapse breccia. The second composite
sequence forms the upper part of the Jubaila Formation, for which the sequence
boundary between the Arab-D Member and Jubaila Formation is located in zone
2b and is marked by a flood of slightly deeper water cycles over the more grain-
dominated cycles in upper zone 3a and lower zone 2b. Several high-frequency
sequences (HFSs) have been identified, each comprising several cycle sets (parase-
quence sets). Each cycle set is composed of approximately five individual car-
bonate cycles (parasequences), and each cycle is composed of one to three beds.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215875M88576

97
98 / Lindsay et al.

These carbonates were deposited approximately 58 south of the equator on a


broad, arid, storm-dominated carbonate ramp. From upslope to downslope, the
ramp consisted of the following subenvironments: (1) inner ramp; (2) ramp-crest
shoal; (3) proximal middle ramp; (4) distal middle ramp; and (5) outer ramp. The
inner ramp was a lagoon with localized intertidal islands composed of grainstones
and packstones and a highly diverse, shallow-marine benthonic foraminiferal mi-
crofauna. The distal or seaward part of this regime consists of packstones character-
ized by dasyclad and encrusting algae. The ramp-crest shoal is composed of skeletal
and oolitic grainstone, mud-lean packstone, and some mud-rich packstone. Skele-
tal sands of micritized foraminiferal tests and broken skeletal detritus also include
larger fragments of transgressive and storm-derived stromatoporoids and corals.
The proximal middle ramp is composed of domed and encrusting stromatoporoid-
coral mounds and intermound sheltered areas dominated by branched stromatopo-
roids. The distal middle ramp, deposited below fair-weather-wave base, is composed
of micritic to very fine-grained sediment capped by Thalassinoides firmgrounds.
These firmgrounds are overlain by storm-derived rudstone and floatstone of inner-
ramp, ramp-crest, and proximal middle-ramp bioclasts. The outer ramp is com-
posed of deeper shallow-marine deposits of micritic to very fine-grained sediment
capped by Thalassinoides firmgrounds. In this setting, smaller, benthonic forami-
nifera are common along with tetraxon and triaxon sponge spicules and coccoliths.
From highest to lowest reservoir quality, the lithofacies or rock types con-
sist of (1) skeletal-oolitic grainstone, mud-lean packstone, and some mud-rich
packstone; (2) stromatoporoid-red and green algae-coral rudstone and float-
stone; (3) Cladocoropsis rudstone and floatstone; (4) dolomite, porous and locally
extremely permeable to nonporous; (5) bivalve-coated grain-intraclast rudstone
and floatstone; and (6) micritic to very fine-grained deposits.
Limestone porosity is a mixture of the following common pore types: interpar-
ticle (dominant), moldic (common), intraparticle (common), and microporosity.
Less common is porosity associated with Thalassinoides burrows, with vertically
oriented tunnels filled by grain-rich sediment. Shelter porosity is uncommon. Do-
lostone porosity, less common than the major limestone pore types, is a mixture of
moldic, intercrystalline, and (least common) intracrystalline porosity. Fractures
(least common) do not contribute much porosity but contribute permeability.
Diagenetic effects common within Arab-D reservoir carbonates include sev-
eral dissolution events, recrystallization, and physical compaction. Cementation,
episodes of dolomitization, and chemical compaction-stylolitization, although
locally important, were less abundant events.
The vertical seal for the reservoir is the overlying Arab C-D anhydrite. It is
more than 30 m (100 ft) thick and is composed of varvelike laminae to very thin
beds of anhydrite (thicker) and carbonate or organic matter (thinner) deposited in
a salina. The salina periodically shallowed upward into peritidal and intertidal
settings. A few porous-permeable carbonate stringers were deposited when rel-
ative sea level rise flooded the evaporitic shelf and temporarily restarted the sub-
tidal carbonate factory, whereas relative sea level fall reestablished the subtidal
brine factory and precipitated more evaporites.

INTRODUCTION derived from thermally mature, organic-rich carbonate


source rock of the Jurassic-age Hanifa and/or Tuwaiq
The Arab-D reservoir of Saudi Arabia is the most pro- Mountain Formations (Ayres et al., 1982; Droste, 1990)
lific oil-producing interval in the world (Bates, 1973; that subsequently migrated into highly porous and
Beydoun, 1988; 1998; 1991; Durham, 2005). Oils were permeable carbonate reservoir rocks, Arab-D Member
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 99

FIGURE 1. Generalized Late


Jurassic stratigraphy and lithol-
ogies in Saudi Arabia, with Arab
and Hith reservoirs named.
Modified from Powers (1968)
and Meyer et al. (1996).

Rus evaporite beds. To ad-


dress this problem, a shallow
drilling program (generally
less than 300 m [1000 ft])
collected stratigraphic and
structural information be-
neath the pre-Neogene un-
conformity, drilling to the
brown crystalline limestone
of Eocene age beneath the
Rus evaporite. Shallow dril-
ling identified true subsur-
face structures from false
structures and outlined an
anticlinal trend that became
of Arab Formation and upper Jubaila Formation, in known as Abqaiq, southwest of Dammam dome.
large structural traps (Figure 1). Overlying highly effi- Saudi Aramco Abqaiq 1 was drilled in 1940 and dis-
cient evaporite seals, the Arab C-D anhydrite and, ulti- covered oil in porous Arab Formation carbonates.
mately, the Hith anhydrite, prevented further vertical At the same time, field mapping near the Rub’Al
migration and ensured the containment of oil in these Khali desert Empty Quarter, 320 km (200 mi) southwest
large structural traps. of Dammam dome, identified Wadi Sahaba, a west-
Early correlation of well logs, before the advent of to east-oriented wadi that abruptly made a turn to
sequence stratigraphy, subdivided the Arab-D reservoir the south before continuing on to the east. Dips taken
in Ghawar field from top to base into zones 1, 2, 3, and throughout the area revealed a large, broad, low-relief
4. Zones 2 and 3 have been subsequently subdivided dome that was named the ‘‘Haradh feature.’’ This was
into zones 2a and 2b and zones 3a and 3b, respectively, the first indication of the southern tip of the large En
with a more detailed zonation scheme for reservoir Nala anticline, the structural feature on which the shal-
management. Arab-D Member of the Arab Formation lowest crest of the Ghawar field was ultimately found.
consists of zone 1, zone 2a, and upper zone 2b, and As the shallow drilling program continued, it re-
the remainder of the Arab-D reservoir, consisting vealed a structural closure at Ain Dar, west of Abqaiq
of lower zone 2b, zones 3a and 3b, and zone 4, is part and, ultimately, through 1941, connected that clo-
of the upper Jubaila Formation (Meyer et al., 2000). sure with a series of closures all the way to the Haradh
closure along the En Nala anticline. In mid-1948,
DISCOVERY OF OIL IN ARABIA Saudi Aramco Ain Dar 1 established production from
the D member of the Arab Formation and the Jubaila
The first wildcat well drilled in Saudi Arabia was Formation and raised the possibility that all of the
on Dammam dome, a surface structure with four-way structurally high areas along the En Nala anticline
closure visible from the island of Bahrain, some 40 km could be productive. This possibility was confirmed
(25 mi) away. Saudi Aramco Dammam 4 discovered oil in late 1948 by the successful completion of Saudi
in Arab Formation carbonates on March 4, 1938. Aramco Haradh 1, some 200 km (120 mi) farther south.
Other surface structures were identified as addi- Between Ain Dar and Haradh, the next closed structure
tional wildcat drilling sites in eastern Saudi Arabia. to be drilled was in an area called ‘‘Ghawar,’’ referring to
However, some surface structures were drilled (e.g., an area below an escarpment that dominated the area.
Ma’agala and El Alat), only to find they were false struc- The wildcat, named ‘‘Ithmaniya’’ well 1, now called
tures created by differential solution of lower Eocene ‘‘Uthmaniyah well 1,’’ after a small djebel east of the
100 / Lindsay et al.

FIGURE 2. Generalized geo-


logic map of the Arabian Pen-
insula and the position of
the central Arabian arch. Mod-
ified from Al-Hinai et al.
(1997) and U.S. Geological
Survey (1963).

To the west is the Arabian


shield, a vast complex of Pre-
cambrian igneous and meta-
morphic rocks with some youn-
ger igneous rocks (Pollastro
et al., 1999). Bordering the
shield to the east are interior
escarpments where long ar-
cuate belts of Paleozoic, Me-
sozoic, and lower Tertiary rocks
crop out and dip basinward
(eastward) at about 18 or less
(Powers, 1968). In eastern
Arabia, almost flat-lying Ter-
tiary and younger sediments
of the interior or Arabian plat-
staked location, was yet another discovery in the Arab-D form effectively cover and mask these older sedi-
and upper Jubaila. Another structural closure east of Ain ments (Powers et al., 1966).
Dar, now called Shedgum, produced an additional dis- The north-northeast-trending En Nala anticline,
covery in the Arab-D and upper Jubaila in 1952. South of the structural trap for Abqaiq and Ghawar fields, is
Uthmaniyah, the Saudi Aramco Huiyah 1 wildcat, later dominated by mostly north – south to north-north-
changed to Hawiyah, was drilled and became yet an- east – south-southwest-trending anticlinal trends
other discovery in the Arab-D and upper Jubaila. and flexures that reflect deep-seated basement faults
Between 1954 and 1957, successful wildcat and (Figure 3) (Ayres, et al., 1982; Al-Husseini, 2000).
production wells left little doubt that all of the dis- Drilling and seismic evidence (Wender et al., 1998;
covery areas, Ain Dar, Haradh, Uthmaniyah, Shed- Konert et al., 2001) supports the presence, at depth,
gum, and Hawiyah, were part of one large oil field. To of large faults bounding basement blocks under
simplify logistics, each area name was used for well the Ghawar structure that extend from Precambrian
names, such as, Ain Dar well 1, Haradh well 1, etc., metasedimentary basement up through Paleozoic sedi-
whereas the entire field became known as Ghawar. mentary rocks and die out in the Mesozoic section
Basic tools used to discover Ghawar were fairly sim- (Konert et al., 2001). Broad flexing of post-Paleozoic
ple and involved the Brunton compass, alidade, and sedimentary rock is the dominant structural style in
plane table for surface mapping; barometer to determine eastern Arabia (Powers, 1968; Konert et al., 2001)
altitude; a shallow drilling program; and good geologic and is exemplified by the En Nala anticline (Powers
insight. Both the historical outline and the techniques et al., 1966), the structural trap for Ghawar and Abqaiq
used are expanded by Stegner (1971), Barger (2000), Keith fields. Although there is some evidence of subtle growth
(2005), and T. A. Pledge (2006, personal communication). of the En Nala structure at least as early as the Perm-
ian, growth was most pronounced in the Late Creta-
STRUCTURAL SETTING ceous, with regional compressive-transpressive stresses
being imposed on eastern Arabia in response to the
The main oil-producing area of Saudi Arabia, in- closing of the Neo-Tethys Ocean (Beydoun, 1991; Nich-
cluding Abqaiq and Ghawar fields, is located on the olson, 2000, 2002). The most recent pulses of structural
northeastern part of the central Arabian arch (Figure 2). growth occurred during the late Eocene–Oligocene
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 101

FIGURE 3. Major tectonic fea-


tures of the Arabian plate and
Iran. Ghawar field is the large
green oil field in the center that
trends north-northeast. Modi-
fied from Konert et al. (2001).

sion-transtension associated
with the continuing vertical
growth of these anticlines. In
core, oblique to vertically ori-
ented stylolites are evidence of
postlithification compression-
transpression associated with
the vertical fractures.

STRATIGRAPHIC
NOMENCLATURE
AND AGE

The uppermost Jubaila For-


mation, or Ju2 Member (Enay,
1987) as defined in outcrop
where the Jurassic section was
first described and named,
is correlative with the middle
and lower part of the Arab-D
reservoir in the subsurface
(Powers et al., 1966; Powers,
1968) (Figures 1, 4). In out-
crop, the top of the Jubaila
Formation is placed immedi-
ately above the highest oc-
curring stromatoporoids, here
interpreted to mean encrust-
ing and domed stromatopor-
oids, almost invariably not
in original growth position,
and at the base of the over-
lying rippled calcarenite (D.
and during the late Miocene–Pliocene, as the Arabian Vaslet, 2002, personal communication to G.W.
plate welded onto the Eurasian plate along the Zagros Hughes) (Figure 4). The upper part of the Arab-D
suture or crush zone (Beydoun, 1991; Glennie, 2000) reservoir is composed of the D member of the Arab
(Figure 3). Minor faulting, fracturing, and jointing are Formation. The Arab-D reservoir is the thickest and
found in the younger formations, especially along oldest of the four reservoirs in the Arab Formation
crests of large flexures where two vertical joint sets are and is overlain by the C-D anhydrite, the upper unit
recognized, including a dominant north-northeast to of the Arab-D Member. Ages for the Arab-D – Jubaila
northeast trend and a subsidiary north-northwest to succession are well constrained, with the Jubaila
northwest joint set (Hancock et al., 1984). Image Formation considered to be early Kimmeridgian
logs suggest that there are also east – west-trending (Arkell, 1952), whereas the overlying Arab Formation
fractures (K. A. T. MacPherson, 2005, personal com- is thought to be late Kimmeridgian to early Titho-
munication). These features may result from exten- nian (LeNindre et al., 1990; De Matos and Hulstrand,
102 / Lindsay et al.

FIGURE 4. Jubaila Formation west of


Riyadh in Wadi Leban. Ju2 member
is the upper, resistive beds. Ju1 member
is the lower, recessive beds. Ju2 cliff
former is 15 m (50 ft) thick.

stromatoporoid and coral banks and


overlying ramp-crest skeletal-oolitic
grain shoals, which subsequently pro-
graded over much of the area of the
intrashelf basins.
Overall, the paleoclimate was prob-
ably hot and arid, similar to the pres-
ent climate on the Arabian Peninsula
(Handford et al., 2002). The general
distribution of evaporites in the Hith
Formation, which formed in response
1995; Hughes, 1997, 2004). The D member is restricted to a restriction of the shelf, and especially of intrashelf
to the Kimmeridgian (Sharland et al., 2001). Although basins, from open-marine circulation, was also influ-
numerous benthonic foraminifera with stratigraphic enced by these intrashelf basins; evaporites are typi-
ranges consistent with a Kimmeridgian age have been cally thickest and most halite prone in the intrashelf
described from the Arab-D reservoir, none establishes basins.
a more precise date.
ARAB-D LITHOFACIES AND
PALEOGEOGRAPHIC SETTING AND REGIONAL DEPOSITIONAL ENVIRONMENTS
DEPOSITIONAL ENVIRONMENTS
The interpretations presented herein are based
During the Late Jurassic, a vast, shallow carbonate primarily on observations from core in the Shedgum
ramp extended from the central Arabian Peninsula, area of Ghawar field (Figures 7, 8 [the foldout located
west of the present site of Riyadh, as far east as the in the back of this publication]).
present Zagros Mountains of Iran, as far north as Previous workers (Mitchell et al., 1988) have de-
central Iraq and south into Oman (Figure 5). A thick veloped a classification scheme to organize Arab-D
succession of carbonate and evaporite sediments was rocks into genetically meaningful packages. This clas-
deposited on this shelf (Figure 1). Differential intra- sification divides Arab-D rocks into six depositional
plate subsidence, partly enhanced by peripheral aggra- lithofacies, which include one anhydrite and five car-
dation of shallow-marine carbonates, led to the devel- bonate lithofacies. Carbonate lithofacies are distin-
opment of intrashelf basins, including the Gotnia, guished on the basis of their typical depositional com-
Arabian, and Rub’Al Khali basins (Al-Husseini, 1997; ponents and include (1) skeletal-oolitic limestones
Ziegler, 2001), during deposition of the Callovian and dolomites (SO); (2) Cladocoropsis limestones and
Tuwaiq Mountain and Oxfordian Hanifa formations. dolomites (CLADO); (3) stromatoporoid-red (green)
Organic-rich carbonate source rocks were deposited algae-coral limestones (SRAC) (In this lithofacies, a
in these intrashelf basins during part of the Oxfordian common organic component is Thaumatoporella par-
(Ayres et al., 1982; Droste, 1990; Carrigan et al., 1994). vovesiculifera (Hughes, 1996), a microbial encruster that
These intrashelf basins tended to be maintained De Castro (1991, 2002) suggests is a chlorophycean,
through the remainder of the Jurassic, influencing the or green alga. The encrusting form gives it the appear-
deposition of the overlying Jubaila, Arab, and Hith ance, both megascopically and under hand lens or bin-
formations (Figure 6) and bounding broad, relatively ocular microscope, of a red alga and was originally
stable shallow shelf or platform areas on which the described as a red alga. The ‘‘SRAC’’ acronym is suffi-
Arab reservoir carbonates were deposited. The pe- ciently widely present in the literature that we continue
ripheral highs around the intrashelf basins are in- to use it despite the fact that few red algal fragments
terpreted to have been the loci for the initiation of actually exist in the lithofacies.); (4) bivalve-coated
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 103

FIGURE 5. Late Jurassic paleogeography of the Arabian Peninsula and surrounding area. Average directions of Arab-D
and Jubaila progradation are shown by arrows. Blue band west of Riyadh is the Tuwaiq escarpment. Modified from
Handford et al. (2002), Al-Husseini (1997), Ayres et al. (1982), R. B. Koepnick and L. E. Waite (1991, personal communication),
Murris (1980), and Scotese (1998).

grain-intraclast limestones (BCGI); and (5) micritic seal of the reservoir. Because a close relationship exists
limestones and dolomites (MIC). Because dolomitiza- between the depositional environments and the litho-
tion frequently destroys evidence of original depo- facies, both will be discussed together below.
sitional lithofacies, the diagenetic lithofacies dolomite
(DOLO) was added (Mitchell et al., 1988) (Figure 9). Anhydrite Lithofacies; Salina Environment
Although some later workers have sought to revise this Anhydrite (nonporous) with some interbedded car-
classification (Meyer and Price, 1993; Meyer et al., 1996; bonate stringers (some of which, especially where
Handford et al., 2002), these later modifications were dolomitized, are porous) makes up most of the C-D
not widely accepted and have since fallen out of use. anhydrite and overlies the Arab-D reservoir (Figure 8,
These lithofacies, with the exception of the dia- the foldout located in the back of this publication).
genetic dolomite lithofacies (DOLO), were deposited Typical anhydrite fabrics include nodular, bedded
approximately 58 south of the equator on a broad, nodular, nodular mosaic, and massive (classification
arid, storm-dominated carbonate ramp. The ramp of Maiklem et al., 1969); abundant, vertically aligned
consisted of the following subenvironments from (elongate) and uncommon mosaic anhydrite are also
upramp to downramp: (1) inner ramp; (2) ramp-crest observed (Figures 11, 12). In thin section, anhydrite
shoal; (3) proximal middle ramp; (4) distal middle locally exhibits a felted replacive texture of very small,
ramp; and (5) outer ramp (Figure 10). Each of these oriented, needlelike anhydrite crystals and clear, coars-
environments is characterized by one or two of the er, rectangular laths of diagenetic anhydrite. On the
depositional lithofacies. In addition, there was a sa- crest of the En Nala anticline in Ghawar field, capil-
lina environment that developed after the deposition lary pressure has been strong enough to force light-end
of the sediments of the Arab-D reservoir, depositing hydrocarbons, which look like wispy films, between
the dominantly anhydrite interval that is the vertical evaporite crystals and nodules.
104 / Lindsay et al.

FIGURE 6. Regional depositional environments of the Arab and Hith formations. Most of the time, deposition of high-
energy grainstones and packstones dominated the broad carbonate shelf. This was succeeded by later widespread
deposition of Hith evaporites. Modified from Ziegler (2001).
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 105

Deposition of the anhydrite occurred after a major


fall of relative sea level that subaerially exposed the
Arab-D carbonate, marking the top of the Arab-D car-
bonate. Small sink holes, filled with collapse breccia,
have been seen in outcrop (Tuwaiq escarpment) and
in Ghawar field in core and on image logs (Figure 13)
where anhydrite overlies and fills in the karstic fea-
tures. Following subaerial exposure, the initial rise of
sea level was slow, and the regional carbonate shelf
and associated interior basins, such as the Arabian
intrashelf basin, became the site of evaporite deposi-
tion (probably a combination of mostly gypsum and
some anhydrite) (Figures 11, 12). Historically, nodu-
lar mosaic to massive anhydrite, typical of the C-D
anhydrite, has commonly been interpreted as form-
ing within a sabkha environment (Kinsman, 1964;
Shearman, 1978). However, vertically aligned crys-
tals, thought to have been deposited originally as
subaqueous gypsum palmate crystals, and bedded an-
hydrite nodules in the C-D anhydrite suggest that
most of the anhydrite probably formed subaqueously
in a salina to peritidal to intertidal setting (McGuire
et al., 1993; Handford et al., 2002) (Figures 11, 12).
Recent detailed descriptions of the anhydrite in
two cored wells show that the sediments were de-
posited in cycles, with each evaporitic cycle starting
with subaqueous salina deposition and growth of
palmate crystals. Slight shallowing to a peritidal po-
sition between the salina proper and the intertidal
resulted in palmate crystals growing at an oblique
angle. Subsequent intertidal environments produced
flat to obliquely oriented gypsum crystals with lami-
nated carbonate mud. Salina and peritidal settings
are the most common, with the intertidal setting as
the least common. The laminae to very thin beds
that are the most common bedding in the evapo-
rite appear to be varvelike and form evaporite cou-
plets that contain thicker evaporite bases, interpret-
ed as summertime deposition, and carbonate- and
organic-rich laminae caps, interpreted as wintertime
deposition.
Uncommon, large transgressive events freshened
the highly saline setting to normal-marine salinity,
which allowed the subtidal carbonate factory to start
up and deposit a cycle of carbonate referred to as a
carbonate stringer. Sometimes, these only formed
very thin to thin beds, but occasionally deposited a
carbonate cycle 1.5–2 m (5–6 ft) thick, with the top
FIGURE 7. Shedgum Arab-D type log (open-hole log portion porous and grain rich. During the following
gamma-ray density-neutron), Ghawar field. Six cored
intervals are shown to the right, with a total of 91 m (299 ft) overall relative sea level fall, restriction and evapora-
cored and depth adjusted to match open-hole logs. Core tion recommenced, and the subtidal evaporite factory
coverage is through the complete Arab-D reservoir. again started to deposit subaqueous evaporites.
106 / Lindsay et al.

FIGURE 9. Summary of characteristics and distribution of carbonate lithofacies in the Arab-D reservoir, Ghawar field.
Modified from Mitchell et al. (1988).

Inner Ramp a normal-marine passageway, through the ramp-crest


The inner ramp, landward and upslope from the shoal, that connected part of the inner ramp with the
landwardmost extent of sediment agitation by sig- proximal part of the middle ramp.
nificant wave activity and behind the shoal crest,
was a lagoon with localized intertidal islands com- Skeletal-oolitic Lithofacies (SO); Ramp-crest
posed of wackestone, packstone, and some grain- Shoal Environment
stone (Figure 10). The sediments of the inner-ramp Skeletal-oolitic (SO) limestone and dolomitic lime-
lagoonal setting are only present at the top of the stone is present in the upper part of the Arab-D reser-
Arab-D reservoir (zone 1) and are approximately 3 m voir (zone 2a) and another interval deeper in the res-
(10 ft) thick (Figure 8, the foldout located in the back ervoir (zone 2b) (Figure 9). This lithofacies is typically
of this publication). This depositional setting con- well sorted and contains grainstone, mud-lean pack-
tains a restricted fauna, suggesting a lack of normal- stone, and some mud-rich packstone (Figures 8 [the
marine circulation. Normal-marine fauna, such as the foldout located in the back of this publication] and
delicate stick-shaped stromatoporoid Cladocoropsis, 14). Overall, the lithofacies represent the best sorted
have been described locally within this setting and and, at the megascopic visual level, the most uniform-
may owe their presence to either storm transport or to appearing carbonate rock type of the Arab-D. Major
FIGURE 10. Idealized depositional model of the Arab-D reservoir carbonate ramp in Ghawar field. This model depicts deposition of one carbonate cycle from the
inner ramp, ramp-crest shoal, proximal, and distal middle ramp to outer ramp. Blue is a transgressive mud-rich deposit capped by a firmground that developed
during maximum flooding and was burrowed by Thalassinoides. Brown with diamond shapes is distal middle-ramp debris-flow deposits of rudstone and floatstone.
Green is proximal middle-ramp biostromes and mounds of stromatoporoids and corals. Blue with fork shapes is proximal middle-ramp Cladocoropsis banks
deposited in sheltered areas between and upslope of biostromes and mounds. Red is the cross-bedded ramp-crest grainstone shoal. Purple is inner-ramp lagoon-
intertidal islands. Those facies and rock types that are abundant in each reservoir zone are underlined by the titles and associated arrows zone 1, zone 2, and zone 3.
Distribution of the common to abundant faunal and floral elements (both laterally at any one time and stratigraphically through the reservoir zones) is shown
below the figure.
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 107
108 / Lindsay et al.

FIGURE 11. Arab C-D evaporite from 2071 – 2073 m (6796 – 6804 ft) in top of core 1 in the Shedgum Arab-D core, Ghawar
field. Evaporite crystals were originally gypsum that grew on the floor of a salina. Note the varvelike cyclicity and
enterolithic folds.

grain types include foraminifera, dasycladacean algae, day-to-day energy to have affected the carbonate ramp
micritized grains and foraminifera, ooids, and bivalves (Figure 10). Fair-weather-wave base is defined as a zone
(Figures 14 – 16). Porosity, from most abundant to where waves first touched the sea floor and continued
least abundant, includes interparticle, moldic, intra- to wash the sea floor with high energy, until friction
particle, microporosity, and, locally common within reduced wave energy, and sediment movement ceased
dolostones, intercrystal pores (Figures 14–16). Cross- (zone Y of Irwin, 1965). The grain-rich setting is cross-
bedding and horizontal laminations are common bedded and slightly cemented by fine, isopachous, ma-
(Figures 8 [the foldout located in the back of this rine, and microspar cement. Skeletal detritus includes
publication] and 14). Other common features include storm-derived fragments and larger to nearly whole
firmgrounds and hardgrounds (Figure 8, the foldout Cladocoropsis, stromatoporoids, and corals from farther
located in the back of this publication), burrows, bor- seaward. Microbial encrustations are fairly common.
ings, and coarsening-upward beds. Common types of The ramp-crest shoal depositional setting and as-
diagenesis include dissolution and recrystallization, sociated skeletal-oolitic lithofacies form the upper
fine isopachous bladed calcite and microspar cementa- Arab-D reservoir (most of zone 2a) (Figures 8 [the fold-
tion, and dolomitization. Although scattered dolomite out located in the back of this publication] and 10),
crystals are commonly found, there does not appear to as well as the uppermost part of the lower sequence
be any complete dolomitization of this lithofacies. (uppermost Jubaila Formation).
The skeletal-oolitic lithofacies were deposited in
the ramp-crest shoal environment, consisting of a se- Cladocoropsis Lithofacies (CLADO); Sheltered
ries of ooid-peloid-intraclast sand shoals that were Proximal Middle-ramp Environment
locally and briefly subaerially exposed as islands in a Cladocoropsis (CLADO) limestone and dolostone
high-energy very shallow subtidal to intertidal setting contain greater than 10% of the branching stromato-
in fair-weather-wave base. It was continuously wave poroid Cladocoropsis mirabilis (Champetier and Four-
washed and occasionally storm swept by the highest cade, 1966) and occur in the upper and middle Arab-D
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 109

FIGURE 12. Arab C-D evaporite


from 6804 to 6809 ft (2073 to
2075 m) in top of core 1 in the
Shedgum Arab-D core, Gha-
war field. Evaporite crystals
were originally palmate gyp-
sum that grew on the floor of a
salina. Note varvelike cyclicity.
Carbonate at the base has
been dolomitized. Cores are
2.5 in. (6.35 cm) wide.

dolomitized, Cladocoropsis
fragments typically undergo
dissolution to form moldic
porosity and, if originally in
contact with each other, de-
velop super-K (high-per-
meability) intervals.
The Cladocoropsis (CLADO)
lithofacies, along with the
stromatoporoid-red algae-
coral (SRAC) lithofacies, de-
scribed below, was deposited
in a proximal middle-ramp
environment, with deposi-
tion just within fair-weather-
wave base, where constant
daily circulation occurred.
This environment was char-
reservoir (zones 2a and upper 2b) (Figures 8 [the acterized by grain-rich, domed, and encrusting stro-
foldout located in the back of this publication] and matoporoid and coral biostromes and mounds.
17). Cladocoropsis rudstones and floatstones contain Intermound sheltered areas and sheltered areas im-
a matrix of grainstone, mud-lean packstone, and mediately upslope of the biostromes and mounds
mud-rich packstone. They are not as well sorted as were populated by Cladocoropsis (Figures 8 [the fold-
those of the skeletal-oolitic lithofacies and are there- out located in the back of this publication] and 10)
fore more heterogeneous (Figure 17). Cladocoropsis (Leinfelder et al., 2005). This interpretation is based, in
fragments, dasycladacean algae, foraminifera, and mi- part, on comparison with domed and branched co-
critized grains and forams are common (Figure 17). rals in modern environments ( James, 1983), where
Pore types, from most common to least common, are domed forms occur in biostrome and mound envi-
interparticle, intraparticle, moldic, and microporous, ronments with moderate wave energy and low rates
with all types abundant (Figure 17), and when do- of sedimentation, whereas branched forms, such as
lomitized, intercrystal and moldic porosity are the Cladocoropsis, occupy lower energy sites with higher
dominant pore types. Sedimentary structures are sub- rates of sedimentation. Cladocoropsis is commonly
tle, but close examination of cores shows that cross- accompanied by the foraminifera Kurnubia palasti-
bedding and horizontal laminations are present in niensis and Nautiloculina oolithica (Figures 15, 16). In
highstand parts of cycles (parasequences), whereas the shallower marine moderate-energy SRAC litho-
soft-body bioturbation and firmgrounds and hard- facies, the microbiota is dominated by a dasyclad
grounds are more common in transgressive parts of alga attributed to Clypeina sulcata; the green alga
cycles (Figure 8, the foldout located in the back of this T. parvovesiculifera is common and is joined by a
publication). Diagenesis is commonly limited to dolo- background population of the forams Kurnubia sp.,
mitization, dissolution, recrystallization, and uncom- Nautiloculina sp., Redmondoides lugeoni, and some
mon equant calcite cementation. When completely biserial agglutinated forms (Figures 15, 16). Microbial
110 / Lindsay et al.

FIGURE 13. A. Examples of breccia-filling sinkholes in the Arab-D in


Ghawar field and in outcrop. The upper photograph is a small sinkhole
in the Arab Formation filled with fitted-fabric and chaotic collapse breccia
in Wadi Birk, 150 km (93 mi) south of Riyadh. Red baseball cap is at the
base of the breccia, and each geologist stands at the side of the breccia
with rock hammer (left) and hand (right) on the breccia-country rock
contact. B. Formation MicroImager image log in a horizontal well in the
south part of Ghawar, with rounded chaotic breccia just beneath the
Arab-D composite sequence boundary and overlying Arab C-D evaporite
(left). (Scale is in feet.) C. Photograph from the central part of Ghawar in
a vertical well. Rounded chaotic breccia has red and brown (terra rosa)
geopetal between breccia clasts, along with plant rootlets. Overlying Arab
C-D evaporite contains palmate gypsum crystal ghosts. (Photo is 6.4 cm
[2.5 in.] wide.)
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 111

tified Shuqria, an earlier relative of Cladocoropsis, to


have populated sheltered intermound areas between
stromatoporoid and coral biostromes and mounds.
Cladocoropsis probably populated a similar depositional
setting. Storm events reworked Cladocoropsis and have
strewn them within the proximal middle ramp, and
more powerful storms reworked fragments back into
the ramp crest (Figure 10).

Stromatoporoid-Red Algae-Coral Lithofacies


(SRAC); Biostrome-mound Proximal
Middle-ramp Environment
Stromatoporoid-red algae-coral (SRAC) limestones
and dolomitic limestones occur in the middle Arab-D
reservoir (zone 2b) and represent the most heterog-
eneous rock types (Figures 8 [the foldout located in
the back of this publication] and 18). Locally, this
lithofacies is present in the lower Arab-D reservoir as
allochthonous debris-flow deposits (Figures 8 [the fold-
out located in the back of this publication] and 10).
This lithofacies is, in general, extremely poorly sorted
FIGURE 14. Ramp crest shoal skeletal-oolitic lithofacies and contains both pores and grains (Figure 18) that
composed of cross-bedded, intraclast-peloid-coated grain- exhibit a wide range of sizes and shapes. Common
ooid grainstone with abundant interparticle porosity grain types include domed, digitate, and encrusting
and some moldic, intraparticle, and microporosity. Ce- stromatoporoids, corals, foraminifera, micritized grains,
mentation was by very fine-bladed calcite and microspar
and microbial encrustations (Figure 18). Scleractinian
cement. Note minor physical and chemical compaction
in the upper and more prevalent compaction in lower corals found in the Arab-D reservoir were originally
photomicrographs. Photomicrograph widths 1.1 mm (up- composed of metastable aragonite and were suscep-
per) and 2.25 mm (lower), Shedgum area. tible to dissolution during diagenesis (Figure 18) and
form large, irregular molds. Stromatoporoids and cor-
als form floatstone to rudstone fabrics (Figure 18),
encrustations are also common (Figure 8, the foldout with a matrix of grainstone, mud-lean packstone, and
located in the back of this publication). mud-rich packstone. Moldic and intraparticle pores
Considerable facies variability exists within and are most common in this lithofacies, with interparticle
between the biostromes and mounds versus shel- pores subordinate and micropores common. Typical
tered parts of this depositional environment. These diagenetic modifications include dissolution, bioero-
two lithofacies contain diverse, normal-marine biota, sion, recrystallization, dolomitization, anhydrite em-
which include common to abundant foraminifera, al- placement, fine isopachous bladed calcite and equant
gae, and microbial encrustations, along with marine calcite cementation, and microspar cementation.
cement, which are evidence of deposition under normal- High-energy storm events buffeted biostromes and
marine conditions. Fairly common cross-bedding also mounds and reworked them locally, with stronger
argues for a moderate to potentially high-energy set- storms reworking bioclasts and grains downslope
ting. Hermatypic corals imply deposition in the photic as debris flows into the distal middle-ramp setting
zone, whereas the presence of such stenotopic organ- (Figure 10) and commonly, but less abundantly, trans-
isms (organisms that have a limited tolerance for porting fragments upslope into the ramp-crest and
changes in environmental conditions) as coral indi- inner-ramp environments. The former effect is seen in
cates normal-marine conditions. outcrops of the Jubaila Formation in the Tuwaiq escarp-
Cladocoropsis have also been reworked with no in- ment west of Riyadh where stromatoporoid and coral
place, upright-standing organisms being identified biostromes and mounds are virtually all reworked. Only
and very few actual intact branches being preserved. one in-place stromatoporoid-coral biostrome, 1 m (3 ft)
Fieldwork in the Tuwaiq Mountain Formation out- thick and 30 m (100 ft) in width (Figure 19), has been
crops in the Tuwaiq escarpment, west of Riyadh, iden- identified in Wadi Hilwah (Hughes, 2004). All other
112 / Lindsay et al.

FIGURE 15. Shedgum area


Arab-D and Jubaila significant
biocomponents. Width of view
is in millimeters. 1 = brachio-
pod valves (2 mm); 2 = cerithid
gastropod (4 mm); 3 – 4 = Tro-
cholina palastiniensis (1 mm);
5–6 = Quinqueloculina sp. (1 mm);
7 – 8 = Pfenderina salernitana
(1 mm); 9 – 10 = Mangashtia
viennoti (1 mm); 11 = Alveosepta
jacardi (1 mm); 12 – 13 = Red-
mondoides lugeoni (1 mm); 14 =
Nautiloculina oolithica (1 mm);
15 = Kurnubia palastiniensis
(1 mm); 16 – 17 = Clypeina sulcata
(1 and 2 mm); 18 = Heteroporella
jaffrezoi (1 mm); 19 = coral
(8 mm); 20 = Cladocoropsis mira-
bilis (4 mm); 21 = Thaumato-
porella parvovesiculifera (2 mm);
22 = tetraxon sponge spicule;
23 –24 = Lenticulina sp. (1 mm).

(Figure 20) and moldic, com-


monly from dissolution of
metastable bivalves, but in-
cludes a significant admixture
of microporosity and intrapar-
ticle pores. Diagenetic modifi-
cations noted in this lithofacies
include dissolution, recrystal-
lization, dolomitization, stylo-
litization, and equant calcite
cementation.
The BCGI lithofacies,
along with the micritic litho-
facies described below, are
the typical sediments of the
outcrops show stromatoporoids and corals to have distal middle ramp, where deposition occurred be-
been at least locally reworked in the proximal middle neath and seaward of fair-weather-wave base. The dom-
ramp, or they have been reworked farther downdip inant sediment is micritic to very fine-grained sediment
into the distal middle ramp (Figure 10). that is capped by firmgrounds (Figure 20) character-
ized by Thalassinoides burrow systems (Figure 10) and
Bivalve-coated Grain-intraclast Lithofacies (BGCI); overlain by storm-derived rudstone and floatstone of
Distal Middle-ramp Environment inner-ramp, ramp-crest, and proximal middle-ramp
Bivalve-coated grain-intraclast (BCGI) limestone bioclasts, most commonly, the BCGI lithofacies but
and dolomitic limestone occur mostly in the lower locally including stromatoporoids and corals (SRAC)
Arab-D reservoir (zones 3a, 3b, and 4) as rudstones (Figure 20).
and floatstones, with grainstone, mud-lean packstone, Burrowing organisms (infauna) were common, and
mud-rich packstone and uncommon wackestone ma- either these created vertically oriented Thalassinoides
trix (Figures 8 [the foldout located in the back of this burrows in micritic to very fine-grained muds, or soft-
publication] and 20). Bivalves, coated grains, intra- body burrowers have homogenized the sediment.
clasts, micritized grains, and foraminifera are domi- Preserved sedimentary structures are uncommon,
nant grain types. Porosity is dominantly interparticle other than those created by bioturbation. These factors
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 113

FIGURE 16. Biozonation and key biocomponents of the Arab-D Member of the Arab Formation and Jubaila Formation in
the Shedgum Arab-D core, Ghawar field.

FIGURE 17. Proximal middle-ramp Cladocoropsis rudstone FIGURE 18. Proximal middle-ramp stromatoporoid-coral
with a packstone matrix containing interparticle, moldic, rudstone, containing a mud-rich packstone matrix and
intraparticle, and microporosity. All Cladocoropsis frag- interparticle, moldic, intraparticle, and microporosity.
ments have been reworked by storms. Note the small mi- Stromatoporoids have been reworked by storms and bio-
cropellets in the matrix (upper). Matrix is composed of eroded by bivalves, still visible. Coral mold, with casts of
intraclasts, micritized forams, micropellets, dasyclad algae coralites, is beneath 223 (lower left). Note the micropellets
(Clypeina sulcata), and stromatoporoid fragments (lower). that acted like microball bearings during storm reworking.
Photomicrograph widths are 1.1 mm (upper) and 7.4 mm Both photomicrograph widths are 1.1 mm, Shedgum Arab-
(lower), Shedgum core. D core on display.
114 / Lindsay et al.

FIGURE 19. Upper photo is Arab-D –


Jubaila contact in Wadi Hilwah. Geolo-
gist (encircled) is standing by the only
in-place coral-stromatoporoid bios-
trome found in outcrop. All others have
been reworked by storms. Arab-D –
Jubaila contact is about 15 ft (4.5 m)
above the geologist. Top of Arab-D is at
the hill top. Lower photo is Arab-D –
Jubaila contact in Wadi Nisah. Arab-D
is the thin, platy outcrops above geol-
ogists, whereas Jubaila is the thicker
beds beneath their feet. Note the large
storm ripple in Jubaila.

setting (Figure 10). Microbial encrus-


tations, coated grains, and ooids in-
dicate that original, coarse-sediment
formation was farther upslope in
the photic zone and fair-weather-
wave base, in shallow-marine wa-
ter of sufficient turbulence to move
grains around and achieve relative-
ly even coatings of grains. Muddy
intraclasts, ripped up from areas of
mud deposition down the ramp, is
another indicator of brief, high-
energy storm events.
The matrix associated with rud-
stone and floatstone tends to be mud-
rich packstone, with mud-lean pack-
stone and grainstone less common.
The muddy matrix contains very
suggest that the sediment accumulated in quiet, low- small micropellets, many of which may have been
energy conditions beneath fair-weather-wave base created by the fruiting bodies of microbial encrusta-
away from strong wave or current action. tions (R. R. Leinfelder, 2003, personal communica-
Firmgrounds formed on the flooding surfaces of tion). The entrained mud-rich matrix and micropellets
individual carbonate cycles. Firmgrounds are typified acted like microball bearings to mediate transport
by vertical Thalassinoides burrows, along with side of storm-generated debris flows much farther
gallery living areas with an average vertical depth of downslope. The relative proportion of mud incorpo-
penetration of 1 m (3 ft) and a range of vertical pene- rated into debris flows was responsible for two types
tration from only a few tenths of a meter (<1 ft) to as of deposits. One is a cohesive debris flow with a
much as 2.4 m (8 ft) (Figure 21). Thalassinoides bur- mud-rich matrix (Figure 20), and the other is a non-
rowers easily mined downward through soft muddy cohesive debris flow with a less mud-rich matrix
sediment, but ceased excavation once a grain-rich sub- (Figure 20). Additional mud matrix assisted debris
strate, containing larger grains and skeletal detritus, flows in transporting sediment farther down the
was intersected. ramp slope.
Bivalve-coated grain-intraclast (BCGI) limestone Weaker storm events were responsible for rework-
and, to a lesser extent, stromatoporoids and corals ing bivalve-coated grain-intraclast (BCGI), whereas
of the SRAC lithofacies were deposited as rudstone larger, more powerful storms were required to dis-
and floatstone by storm-induced high-energy debris lodge and transport stromatoporoids and corals be-
flows that transported debris and matrix downslope cause they formed biostromes and mounds and were
into a deeper, shallow-marine distal middle-ramp better anchored than were BCGI lithofacies.
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 115

the surrounding lime mudstone matrix is only partially


dolomitized.
As described above, micritic (MIC) limestones,
bivalve-coated grain-intraclast (BCGI) limestones, and,
to a lesser extent, stromatoporoid-red algae-coral (SRAC)
limestones, are typically interbedded in the distal
middle-ramp environment. In the outer-ramp environ-
ment, beyond the reach of storm-generated debris flows
(Figure 10), lime mudstones (micritic lithofacies) were
the only sediments deposited. Each cyclic package is
capped by a firmground (Figures 8 [the foldout located
in the back of this publication] and 10) that are typified
by Thalassinoides burrows that penetrated the muddy
carbonate layer. Although a lagoon paleoenvironment
has been suggested for this lithofacies (Enay, 1987),
the presence of the forams Lenticulina spp., Nodosaria
spp., and Pseudomarsonella spp., along with monaxon
and tetraxon sponge spicules, coccoliths, and juvenile
costate brachiopods, is more consistent with open,
normal-marine conditions below fair-weather-wave
base, in perhaps 25 – 30 m (80– 100 ft) of water depth.
Modern assemblages, including Lenticulina and Nodo-
saria, are confined to water depths greater than 100 m
FIGURE 20. Distal middle-ramp bivalve-coated grain-
intraclast rudstone. The upper part is the cohesive debris
flow (mud rich) and the lower part is the noncohesive
debris flow (grain rich). Debris flows were created by storms
and rest atop firmgrounds burrowed by Thalassinoides.
Photomicrograph widths are 7.4 mm (upper) and 2.2 mm
(lower), Shedgum Arab-D core.

Micritic Lithofacies (MIC); Distal Middle-ramp


and Outer-ramp Environments
Micritic (MIC) limestone and dolostone (Figures 8 [the
foldout located in the back of this publication] and 21)
are common and tend to dominate the lower half of
the Arab-D reservoir (zones 3a, 3b, and 4), less common
in the middle Arab-D reservoir (zone 2b), and least
common to nonexistent in the upper Arab-D reservoir
(zone 2a and 1). In contrast to the other depositional
lithofacies, micritic limestones are mostly mud sup-
ported. Grains present in this lithofacies include mi-
critized grains, bivalves, ostracods, foraminifera, and
FIGURE 21. Outer ramp mudstone, bioturbated by
intraclasts. Although porosity is generally low, typical Thalassinoides burrows and by soft-body burrowers, that
pore types include microporosity, moldic, and intra- contains only low amounts of microporosity. In the upper
particle pores. In addition, interparticle and occasional photomicrograph, note a few dolomite crystals and one
intercrystal porosity is preserved in the grain-rich sedi- quartz silt grain (center) in the pelleted matrix. In the
lower photomicrograph is the contact of a Thalassinoides
ment that filled vertically oriented Thalassinoides bur-
burrow tube that was filled with fine sediment and partially
row tubes (Figure 21). Diagenetic alterations include dolomitized and the mud-rich matrix containing a few
dolomitization, dissolution, recrystallization, and stylo- windblown quartz silt grains. Photomicrograph widths are
litization. Burrows may be highly dolomitized, whereas 1.1 mm (upper) and 7.4 mm (lower), Shedgum Arab-D core.
116 / Lindsay et al.

typically consists of anhedral to euhedral crystals


(30– 500 mm in size) that can be sucrosic or mosaic,
depending on whether intercrystal porosity is pres-
ent. Relict burrows and firmground surfaces are lo-
cally visible (Figure 23). Intercrystal porosity is domi-
nant, with moldic porosity common (Figures 22, 23)
and fracture porosity less common. Thin layers of su-
crosic dolomite locally have sufficiently high poros-
ity that they have anomalously high permeability re-
ferred to as super-K. Uncommon intracrystal porosity
developed when dolomite crystals were partially dis-
solved to form dolo-donuts (Figure 23). Dolomite crys-
tals commonly have cloudy, inclusion-rich centers and
clear, limpid rims (Figures 22, 23). Uncommonly, three-
to four-generation dolomite crystals have been identi-
fied that contain inclusion-rich centers, surrounded
by a clear, limpid layer, with an additional inclusion-
rich layer, followed by a clear, limpid, outermost layer.
Dolomitization of Cladocoropsis lithofacies has lo-
cally resulted in the dissolution of the Cladocoropsis
fragments and the development of moldic porosity.
FIGURE 22. Proximal middle-ramp dolomitized, cross-
bedded, Cladocoropsis, peloid rudstone (upper) and dolo-
mitized, cross-bedded mud-lean packstone (lower), con-
taining moldic and intercrystal porosity. The upper
photomicrograph contains saddle dolomite with ghosts
of peloids. Photomicrograph widths are 2.25 mm (upper)
and 1.1 mm (lower), Shedgum Arab-D core.

(330 ft), but the Mesozoic equivalents are known to


have occupied depths shallower than this.
The outer ramp was probably affected by storms.
Using the Persian Gulf as a modern analog, yearly
storms have been found to stir bottom sediment to a
depth of 50 m (164 ft), whereas a 100-yr storm can stir
bottom sediment to 110 m (360 ft) (Saudi Aramco,
oceanographic department database). However, bio-
turbation from soft-body burrowing organisms and
Thalassinoides probably churned the muddier sedi-
ment and erased most traces of storm events. No
rudstone or floatstone deposits were reworked and
carried by debris flows this far down the ramp from
the inner ramp, ramp crest, and proximal middle ramp
(Figure 10), so distinction between burrow fill and
FIGURE 23. Dolomitized, cross-bedded, Cladocoropsis,
the surrounding matrix is subtle except in those in- skeletal, peloid rudstone, with a mud-rich packstone
stances where differential partial dolomitization re- matrix and containing intercrystal and moldic porosity
sulted from slight depositional differences. and very high permeability (super-K). Long laths are
anhydrite crystals. The lower part is dolomitized, peloid
Dolomite Lithofacies (DOLO) mud-lean packstone at a potential subaerial exposure
surface, with plant rootlets and/or Skolithos – Arenicolites
Dolomite (DOLO) (Figures 8 [the foldout located
burrow traces. The pore space is lined with insoluble
in the back of this publication], 22, 23) generally oc- residue and one dolo-donut intraparticle pore. Photomi-
curs as thin to thick (0.3–4.6-m, 1–14-ft), sheetlike beds crograph widths are 2.24 mm (upper) and 1.1 mm
throughout most of the Arab-D reservoir. Dolomite (lower), Shedgum Arab-D core.
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 117

If individual moldic pores are touching or if the matrix posited the Arab-D reservoir as a series of submeter-
dolomite has sufficient intercrystal porosity, they de- to meter-scale and thicker carbonate cycles (para-
velop thin intervals with low porosity but very high sequences), with each carbonate cycle composed of
permeability, also referred to as super-K (Figure 23). one to three beds. Approximately five individual car-
The least common dolomite occurrence is saddle bonate cycles combine to form a cycle set (parase-
(baroque) dolomite with crystals 500 mm up to a few quence set), with two to six cycle sets forming a HFS
thousand micrometers in size (Figure 22). This type and four to five HFSs forming a composite sequence.
of dolomite has been interpreted to result from hy- Two composite sequences form the Arab-D reservoir.
drothermal emplacement along widely spaced frac- One composite sequence forms the Arab-D Member
ture systems (Cantrell et al., 2004). Distribution is of the Arab Formation; the other forms the upper
areally restricted, but may extend vertically through part of the Jubaila Formation (Figure 24). It is tempt-
much of the reservoir. ing to relate these different cycle scales to Milanko-
vitch cycles (Milankovitch, 1930, 1941).
ARAB-D RESERVOIR VERTICAL SUCCESSION
AND SEQUENCE STRATIGRAPHY Carbonate Cycles (Parasequences)
On the individual carbonate cycle (parasequence)
In general, from bottom to top, the Arab-D reser- scale, the Arab-D reservoir is composed of a trans-
voir consists of (1) outer-ramp, mud-rich carbonate gressive mudstone, wackestone, or packstone base,
cycles (parasequences); (2) distal middle-ramp, mud- deposited below fair-weather-wave base, and over-
rich, and rudstone- and floatstone-capped cycles; lain by highstand cross-bedded grainstone, mud-lean
(3) proximal middle-ramp biostromes and mounds of packstone, or mud-rich packstone, deposited with-
stromatoporoids and corals and sheltered Cladocoropsis in fair-weather-wave base and/or storm-wave base
bank cyclic deposition; (4) ramp-crest shoal, high- (Figures 8 [the foldout located in the back of this
energy cyclic deposition; and (5) inner-ramp lagoon publication], 10, 24).
and localized intertidal islands that form a few cycles A total of 120 cycles (parasequences) have been
at the top of the section (Figures 8 [the foldout located identified in the Shedgum Arab-D core presented
in the back of this publication] and 10). here (Figures 8 [the foldout located in the back of this
This shoaling-upward succession can be subdi- publication] and 24). A Fischer plot of these cycles
vided into 11 informal biozones, with the depths for has been constructed and annotated to show which
the Shedgum Arab-D well shown (Figure 16). In this cycles were deposited in outer-ramp, distal to proxi-
figure, the term ‘‘outer ramp’’ has to be understood in the mal middle-ramp, ramp-crest, and inner-ramp depo-
context of normal, tropical carbonate deposition to in- sitional settings (Figure 24). Individual carbonate cycles
dicate water depths of perhaps as much as 30 m (100 ft). (parasequences), from bottom to top, in the Arab-D
In each of the biozones, there is a characteristic asso- reservoir consist of the cycles discussed below.
ciation of biotic elements indicative of the water depth
and associated paleoenvironment shown (Figure 16). Outer-ramp Cycles
Outer-ramp and distal middle-ramp cyclic depo- Outer-ramp carbonate cycles are submeter to meter
sition was seaward and beneath fair-weather-wave scale and occasionally thicker and consist of mudstone,
base, but within or at least influenced by transport wackestone, and local mud-rich packstone that have
from sediments that were within storm-wave base. an uppermost firmground with Thalassinoides burrows,
Proximal middle-ramp and ramp-crest shoal depo- with no highstand grain-rich cap (Figures 8 [the fold-
sition was within fair-weather-wave base, and inner- out located in the back of this publication], 10, 24),
ramp deposition was landward of wave base. This although a thin layer of highstand-equivalent sediment
succession of carbonate deposition on a carbonate may be preserved at some localities. Thalassinoides-
ramp recorded an overall shallowing-upward history burrowed firmgrounds represent the maximum flood-
related to a long-term base-level fall (Mitchell et al., ing surface (MFS) of an individual cycle.
1988; Meyer and Price, 1993; Handford et al., 2002).
This overall shallowing-upward, long-term base- Distal Middle-ramp Cycles
level fall and loss of accommodation space was punc- Distal middle-ramp carbonate cycles are sub-
tuated by a series of smaller relative sea level rises and meter to meter scale and thicker and contain trans-
falls (Figure 8, the foldout located in the back of this gressive mud-rich bases, similar to the outer-ramp
publication). These rises and falls of sea level de- setting, with the maximum flooding surface (MFS)
118 / Lindsay et al.

FIGURE 24. Fischer plot of the Shedgum Arab-D cored interval. Each carbonate cycle (parasequence) is color coded to
show were it was deposited on the carbonate ramp. The small dimple on each cycle represents maximum flooding.

of the cycle at the top of the mud-rich base, where stand rudstones-floatstones thicken, and transgres-
a firmground with Thalassinoides burrows formed. sive cycle bases thin to a point where they are of equal
Thalassinoides burrows in Jubaila and Arab Formation thickness (Figure 10).
outcrops are spaced 0.1 m (4 in.) from each other.
Highstand progradation caused by relative sea level Proximal Middle-ramp Cycles
fall brought the proximal middle ramp, ramp crest The proximal middle ramp contains a thinner
shoal, and inner ramp close enough so that storms transgressive base, with the maximum flooding sur-
could rework inner-ramp, ramp-crest, and proximal face forming a firmground or the top of a bioturbated
middle-ramp mud, grains, and skeletal bioclasts interval and a thicker highstand cap composed of bio-
downslope as debris flows that covered the firm- stromes and mounds of stromatoporoids and corals
ground surface with grainstone or rudstone and float- and Cladocoropsis banks. Cladocoropsis banks were de-
stone (Figures 8 [the foldout located in the back of posited in sheltered areas between biostromes and
this publication], 10, 24). In some cases, more than mounds and in sheltered areas immediately upslope
one debris flow (evidenced by two or more stacked, of the biostromes and mounds, in slightly deeper parts
fining-upward packages) covered the firmground, of the proximal middle ramp (Figures 8 [the foldout
attesting to multiple storms passing through the area. located in the back of this publication], 10, 24). Stro-
Farthest downslope, in the distalmost parts of the mid- matoporoid and coral biostromes and mounds were
dle ramp, highstand rudstone and floatstone caps reworked by storms, although stromatoporoids and
are thin and are locally reduced to coarse packstones. corals have been found to be still in place (Figure 19).
In contrast, upslope and closer to the transition from Cladocoropsis was easily reworked by storms, and none
proximal to distal parts of the middle ramp, high- have been found in their original growth position.
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 119

Ramp-crest Shoal Cycles in a deeper shallow-marine setting, shallowing upward


The ramp-crest shoal contains thinner transgres- only extended into the influence of storm-wave base
sive bases that are only locally preserved, with the (Figure 10).
remainder eroded away or not deposited as wave base
stirred the floor of the ramp crest and continuously High-frequency Sequences
washed grains with high energy and deposited a thin In the Shedgum Arab-D core on display, eight HFSs
to thick cross-bedded highstand cycle cap (Figures 8 have been identified; the top of a possible ninth HFS
[the foldout located in the back of this publication], occurs at the base of the cored interval (Figures 8 [the
10, 24). In this setting, transgressive bases of HFSs foldout located in the back of this publication] and
may contain Cladocoropsis, stromatoporoids, and co- 24). Because the entire reservoir interval was depos-
rals. Storms during the highstand also reworked frag- ited in an overall shallowing-upward depositional sys-
ments of Cladocoropsis, stromatoporoids, and corals tem with progressive loss of accommodation space,
into cross-bedded, intraclast-peloid-ooid grainstone each successive HFS has a somewhat different and,
and mud-lean packstone (Figure 10). therefore, unique, lithofacies succession from those
above and below it.
Inner-ramp Cycles The general stacking pattern of HFSs is a transgres-
A lagoon formed behind the ramp crest and included sive systems tract (TST) of mud-rich cycles and a high-
uncommon, small intertidal islands (Figures 8 [the stand systems tract (HST) of grain-rich cycles, with the
foldout located in the back of this publication], 10, maximum flooding surface (MFS) at the top of the
24). Transgressive bases contain bioturbated, mud- mud-rich or muddier cycle (Figures 8 [the foldout
rich to grain-rich strata with Thalassinoides-burrowed located in the back of this publication], 10, 24). In the
firmgrounds overlain by highstand-associated grain- lower part of the Arab-D reservoir (zone 3), HFSs are
rich caps. These form the last few cycles at the top of the 46 – 52 ft thick, with thick transgressive systems tracts
Arab-D reservoir. Intertidal islands, thought to range (TSTs) and thin highstand system tracts (HSTs). Toward
in size from a few well spacings to less than one well the top of the lower part of the reservoir, transgressive
spacing (one well space = 1 km [0.6 mi]) in aerial ex- systems tracts (TSTs) become thinner, and highstand
tent, have been identified in a few cored wells. Gen- system tracts (HSTs) thicken in an expected trend. In
erally, the last cycle of deposition of Arab-D carbonate the upper part of the Arab-D reservoir (zones 2 and 1),
was a transgressive lagoon cycle with a Thalassinoides- HFSs progressively thin upsection from 46, 40, 29, 19,
burrowed firmground that lacked a highstand cap. We and finally, to 9 ft (14, 12, 9, 6, and finally, to 2.7 m)
interpret this facies to be the upslope deposit of a (Figure 8, the foldout located in the back of this publi-
series of cycles deposited during a forced regression, cation). This thinning of HFSs demonstrates an overall
terminating the deposition of the Arab-D. shallowing-upward history and base-level fall with the
progressive loss of accommodation space. Transgres-
Cycle Sets (Parasequence Sets) sive systems tracts (TSTs) are initially thick and drama-
A total of 28 cycle sets (parasequence sets) has been tically thin, whereas HSTs become the thicker and
identified in the Shedgum Arab-D core (Figures 8 [the dominant overall deposition in the ramp crest (Figure 8,
foldout located in the back of this publication] and the foldout located in the back of this publication).
24). These cycle sets are generally made up of approxi- In the lower part of the reservoir, transgressive sys-
mately five individual cycles (parasequences), ranging tems tracts (TSTs) tend to be composed of outer-ramp
from two to six, which record an increase in accom- and distal middle-ramp, mud-rich cycles (Figure 10).
modation space that progressively declined with sub- At the base of the cored interval, HSTs are composed
sequent deposition. In the Arab-D reservoir, the last of thin, storm-generated rudstone and floatstone with
few shallowing-upward cycles may thicken, which a mud-rich matrix (Figure 8, the foldout located in the
implies that the earlier deposited cycles did not back of this publication). Upsection, transgressive sys-
shallow up to sea level and only partly filled accom- tems tracts (TSTs) become less mud rich and more
modation space (Figure 24). An overall increasing ac- grain rich and are composed of distal and proximal
commodation space trend that is matched by sedi- middle-ramp cycles (Figures 8 [the foldout located
ment fill at the beginning of a cycle set resulted in in the back of this publication], 10, 24). Highstand
shallowing to near sea level, resulting in the last few systems tracts are thicker, grain rich, and cross-bedded
cycles being deposited in higher energy within fair- as they shallow upward into fair-weather-wave base
weather-wave base. In the lower part of the reservoir, and develop a ramp-crest shoal (Figure 8, the foldout
120 / Lindsay et al.

located in the back of this publication). Two paleosol The remaining 28 cycles, incorporated into 7 cycle sets
horizons (described below) have been identified in making up 2 HFSs and part of a third, are composed
the lower part of the reservoir (zone 3), with no pa- of proximal middle-ramp cycles, with only one distal
leosols identified in the upper part of the reservoir middle-ramp cycle and no outer ramp cycles (Figure 24).
(zones 1 and 2) (Figure 23). In the ramp-crest shoal, The lower composite sequence is the upper part of
the HSTs become thick and dominated deposition in the Jubaila Formation (6945.75-ft [2117.06-m] core
the upper part of the reservoir (Figure 8, the foldout depth to below the base of the core — lower half of
is located in the back of this publication). zone 2b and all of zone 3) (Figures 8 [the foldout lo-
Base-level fall led to the deposition of several cated in the back of this publication] and 24). This
seaward-stepping ramp-crest shoal complexes with composite sequence is referred to as the Jubaila com-
low-angle clinoform geometries. Downlapping clino- posite sequence. The top of the composite sequence was
form progradation was both northward and south- briefly subaerially exposed and contains uncommon
ward from an area in the vicinity of the Shedgum meniscus and flowstone (gravity) cements. Two HFSs
Arab-D core. Perhaps, several HFSs prograded and in other cored wells in the southern part of Ghawar
downlapped to fill accommodation space through- shallowed upward and developed short-lived paleosols.
out the length of Ghawar field. One paleosol is a Psamment paleosol, with plant root-
lets that were in a shoreface to localized island setting
Composite Sequences and the other deeper in the section contained man-
Two composite sequences have been identified. The grovelike rootlets in a paleosol at the zone 3a–3b contact
upper composite sequence forms the Arab-D Member (paleosols were identified by G. Retallack, 2003, 2004,
of the Arab Formation and is herein referred to as the personal communication). The lower half of the Jubaila
Arab-D composite sequence and covers a vertical in- composite sequence also contains a minor amount
terval from 6809.2 to 6945.75 ft (2075.4 to 2117.06 m) of windblown quartz silt (<1%) discussed below.
(zone 1, zone 2a, and the upper half of zone 2b) The Jubaila composite sequence in the example
(Figures 8 [the foldout located in the back of this core consists of 65 cycles, 15 cycle sets, and 4 HFSs
publication] and 24). The top of the composite se- (Figure 24). The ramp-crest shoal is composed of only
quence is located at the top of the Arab-D reservoir, three cycles within one cycle set in the uppermost
where the contact between the uppermost Arab-D HFS (Figures 8 [the foldout located in the back of this
carbonate and the overlying Arab C-D evaporite forms publication] and 24). No inner-ramp lagoon and in-
the sequence boundary. Along the composite sequence tertidal island setting was deposited in this vicinity.
boundary, uncommon, small, shallow, karst-generated However, sediments containing the shallow-marine
sink holes (not present in this well) filled with a few feet benthonic foraminifera Trocholina alpina (Figures 15, 16)
of collapse breccia are present. These can be seen in were reworked from a nearby lagoon and mixed into
core, in outcrop, and on FMI Fullbore Formation Mi- the ramp crest at this level. Outer-ramp and distal
croImager logs (Figure 13). In outcrop, small teepee middle-ramp depositional settings represent most of
structures characterize the top of the Arab-D compos- the deposition, with 55 cycles in 10 cycle sets that
ite sequence boundary in Wadi Hilwah (Figure 19). are in 2 HFSs (Figures 8 [the foldout located in the
The Arab-D composite sequence in the Shedgum back of this publication] and 24). The proximal mid-
Arab-D core consists of 55 cycles (parasequences), 13 cycle dle ramp is represented by 12 cycles and 3 cycle sets
sets (parasequence sets), and 5 HFSs (Figures 8 [the in the uppermost HFS (Figures 8 [the foldout located
foldout located in the back of this publication] and 24). in the back of this publication] and 24).
If the minimal amount of time for each cycle to be Meyer et al. (1996, 2000) and Al-Dhubeeb (2005)
deposited is 20,000 yr (assuming association with used biostratigraphy to correlate the contact between
20,000-yr astronomical precession cycle), then the the Arab-D Member of the Arab Formation and the
Arab-D Member may represent approximately 1.1 m.y. underlying Jubaila Formation in outcrop sections in
of time. Most of the Arab-D Member deposition was the Tuwaiq escarpment into the subsurface through
in the ramp-crest shoal, which is composed of approxi- the Khurais and Ghawar fields. Hughes (2004) de-
mately 22 cycles, 5 cycle sets (parasequence sets), and fined nine paleonvironmentally influenced biozones
2 HFSs, with a hardground separating the ramp crest (D1 to D9) in Ghawar field (Figure 25). Al-Dhubeeb
from the inner ramp (Figure 24). The inner ramp con- (2005) has demonstrated that three of those biozones
sists of five cycles that compose one cycle set that (D3 to D5) are not represented in Khurais field, and
forms a 9-ft (2.7-m)-thick, potential, HFS (Figure 24). that three slightly younger zones (D5 to D7) are not
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 121

FIGURE 25. Biostratigraphic


zonation of the Arab-D Mem-
ber and upper Jubaila Forma-
tion from outcrops in the
Tuwaiq escarpment into the
subsurface through Khurais
and Ghawar fields. Three
zones belonging to the Arab-D
Member are missing in out-
crop. In Khurais field, three
zones are also missing (upper
and middle in the Arab-D and
the lower from uppermost
Jubaila). In Ghawar field, all
zonations are present. The
distance from the Tuwaiq es-
carpment to Ghawar field is
300 km (186 mi). From
A. G. Al-Dhubeeb (2005, per-
sonal communication).

overlying Arab-D Member


(Steineke et al., 1958). In the
subsurface in Ghawar field,
thin sections from cored
represented in the Tuwaiq escarpment outcrop west wells contain quartz silt in the upper part of zone 2b
of Riyadh (Figure 25). In outcrop, the Arab-D Member (Figure 8, the foldout located in the back of this
is composed of inner-ramp lagoonal deposits of thin, publication). The quartz silt is interpreted to have
platy, mud-rich beds, which rest upon reworked prox- been reworked by eolian processes off the subaerially
imal middle-ramp stromatoporoid and coral rudstone exposed western shelf (Tuwaiq escarpment area) Ju-
and floatstone debris flows of the Jubaila Formation baila composite sequence boundary and blown into
(Figure 19) (Meyer, 2000). In outcrop, the composite the basin (Figure 26). In the Arabian intrashelf basin,
sequence boundary separating the overlying Arab-D the ultimate resting site for the windblown silt is
Member and the underlying Jubaila Formation is lo- within the transgressive systems tract (TST) of the
cated at a datum where stromatoporoids and corals first HFS of the Arab-D composite sequence. This HFS
disappear upsection and are overlain by thin, platy, is interpreted to onlap onto the Jubaila composite
and muddy inner-ramp lagoon beds (Figure 19) (Meyer sequence boundary regionally, but did not toplap it
et al., 1996; D. Vaslet, 2002, personal communica- (Figures 8 [the foldout located in the back of this
tion). In Ghawar field, with the complete biostrati- publication] and 27). The second HFS of the Arab-D
graphic succession and complete inner-ramp to outer- composite sequence only received windblown silt in
ramp succession present, identifying the composite the lower part of the transgressive systems tract (TST)
sequence boundary is more difficult. However, the (Figure 8, the foldout located in the back of this
same general criteria, upward decrease in the abun- publication). This second HFS is interpreted to have
dance and consistent presence of stromatoporoids also initially onlapped onto the Jubaila composite
and corals, coupled with an increase in abundance of sequence in the vicinity of the present-day Tuwaiq
the branched stromatoporoid Cladocoropsis, can be escarpment and, during maximum flooding (MFS),
used to identify this boundary. toplapped the composite sequence (Figure 27). It has
Although the top of the Jubaila composite se- been estimated that approximately 47 ft (14 m) of
quence is difficult to identify, a few additional pieces topographic relief existed between the subject well in
of information have helped identify the composite the Arabian intrashelf basin and outcrops in the
sequence boundary. First, at one outcrop locality in Tuwaiq escarpment (Figure 27). The topographic re-
the Tuwaiq escarpment, 1 m (3 ft) of sandstone was lief implied by this onlap-toplap series of events is in-
found resting upon the composite sequence bound- terpreted to have been 14 m (47 ft) because the first
ary that separates the Jubaila Formation from the two HFSs of the Arab-D Member (some 14 m [47 ft]
122 / Lindsay et al.

FIGURE 26. Windblown quartz


and dolomite silt set in a do-
lomitized matrix that contains
intercrystal and moldic porosi-
ty. Quartz silt was blown off
the subaerially exposed Jubaila
Formation composite sequence
boundary (Tuwaiq escarpment)
into the Arabian intrashelf ba-
sin. Black crystals near the top
of the photomicrograph are
pyrite cubes. Photomicrograph
width is 0.55 mm (6903.7 ft),
Shedgum Arab-D core.

highstand conditions of the


Jubaila composite sequence
as overall accommodation
space was lost. Second, Powers
et al. (1966) described a mud-
rich interval in outcrop that
could be regionally mapped
throughout the subsurface.
thick in the Shedgum Arab-D core) onlapped and even- This interval may represent maximum flooding
tually toplapped onto the Jubaila composite sequence (MFS) of the Arabian intrashelf basin during Arab-D
boundary in the Tuwaiq escarpment (Figure 27). Member deposition and may be the dolomitized in-
Based on this interpretation, the onlapping, HFSs terval separating zones 2a and 2b or, more likely, the
imply that the western shelf in the Tuwaiq escarp- next lower, somewhat persistent, dolomitized in-
ment was subaerially exposed. The interpretation of terval (upper zone 2b) (Figure 8, the foldout located
subaerial exposure is further supported by several fea- in the back of this publication). Dolomitization in
tures. First, deep incisions have been found in the Ghawar is generally in mud-rich transgressive sys-
uppermost Jubaila shelf margin in an outcrop in the tems tracts (TSTs) (Figure 8, the foldout located in
vicinity of the Diplomatic Quarter (DQ) on the south- the back of this publication) (although significant
west side of Riyadh. The last and deepest incision dolomitization locally occurs in grainy intervals,
forms the top of the Jubaila Formation and the top of most notably in the CLADO facies, where disso-
the Jubaila composite sequence (Figure 28). These lution of the Cladocoropsis colonies may produce
observations could easily have been produced under superpermeability).

FIGURE 27. Conceptual model of basal Arab-D Member onlap and toplap onto the Jubaila composite sequence
boundary, using the Shedgum Arab-D core (Ghawar field) in the Arabian intrashelf basin compared to outcrops in the
Tuwaiq escarpment, west of Riyadh, a distance of 310 km (189 mi).
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 123

FIGURE 28. Photographs of


the DQ section along the Makkah
Highway, west side of Riyadh,
where the Jubaila composite se-
quence boundary, three quarters
up the road cut, has been
downcut into by erosion and
channelized. Storms have re-
worked corals and stromatopo-
roids along the boundary and
into the channels. About 3 – 5 m
(10 – 15 ft) of Arab-D overlies
the Jubaila at this section. Ver-
tical height of the bottom pho-
to is 4.6 m (15 ft).

thickness change has been


alternatively interpreted as a
facies change from carbonate
In outcrop, the upper part of the Jubaila Formation to evaporite at the top of the Arab-D reservoir (Mitchell
contains a rich assemblage of reworked stromato- et al., 1988) or as a transition from a paleotopographic
poroids and corals with only a few whole pieces of high in northern Ghawar into an intrashelf basin to the
Cladocoropsis present (Meyer, 2000). In the subsurface south (Handford et al., 2002).
in Ghawar field, a rich assemblage of stromatoporoids In addition, smaller scale heterogeneity occurs
and corals is also present, whose upsection give way to internally within the Arab-D reservoir and varies in
Cladocoropsis bank assemblages that are not repre- a systematic way both temporally and geographically
sented in outcrop (Figure 8, the foldout located in the in the field. Lateral variability is least pronounced
back of this publication). Stromatoporoid and coral in the lower Arab-D reservoir, where meter-scale,
biostromes and mounds are most common in the up- coarsening-upward cycles consist of transgressive mi-
per Jubaila Formation, whereas Cladocoropsis banks critic wackestone and mudstone overlain by storm-
are most common in the lower part of the Arab-D transported, bivalve-coated grain-intraclast (BCGI)
Member. Exclusion of extensive Cladocoropsis in the rudstone and floatstone and occasionally by stroma-
Arab-D outcrop can be also explained by unsuitable, toporoid-red algae-coral (SRAC) deposits. They repre-
possibly too shallow environmental conditions, such sent either episodic storm deposition or cyclic varia-
as inner-ramp deposition. tion in sea level. The facies contained therein appear
In the subsurface at Ghawar field, a combination to be relatively continuous and create a highly strati-
of information can be used to identify the top of the fied reservoir interval (Mitchell et al., 1988) that are
Jubaila composite sequence boundary: (1) the strat- interpreted to very gently downlap. In contrast, the
igraphic level above which stromatoporoids and cor- middle and upper Arab-D reservoir is characterized
als are less abundant; (2) the stratigraphic level above by a much greater diversity of facies that tend to
which Cladocoropsis is more abundant; (3) the interval lack lateral continuity. Locally, thick accumulations
through which windblown quartz silt is present in the of stromatoporoid-red algal-coral rudstone and float-
overlying onlapping, HFSs; and (4) the stratigraphic stone with an internal matrix of mud-rich to mud-lean
level above which overlying transgressive mud-rich packstone and grainstone are present and have been
intervals have been dolomitized. interpreted as local carbonate buildups or mounds
(Mitchell et al., 1988; Handford et al., 2002) or as a
ARAB-D LATERAL FACIES VARIATIONS stromatoporoid bank (Hughes, 1996; Meyer et al.,
1996). Skeletal-oolitic (SO) grain-rich shoals cap the
Although impossible to interpret from a single core upper Arab-D interval and are thought to have a
from the Shedgum area, significant lateral variability shingled geometry, probably derived from the south-
in the distribution of lithofacies does occur in the field ward progradation in the southern 80% of Ghawar
(Figure 29). Overall, a general pattern of thinning of the and northward progradation in the northern 20%.
carbonate section and thickening of overlying evapo- Highstand systems tracts (HSTs) contain the great-
rite occurs from north to south across Ghawar. This est lateral variability and exhibit potential downlap
124 / Lindsay et al.

surfaces, whereas transgressive systems tracts (TSTs)


contain the best and most extensive lateral continu-
ity. Variations in continuity and development of
downlapping shingled or clinoform surfaces may be
related to relative sea level rise and fall that are not
necessarily representative of the Late Jurassic green-
house setting. Instead, the Late Jurassic appears to be
intermediate between icehouse and greenhouse con-
ditions (J. F. Read, 2004, personal communication). If
truly greenhouse, tidal-flat caps should be common,

FIGURE 29. South to north cross section of Arab-D reservoir lithofacies. Modified from Mitchell et al. (1988) and Cantrell (2004).
and if icehouse, numerous exposure surfaces should
be common (Read, 2004), although perhaps obscured
if little meteoric water was available during subaerial
exposure. What is described with respect to the Arab-D
reservoir is considered to represent intermediate or
transitional conditions between the greenhouse and
icehouse extremes.

DIAGENESIS

Diagenesis has modified the original sedimentary


textures and fabrics of the Arab-D reservoir and is a
significant factor in shaping the final reservoir quality
of these rocks. Major diagenetic processes active in
the Arab-D reservoir include dolomitization, dissolu-
tion and formation of microporosity, cementation,
and compaction (Mitchell et al., 1988). The most sig-
nificant of these processes are dolomitization, dis-
solution, and microporosity formation because they
created the most variation in reservoir quality.

Dolomitization
Dolomitization has long been recognized as a sig-
nificant component of the Arab-D reservoir in Ghawar
field (Powers, 1962; Mitchell et al., 1988; Meyer et al.,
1996) and contributes to vertical and lateral hetero-
geneity to the reservoir in terms of textural, miner-
alogical, and pore-type alteration (Cantrell et al., 2001,
2004; Swart et al., 2005). At least three dolomite
phases exist in the Arab-D, all of which are petro-
graphically, geochemically, and stratigraphically dis-
tinguishable (Table 1).
Fabric-preserving dolomite (Cantrell, 2004) con-
sists of very fine crystals of dolomite (10 – 50 mm), in
which the original limestone fabric is well preserved.
Fabric-preserving dolomite typically occurs in the up-
permost Arab D (zone 1) as thin, sheetlike, or strati-
form layers that are always intimately associated with
overlying laminated and bedded anhydrite. Reservoir
quality is generally fair to poor, with porosity typi-
cally less than 10% and permeability ranging from a
few millidarcys to a few tens of millidarcys. Because of
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 125

Table 1. Summary of Arab-D reservoir dolomite types.*


Fabric-preserving Dolomite NFP Dolomite Baroque

Style of dolomitization Fabric-preserving Nonfabric preserving, Obliterative, with saddle-


typically obliterates all shaped crystals that display
traces of the original undulose extinction
limestone fabric
Petrography Cloudy crystals, 10 – 50 mm Crystals may be clear or Crystals are typically
in size internally zoned, 50 to somewhat cloudy,
more than 150 mm in size 100 – 700 mm in size
Cathodoluminescence Nonluminescent Nonluminescent Patchy luminescence
Blue-light flourescence Flouresces strongly Flouresces weakly Flouresces weakly
Distribution Stratiform Stratiform Nonstratiform
Average d18O ratio 2.08% PDB 2.66% PDB 7.37% PDB
Average 87Sr/86Sr ratio 0.70681 0.70682 0.70712
Time formed Very early, almost Early Later
penecontemporaneously(?)
Water compositions Hypersaline Hypersaline Hot, mineral-laden fluids
from depth

*From Cantrell (2004).

its relatively heavy oxygen isotope values, intimate may exceed 20%, and permeability may approach sev-
association with the overlying anhydrite, and fabric- eral darcys. NFP dolomite contains heavy oxygen iso-
preserving nature, fabric-preserving dolomite is in- tope values (Table 1). Because of its general geochemical
terpreted to have formed early in the diagenetic his- similarity to fabric-preserving dolomites, it is inter-
tory of the sediment by dense, highly evaporated, preted to have formed from hypersaline fluids derived
magnesium-rich brines percolating down into the from the overlying salina anhydrite. In a few uncom-
underlying sediment from the overlying salina. The mon examples, some NFP dolomite contains light oxy-
early dolomitization interpretation is in agreement gen isotope values, has good reservoir quality (similar
with many previous studies of fabric-preserving or to above), and is thought to represent a transitional
mimetic dolomites (Dawans and Swart, 1988; Pleydell form, with the third dolomite type known as saddle or
et al., 1990; Land, 1991; Kimbell, 1993), and it is gen- baroque dolomite. Strontium isotopic ratios (Table 1)
erally recognized that fabric-preserving texture will suggest that both the fabric-preserving and NFP dolo-
not form if the precursor limestone has already sta- mite formed early or shortly after deposition of the
bilized to low-Mg calcite (Sibley, 1982, 1991). It should original sediment or during early burial (Cantrell, 2004).
be noted that there is dolomite lower in the section Saddle or baroque dolomite is a coarse-crystal do-
(zones 3a and 3b) where some aspects of the original lomite with saddle-shaped crystals displaying undu-
sediment may be interpreted (e.g., grain size and shape, lose extinction in thin section. It is uncommon in the
cross-bed inclination, etc.), but where fine details reservoir and appears to be limited to wells that con-
are not preserved. These dolomites are included with tain abnormally thick sections of dolomite. In ex-
the non-fabric-preserving dolomites (NFPs) discussed treme cases, saddle dolomite is vertically pervasive
below. and crosscuts stratigraphy. Geochemically, saddle do-
Non-fabric-preserving dolomite consists of medium- lomite is distinctive with high iron and very low
size crystals (50 to >150 mm), nonsaddle (nonba- (light) oxygen isotopic compositions (Table 1) and is
roque) dolomite, in which the original limestone fab- interpreted to have formed from high-temperature
ric has been lost, and in many cases, all fabric evidence fluids during burial diagenesis. Reservoir quality is
has been completely obliterated (Table 1) (Cantrell, fair to poor. The interpretation that saddle dolomite
2004). Typically, NFP dolomite occurs as stratiform formed relatively late in the diagenetic history of a
layers scattered throughout the reservoir. It gener- rock from high-temperature fluids during deeper
ally has very poor reservoir quality. Intercrystal or burial diagenesis is consistent with many previous
vuggy porosity is typically less than 10%, and perme- studies (Folk and Assereto, 1974; Radke and Mathis,
ability is less than 1 md. Locally, intercrystal porosity 1980) and is in agreement with the fluid-inclusion
126 / Lindsay et al.

FIGURE 30. Cross sections con-


trast the lateral persistence and
continuity of stratigraphic dolo-
mite (section A) with vertically
pervasive, nonstratigraphic dolo-
mite (section B). Modified from
Cantrell et al. (2004).

is more mud rich. Descending


brines sourced from the over-
lying C-D evaporite will per-
colate through the more grain-
rich HST and be dramatically
slowed to preferentially dolo-
mitize the TST.
In contrast, the more un-
common saddle dolomite and
some isotopically light forms
of NFP dolomite (more coarse-
ly crystalline than most NFP
dolomite) are nonstratiform in
nature, in that they typically
occur as abnormally thick do-
lomitized intervals that cross-
cut stratigraphic reservoir
zones (Figure 30B). Mapping
reveals that this type of dolo-
results from the Arab-D (mean homogenization tem- mite occurs as linear northeast–southwest-oriented
perature = 97.48C, Table 1) (Cantrell, 2004). fracture and fault trends across Ghawar (Figure 31).
In addition, quantification and mapping of the Dolomite along these trends typically makes up 40–
dolomite content of the Arab-D across Ghawar re- 60% (and locally as much as 100%) of the entire Arab-D
vealed that dolomite does not occur randomly or reservoir interval, whereas offtrend dolomite typically
uniformly across the field. Fabric-preserving and most makes up less than 20% of the reservoir. These fracture-
NFPs typically occur as bed-parallel, thin, lenticular and fault-related saddle dolomite trends appear to be
or sheetlike units that range in thickness from 0.3 very limited areally (1 – 2 km [0.6 – 1.2 mi] in width)
to 4.6 m (1 to 15 ft). These stratiform dolomite bod- but extend obliquely across the field for distances
ies commonly occur within fairly well-defined strati- greater than 60 km (37 mi). Because of their narrow
graphic intervals in the Arab-D (Figure 30A). As such, width (which is on the order of one to two well spac-
these dolomites may vertically stratify the reservoir ings), these trends are interpreted to be vertical in
and result in local impediments to vertical flow. Note, orientation and to potentially create horizontally
however, that thin intervals (locally as little as 0.3 m separated compartments in the reservoir and would
[1 ft] thick) of high-porosity and high-permeability generally act as vertically oriented, elongate, low-
NFP dolomite may contribute a high proportion of permeability barriers that would drastically effect
the flow from an individual well, creating a high-flow horizontal flow in the reservoir. When porous, these
or super-K zone. trends of saddle (baroque) dolomite can contain high
Detailed core descriptions throughout the field permeability and act as sites of high fluid flow, or
have begun to show a preference for dolomitization super-K, in the reservoir (the occurrence of which is
to be more pervasive in the transgressive systems tract currently unpredictable).
(TST) and at the maximum flooding surface (MFS), A model of dolomitization (Figure 32) is proposed
with less dolomite associated with the HST. The ap- to explain the occurrence of both stratiform and
parent reason for this phenomenon is that the TST nonstratiform types of dolomite. Stratiform dolomite
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 127

originated via several mech-


anisms that occurred early in
the diagenetic history of the
sediment and are related to
downward percolation of hy-
persaline brines as the over-
lying C-D evaporite was be-
ing deposited (Figure 32). In
contrast, nonstratiform, pre-
dominantly saddle dolomite
formed from hot, mineral-
laden fluids that ascended
upsection and emplaced in-
to the reservoir from depth
along a series of faults and/
or fractures and preferen-
tially dolomitized the sec-
tion adjacent to faults and
fractures (Figure 32). In some
instances, the fractures were
first solution widened prior
to emplacement of the sad-
dle dolomite. Saddle dolo-
mite has been observed in
faulted blocks of Jubaila –
Arab-D carbonates along the
north flank of Wadi Nisah,
which provides some evi-
dence for fault-related do-
lomitization of the saddle
dolomite.

Implications
With the exception of
saddle dolomite, which is
thought to be associated with
the movement of fluids
along fracture and fault zones
during later burial, the oc-
currence of dolomite in the
Arab-D reservoir reflects spe-
cific physical and chemical
conditions that occurred dur-
ing deposition to early buri-
al of the Arab-D. As a result,
placing these dolomites in
the framework of the overall
depositional and stratigraph-
FIGURE 31. The left is a map of percent dolomite in the Arab-D reservoir, Ghawar field. ic architecture of the Arab-D
The right is a map of percent dolomite in zone 2b, Arab-D reservoir, Ghawar field. is essential to understand the
Modified from Cantrell et al. (2004). proper temporal and spatial
128 / Lindsay et al.

and nearest to the source


of descending brines, where
brine concentrations were
highest and reaction rates
were rapid. In addition, al-
though NFP dolomite occurs
in both Arab-D composite se-
quences, its distribution is
not uniform. No significant
dolomitization of any kind
occurs in the skeletal-oolitic
(SO) grainstone-dominated
upper Arab-D (zone 2A), but
NFP dolomite is common in
the middle part and the up-
per part of the lower Arab-D
reservoir (zones 2B and 3A).
This NFP dolomite occurs in
association with stromato-
poroid-red algal-coral (SRAC)
and Cladocoropsis (CLADO)
rudstone and floatstone (pack-
stone and grainstone matrix)
in the basal part zone 2B and
with mudstone and wacke-
stone interbedded with bi-
valve-coated grain-intraclast
(BCGI) rudstone and float-
stone (grainstone, mud-lean
and mud-rich packstone ma-
trix) in zone 3A. Little or no
dolomite occurs in zone 3B
or below in the northern
half of Ghawar, but some
zone 3B mudstone is dolomi-
tized in the southern half.
Although fabric-preserving
and NFP dolomites are inter-
preted to have similar ori-
gins, their different dolomite
FIGURE 32. Dolomitization model for the Arab-D reservoir in Ghawar field. Modified fabrics reflect their different
from Cantrell et al. (2004). stratigraphic contexts and
potentially slightly different
times of formation. Fabric-
context in which dolostones formed and to provide a preserving dolomite formed early (penecontempora-
predictive framework for the occurrence of similar do- neously), prior to extensive stabilization of the sedi-
lostones in other intervals. ment to low-Mg calcite, whereas NFP formed later
It is clear that the partitioning of dolomite types (but still early, as suggested by the Sr isotope data),
occurs in the Arab-D, with NFP dolomite occurring in subsequent to mineralogical stabilization of the sedi-
both composite sequences of the Arab-D carbonate, ment, but after initial compaction.
whereas fabric-preserving dolomite occurs only in Overall, the vertical distribution of dolomite re-
zone 1, at the top of the upper composite sequence flects both the vertical succession of the Arab-D as well
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 129

FIGURE 33. Arab-D reservoir


rock classification for Ghawar
field. Modified from Cantrell
and Hagerty (2003).

dolomitizing fluids. The upper


composite sequence bound-
ary at the top of the Arab-D
served to help partition fabric-
preserving from NFP dolo-
mite; mud-rich transgressive
systems tracts in HFSs also
help localize dolomite. The
as the hydrodynamics of the top-down dolomitiza- presence of certain sedimentary fabrics, such as poorly
tion mechanism proposed for both types of dolomite. sorted and mud-prone rocks in these cycles, appears
Dense, hypersaline brines that migrated downward to act as the primary control over the occurrence of
from the overlying salina initially moved into zone 1 nonsaddle dolomite. Provided that these three con-
shortly after deposition and produced fabric-preserving trolling factors are present, it is possible to under-
dolomite in unstabilized sediment. As deposition of stand and predict patterns of dolomitization in the
salina evaporites continued, brines percolated down- Arab-D, as well as in other similar carbonate-evaporite
ward and moved into the older Arab-D reservoir com- cycles that formed elsewhere. Vertically oriented high-
posite sequences below. Initially, these fluids moved temperature saddle (baroque) dolomite, in fracture-
unrestricted through well-sorted skeletal-oolitic (SO) and fault-related structural trends, across the Arab-D
grainstones in the Arab-D (zone 2A) but slowed sig- reservoir have the potential to influence horizontal
nificantly as they passed through the highly heteroge- flow in the reservoir.
neous, poorly sorted, and more mud-prone sediment-
rock at the zone 2A – 2B contact and in zones 2B RESERVOIR ROCK CLASSIFICATION
and 3A. There, increased residence time of the fluids
resulted in increased probability of dolomitization An integrated petrographic and petrophysical
of these mud-prone sediments. In addition, the finer study of Arab-D carbonates in the Ghawar field
grain size in the muddier parts (TST) of these sequences has provided a new carbonate reservoir rock clas-
promoted dolomitization by providing a greater num- sification (Figure 33) (Cantrell and Hagerty, 2003).
ber of nucleation sites for dolomitization (Sibley and Each reservoir rock type recognized in this classi-
Gregg, 1987) than would the less mud-prone, more fication has a distinct pore network as defined by
grain-rich sediment-rock above. Ultimately, these porosity-permeability relationships and pore-throat
fluids continued moving downsection (into lower radii connectivity expressed as pore-throat radii size
zone 3A and zone 3B) and eventually equilibrated distribution and J-function curves. This classification
with the local host sediment-rock-water system and divides Arab-D carbonates into seven limestone and
were no longer capable of dolomitizing. As a result, four dolomite rock types (Figure 33). The amount of
the amount of dolomite generally decreases down- matrix (lime mud) and pore types are the primary de-
section below the middle Arab-D reservoir. fining parameters for limestones. Reservoir rock type
Dolomitization thus induced heterogeneity in a for dolomite is based on crystal texture. Note that
variety of ways that are generally predictable in the numerous pigeon holes in the classification scheme
Arab-D reservoir. In a stratigraphic sense, dolomite (Figure 33) have been left blank, indicating that those
occurrence is predictable based on the following criteria are of no significance in defining that par-
criteria: (1) proximity to a thick, laterally extensive ticular rock type.
evaporite unit (C-D evaporite), the deposition of which The seven limestone reservoir rock types are based
provided a source of downward-migrating hypersa- on the values of five petrographic parameters, ap-
line brines; (2) a major sequence boundary that sepa- plied in the following order (that is, starting from
rated the overlying evaporite from underlying car- the top of the table in Figure 33 and working toward
bonate; and (3) cyclic carbonates that contain poorly the bottom): (1) abundance of cement; (2) abun-
sorted, mud-prone sediments that are prone to retain dance of matrix (lime mud); (3) sorting (Krumbein and
130 / Lindsay et al.

FIGURE 34. Typical core-plug photo-


graph examples of the major Arab-D
reservoir limestone rock types. Divi-
sions at top of photographs are in
millimeters. Modified from Cantrell
(2004).

The four dolomite reservoir rock


types are classified according to do-
lomite crystal texture, although strati-
graphic position and abundance of
porosity can also be effective in their
classification. The four textures are
fabric preserving (Vfp), sucrosic (Vs),
intermediate (Vi) and mosaic (Vm).
As noted previously, the Vfp dolo-
mite is only found in zone 1 of the
Arab-D, where it is the major dolo-
mite type. Vs dolomite occurs in
dolomites with more than 12% po-
rosity; Vm is less than 5% porosity,
and Vi is between 5 and 12% porosity.
Vfp dolomites have pore systems sim-
ilar to their precursor limestone, but
the pore systems of the other dolo-
mite types are unique.

Implications
The distribution of these rock
types is not random in the reservoir
and reflects the stratigraphy and
Pettijohn, 1988); (4) dominant pore type; and (5) size of vertical succession of rock types previously described
the largest molds. Of the five, the amount of matrix is in the Arab-D reservoir (Figure 8, the foldout located
the most important parameter, although the amount in the back of this publication). As noted previously,
of cement is the first criterion for subdividing the zone 2A is dominated by relatively uniform, well-
limestone rock types (Figure 33). In general terms, six sorted, skeletal-oolitic grainstone that falls into res-
of these seven types fall into two broad families, A ervoir rock type IB, whereas more heterogeneous
and B, each of which can then be subdivided into Cladocoropsis (CLADO) and stromatoporoid-red algal-
three members (types I, II, and III) according to their coral (SRAC) rudstone and floatstone that contain a
matrix content. The first family, A, is a fairly coarse- matrix of packstone and grainstone occur in zone 2B
grained, poorly sorted rock with relatively large molds. and fall into reservoir rock type IA or IIA. Zone 3A
The second family, B, is a generally fine- to medium- tends to be dominated by heterogeneous and rela-
grained, well-sorted rock with few or small molds. The tively mud-prone, bivalve-coated grain-intraclast
seventh rock type contains more than 10% cement, (BCGI) rudstone and floatstone that contain a ma-
which modifies the pore-size distribution enough to trix of packstone that tend to fall in reservoir rock
warrant a separate reservoir rock type. Each of the type IIB, whereas zone 3B typically contains abun-
reservoir rock types exhibits a distinctive pore-throat dant micritic wackestone to mudstone that would
size distribution and, in turn, a Leverett J-function. be classified in reservoir rock type IIIB. The occurrence
The seven types are also characterized by distinctive of these reservoir rock types is thus predictable in
porosity-permeability relationships (Figures 34, 35). the reservoir.
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 131

FIGURE 35. Porosity-perme-


ability crossplots of individual
limestone reservoir rock types.
Outliers are labeled according
to their nonuniformity or tex-
tural characteristics (Bu = bur-
rows; Cg = coarse grained;
Fr = microfractures; Tx = tex-
tural variations). Green are for
zones 1 and 2, and red are
for zones 3 and 4. Modified
from Cantrell (2004).

RESERVOIR
CHARACTERIZATION
USING IMAGE LOGS

Background
Understanding the hetero-
geneity of porosity and per-
meability is critical to evalu-
ating production behavior of
Ghawar field (Figure 37). De-
velopment scenarios have
been implemented that em-
ploy long-radius, multilater-
This new classification scheme is significant, in al horizontal wellbores, termed ‘‘maximum reservoir
that it focuses on observed heterogeneities in the contact’’ (MRC) wells, which target the uppermost
rocks and groups them according to impact on flow Arab-D reservoir while avoiding deeper potential high-
properties. Thus, this classification is quite different permeability zones. Because nearly all of these wells
from other, more widely used carbonate classifica- are not cored, few available data characterize reser-
tions. Most other classifications have relied on either voir quality in multilateral wells, with the exception
textural (Dunham, 1962; Powers, 1962; Lucia et al., of image and conventional logs.
2001) or lithofacies information (Mitchell et al., 1988; Several challenges exist to characterizing reservoir
Meyer and Price, 1993). Figure 36 compares two stan- porosity and permeability variations in the Ghawar
dard classifications with the scheme used in this study. field. First, in some instances, a complex secondary
Although collection of texture and lithofacies infor- diagenetic overprint (such as dolomitization) has altered
mation can provide the information required to de- the primary texture. Quantifying petrophysical prop-
velop the depositional and stratigraphic framework erties of the reservoir and correlation of wire-line-log
of the reservoir and its gross reservoir quality, it does signatures to depositional facies become challenging.
not provide sufficient information to define pore In these instances, porosity and permeability are con-
network types in the reservoir. These textural and trolled much more by diagenesis than by deposition-
lithofacies classifications only indirectly relate to al facies, although depositional facies influenced the
pore types and, in most instances, are unsuccessful degree of diagenetic overprint. Second, conventional
in subdividing the reservoir rocks into groupings that wire-line logs lack the fundamental vertical and azi-
are significant relative to their flow characteristics. muthal resolution to provide an adequate characteri-
The recognition and codification of the present clas- zation of the reservoir heterogeneity. Significant tex-
sification, producing reservoir rock groupings that tural and stratigraphic variations between adjacent
have significantly similar flow characteristics, in- thin layers hold the key to understanding the sequence
vites development of a coordinated scheme for the stratigraphy. Third, standard core plugs, taken at ver-
prediction of occurrence of these groupings in the tical intervals of 6 in. (15 cm) or more, may not be
reservoir. adequate to characterize the reservoir quality.
132 / Lindsay et al.

permeability prediction (Russell et al., 2002, 2004),


and extrapolation of image log-derived permeability
data into wells without image logs using sonic-derived
vuggy porosity measurements (Xu et al., 2006).
Porosity and permeability in vuggy carbonate res-
ervoirs is quantified through an optimized workflow
of borehole image and conventional log processing
that is calibrated to core data. Permeability from cores,
which commonly has a poor correlation with con-
ventional log-derived permeability, is found to have
an exponential correlation to the vug porosity com-
ponent of the total porosity computed from borehole
image porosity analysis.

Results
Primary depositional texture exerts a strong in-
fluence on reservoir quality in the Arab-D because
of the effect of lime mud on permeability. Mud-
dominated textures, such as mudstone and wacke-
stone, have generally low porosity and permeability,
FIGURE 36. Distribution of limestone samples by standard with the exception of diagenetically altered dolo-
carbonate rock classifications and reservoir rock types.
stone that has relatively high permeability in in-
Dunham textures: GRN = grainstone; MLP = mud-lean
packstone; PCK = packstone; WCK = wackestone; MUD = termediate porosity. Packstone and floatstone have
mudstone. Depositional facies: SO = skeletal-oolitic; MIC = higher porosity and permeability, and grain-supported
micritic; BCGI = bivalve-coated grain-intraclast; SRAC = textures, such as grainstone and rudstone, have the
stromatoporoid-red algae-coral; CLAD = Cladocoropsis. best reservoir quality. Traditional log-based porosity-
Modified from Cantrell (2004).
permeability transforms do not adequately predict
permeability in different textures with similar total
porosity values because the vuggy porosity com-
Methodology ponent of the total porosity is not measured on con-
Carbonate rocks are characterized by complex po- ventional wire-line logs, which provide porosity
rosity systems (Lucia, 1995; Kazatchenko and Mousa- values averaged more than 0.6 – 1.2 m (2 – 4 ft) ver-
tov, 2002). Small-scale heterogeneity of porosity and tically. Stromatoporoid floatstone and rudstone
permeability can only be resolved in wellbore logs with contain significant vuggy porosity in image logs,
image log data (Newberry et al., 1996). Microresistivity and the clearly multimodal porosity spectrum mea-
image logs have ideal, small-scale vertical resolution as sured azimuthally in image logs contributes to anom-
well as azimuthal borehole coverage for detailed res- alously high permeability that matches production
ervoir heterogeneity analysis. Core data are necessary data. Sucrosic dolostone, commonly characterized by
to calibrate all the log results, especially in heteroge- large intercrystalline vugs, accounts for extremely
neous rocks. In wells without borehole images, open- high flow from very thin intervals (<0.3 m [<1 ft]
hole lithology logs are required to determine reservoir thick) because of permeability that is measured from
rock types for heterogeneity analysis. image log analyses at two to three orders of magnitude
Image logs that detect relative microresistivity higher than conventional log transforms. This dramatic
changes (e.g., pores invaded by conductive drilling increase of permeability with vug porosity partition-
mud that contrast with resistive rock matrix) in the ing without total porosity change can result in the
wellbore may aid in the identification of rock tex- inverse correlation between porosity and permeabil-
ture. Therefore, image logs have been used for various ity (Russell et al., 2002). Conventional wire-line-log
quantitative analyses in carbonate reservoir charac- porosity-permeability transforms can neither predict
terization, such as partitioning of matrix and vuggy nor resolve such heterogeneity of permeability. Fi-
porosity through borehole porosity image analysis nally, well-sorted, highly permeable oolitic, peloidal
(Newberry et al., 1996; Chitale et al., 2004), vug con- grainstones of the upper Arab-D are commonly very
nectivity through image texture analysis leading to thin (less than 0.3 m [1 ft]) and account for high flow
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 133

FIGURE 37. Cross-bedded peloid grainstone in zone 2A. Image logs resolve thin beds of grainstone (black lines on left)
and commonly show hints of cross-bedding (as indicated by the enlarged view of the red outlined interval shown
below).

rates (Figure 37). These grainstones are well sorted base) to a depth of approximately 30 m (100 ft)
and exhibit a high degree of interconnectedness of water depth characterized by micritic cycles with
intergranular pores. Image-derived permeability data Thalassinoides burrows. Updip, the distal middle
strongly correlate with core-derived permeability in ramp is characterized by micritic cycles capped
these facies. Image logs clearly resolve the thin grain- by Thalassinoides firmgrounds with overlying float-
stone layers, and calibration of image facies with high- stone to rudstone of dominantly bivalve-coated
resolution core description data provides a method- grain-intraclast sediment. The proximal middle
ology for extrapolating grainstone layers into uncored ramp is characterized by domed and encrusting
wells using image logs. A methodology is shown for stromatoporoid and coral biostromes and mounds
predicting flow behavior from image logs alone using with delicately branched Cladocoropsis stromato-
image-derived permeability. After calibration in verti- poroids in sheltered areas. The ramp-crest shoal
cal cored wells, the classification of image textures and is characterized by skeletal-oolitic grainstone to
facies is the basis for estimating grainstone, stroma- packstone, and the inner-ramp lagoon is char-
toporoid, and dolostone facies, cyclicity, and their acterized by more micritic sediment with local
associated permeabilities in image logs from hori- intertidal islands.
zontal and uncored wells. 2) The reservoir is made up of two composite se-
quences. The lower composite sequence is the
CONCLUSIONS upper part of the Jubaila Formation. The se-
quence boundary occurs at the top of the ramp-
1) The Arab-D reservoir in the Ghawar field was de- crest shoal, which is overlain by middle-ramp
posited on a tropical carbonate ramp that ex- lithofacies. The upper composite sequence is the
tended from outer ramp (below fair-weather-wave D member of the Arab Formation, the sequence
134 / Lindsay et al.

boundary at its top marked by local karst-collapse of the manuscript. We thank Nassir Alnaji, David
breccias and flooded over by subaqueous evapo- Bacchus, and the cartography section of Saudi Aramco
rites interpreted as the TST of the following com- for their help with the creation of the digital and
posite sequence. scanned images.
3) Major diagenetic features of the Arab-D reservoir
include dolomitization, dissolution, cementation, REFERENCES CITED
and compaction. Dolomitization was mostly
non-fabric-preserving and stratiform, occurring Al-Dhubeeb, A. G., 2005, Biofacies as a tool for calibrat-
in more micritic TST sediments and probably ing the Jurassic Jubaila – Arab formational contact
from outcrop in Riyadh area to subsurface in eastern
formed by downward seepage of Mg-rich brines
Saudi Arabia: Master’s thesis, King Fahd University
during deposition of the overlying C-D anhydrite. of Petroleum and Minerals, Dhahran, Saudi Arabia,
Locally, saddle dolomite occurs in bodies that 219 p.
crosscut stratigraphy, perhaps resulting from Al-Hinai, K. G., A. E. Dabbagh, W. C. Gardner, M. Khan,
late fluids moving up local fracture zones. Dis- and S. Saner, 1997, Shuttle imaging radar views of
solution is widespread, especially in the grainier some geological features in the Arabian Peninsula:
parts of the reservoir, but the cause is less obvious GeoArabia, v. 2, p. 165 – 178.
Al-Husseini, M. I., 1997, Jurassic sequence stratigraphy of
given the evaporitic nature of the overlying hy-
the western and southern Arabian Gulf: GeoArabia,
drographic system.
v. 2, p. 361 – 382.
4) Porosity includes interparticle, moldic, intrapar- Al-Husseini, M. I., 2000, Origin of the Arabian plate struc-
ticle, and microporosity. Permeability is mostly tures: Amar collision and Najd rift: GeoArabia, v. 5,
dominated by interparticle pores, which are typi- no. 4, p. 527 – 542.
cally only lightly cemented. Although moldic pores Arkell, W. J., 1952, Jurassic ammonites from Jebel Tuwaiq,
are abundant, they typically only contribute sig- central Arabia: Royal Society of London Philosophical
nificantly to permeability in dolomite, where the Transactions, Series B, v. 236, p. 241 – 313.
Ayres, M. G., M. Bilal, R. W. Jones, L. W. Slentz, M. Tartir,
delicately branched stromatoporoid Cladocoropsis
and A. O. Wilson, 1982, Hydrocarbon habitat in main
has been leached and the matrix has been re- producing areas, Saudi Arabia: AAPG Bulletin, v. 66,
placed by tightly intergrown dolomite. Dolomite p. 1 – 9.
is porous to nonporous (tight), with porosity a Barger, T. C., 2000, Out in the blue: Vista, Selwa Press,
mixture of moldic, intercrystal, and intracrystal 320 p.
pores. Bates, B. S., 1973, Oscar for an oilfield: Saudi Aramco World,
5) Image logs, in conjunction with other logs, show v. 24, no. 6, p. 14 – 15.
significant promise for the identification of litho- Beydoun, Z. R., 1988, The Middle East: Regional geology
and petroleum resources: Beaconsfield, United King-
facies, especially coarse-grained lithofacies, in the
dom, Scientific Press, 292 p.
Arab-D reservoir. Image log data clearly resolve Beydoun, Z. R., 1991, Arabian plate hydrocarbon geology
thin layers of connected vugs in sucrosic dolo- and potential — A plate tectonic approach: AAPG
stone and stromatoporoid rudstone and float- Studies in Geology 33, 77 p.
stone that contribute to high permeability and Beydoun, Z. R., 1998, Arabian plate oil and gas: Why so
flow rates. Permeability from image log analyses rich and so prolific?: Episodes, v. 21, p. 74 – 81.
correlates strongly with core permeability, yet Cantrell, D. L., 2004, Carbonate heterogeneity during
global greenhouse time: Examples from the Jurassic of
image-derived permeabilities are two to three or-
the Middle East: Ph.D. thesis, University of Manchester,
ders of magnitude higher than conventional log
Manchester, England, United Kingdom, 368 p.
transforms in intervals with high vuggy poros- Cantrell, D. L., and R. M. Hagerty, 2003, Reservoir rock
ity. Predicted image-derived flow behavior cor- classification, Arab-D reservoir, Ghawar field, Saudi
relates strongly with actual flowmeters. Arabia: GeoArabia, v. 8, p. 435 – 462.
Cantrell, D. L., P. K. Swart, C. R. Handford, C. G St. C.
ACKNOWLEDGMENTS Kendall, and H. Westphall, 2001, Geology and produc-
tion significance of dolomite, Arab-D reservoir, Ghawar
field, Saudi Arabia: GeoArabia, v. 6, p. 45 – 60.
The authors thank the management of Saudi
Cantrell, D. L., P. K. Swart, and R. M. Hagerty, 2004,
Aramco for granting permission to publish this con- Genesis and characterization of dolomite, Arab-D res-
tribution. In particular, the authors thank Abdulla ervoir, Ghawar field, Saudi Arabia: GeoArabia, v. 9,
Al-Naim, Aboud Al-Afifi, Sa’id Al-Hajri, Adnan Al- p. 11 – 36.
Sharif, and George Grover for their support and review Carrigan, W. J., G. A. Cole, E. L. Colling, and P. J. Jones, 1994,
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 135

Geochemistry of the Upper Jurassic Tuwaiq Mountain Handford, C. R., D. L. Cantrell, and T. H. Keith, 2002,
and Hanifa formations petroleum source rocks of east- Regional facies relationships and sequence stratig-
ern Saudi Arabia: in B. J. Katz, ed., Petroleum source raphy of a super-giant reservoir (Arab-D Member),
rocks, casebooks in earth sciences series: Berlin, Springer- Saudi Arabia, in J. M. Armentrout, ed., Sequence strati-
Verlag, p. 67–87. graphic models for exploration and production: Evolv-
Champetier, Y., and E. Fourcade, 1966, A propos de ing methodology, emerging models and application
Cladocoropsis mirabilis Felix dans le Jurassique superieur histories, 22nd Annual Bob F. Perkins Research Confer-
de Sud-Est de l’Espagne: Estudios Geologicos, v. 22, ence: Gulf Coast Section SEPM, p. 539 – 564.
p. 101 – 111. Hughes, G. W., 1996, A new bioevent stratigraphy of late
Chitale, D. V., J. Quirein, T. Perkins, G. B. Lambert, and Jurassic Arab-D carbonates in Saudi Arabia: GeoArabia,
J. C. Cooper, 2004, Application of a new borehole v. 1, p. 417 – 434.
imager and technique to characterize secondary po- Hughes, G. W., 1997, The Great Pearl Bank barrier of the
rosity and net-to-gross in vugular and fractured car- Persian Gulf as a possible Shu’aiba analogue: Geo-
bonate reservoirs in Permian basin: 45th Society of Arabia, v. 2, p. 279 – 304.
Professional Well Log Analysts Annual Logging Sym- Hughes, G. W., 2004, Middle to Upper Jurassic Saudi
posium, Noordwijk, June 6 – 9, 2004. Arabian carbonate petroleum reservoirs: Biostratigra-
Dawans, J. M., and P . K. Swart, 1988, Textural and geochem- phy, micropaleontology and paleoenvironments: Geo-
ical alternations in late Cenozoic Bahamian dolomites: Arabia, v. 9, no. 3, p. 79– 114.
Sedimentology, v. 35, p. 385 – 404. Irwin, M. L., 1965, General theory of epeiric clear water
De Castro, P., 1991, On the cortical layer of Thaumato- sedimentation: AAPG Bulletin, v. 49, p. 445 – 459.
porellaceans (green algae) (abs.): 5th International James, N. P., 1983, Reef environment, in P. A. Scholle,
Symposium on Fossil Algae, Capri, April 7 – 12, p. 16 – D. G. Bebout, and C. H. Moore, eds., Carbonate de-
17. positional environments: AAPG Memoir 33, p. 345 –
De Castro, P., 2002, Thaumatoporella parvovesiculifera 440.
(Raineri): Typification, age and historical background Kazatchenko, E., and A. Mousatov, 2002, Primary and
(Senonian, Sorrento Peninsula, southern Italy): Bollet- secondary porosity estimation of carbonate formation
tino della Societa Paleotologica Italiana, v. 41, no. 2 – 3, using total porosity and the formation factor: Pre-
p. 121 – 129. sented at 2002 Society of Petroleum Engineers Annual
De Matos, J. E., and R. F. Hulstrand, 1995, Regional Technical Conference and Exhibition, San Antonio,
characteristics and depositional sequences of the Texas, SPE Paper 77787, 6 p.
Oxfordian and Kimmeridgian, Abu Dhabi, in M. I. Al- Keith, T. H., 2005, Finding the super-giant: The discovery
Husseini, ed., Middle East Geosciences Conference, of Ghawar field: GeoFrontier, Dhahran Geoscience
GEO’94: Bahrain, Gulf PetroLink, v. 1, p. 346 – 356. Society, v. 2, no. 1, p. 29 – 44.
Droste, H. H. J., 1990, Depositional cycles and source rock Kimbell, T. N., 1993, Sedimentology and diagenesis of late
development in an epeiric intra-platform basin, the Pleistocene fore-reef calcarenites, Barbados, West In-
Hanifia Formation of the Arabian Peninsula: Sedimen- dies: A geochemical and petrographic investigation of
tary Geology, v. 69, p. 281 – 296. mixing zone diagenesis: Ph.D. dissertation, University
Dunham, R. J., 1962, Classification of carbonate rocks of Texas, Dallas, 296 p.
according to depositional texture, in W. E. Ham, ed., Kinsman, D. J. J., 1964, Recent carbonate sedimentation
Classification of carbonate rocks: AAPG Memoir 1, near Abu Dhabi, Trucial Coast, Persian Gulf: Ph.D.
p. 108 – 121. thesis, University of London, London, 302 p.
Durham, L. S., 2005, Saudi Arabia’s Ghawar field: The ele- Konert, G., A. M. Al-Afifi, S. A. Al-Hajri, and H. J. Droste,
phant of all elephants: AAPG Explorer, January 2005, 2001, Paleozoic stratigraphy and hydrocarbon hab-
p. 4 and 7. itat of the Arabian plate: GeoArabia, v. 6, p. 407 –
Enay R., 1987, Le Jurassique d’Arabie Saoudite Centrale: 442.
Geobios Memoir Special 9, 314 p. Krumbein, W. C., and F. J. Pettijohn, 1988, Manual of
Folk, R. L., and R. Assereto, 1974, Giant aragonite rays and sedimentary petrography: SEPM Reprint Series 13,
baroque white dolomite in tepee-fillings, Triassic of 549 p.
Lombardy, Italy (abs.): SEPM Annual Meeting Pro- Land, L. S., 1991, Dolomitization of the Hope Gate
grams with Abstracts, v. 164, p. 34 – 35. Formation (north Jamaica) by seawater: Reassessment
Glennie, K. W., 2000, Cretaceous tectonic evolution of of mixing-zone dolomite, in H. P. Taylor, J. R. O’Neill,
Arabia’s eastern plate margin: A tale of two oceans: in and I. R. Kaplan, eds., Stable isotope geochemistry: A
A. S. Alsharhan and R. W. Scott, eds., Middle East tribute to Samuel Epstein: Geochemical Society
models of Jurassic/Cretaceous carbonate systems: (London) Special Publication 3, p. 121 – 133.
SEPM Special Publication 69, p. 9 – 20. Leinfelder, R. R., F. Schlagintweit, W. Werner, O. Ebli,
Hancock, P. L., A. Al-Kadhi, and N. A. Sha’at, 1984, Re- M. Nose, D. U. Schmid, and G. W. Hughes, 2005,
gional joint sets in the Arabian platform as indicators Significance of stromatoporoids in Jurassic reefs and
of intraplate processes: Tectonics, v. 3, p. 27 – 43. carbonate platforms — Concepts and implications:
136 / Lindsay et al.

Facies: International Journal of Paleontology, Sedi- sis of carbonate dual porosity systems from borehole
mentology and Geology, v. 51, p. 287 – 325. electrical images: Society of Petroleum Engineers
LeNindre, Y.-M., J. Manivit, H. Manivit, and D. Vaslet, Paper 35158, 7 p.
1990, Stratigraphie sequentielle du Jurassique et du Nicholson, P. G., 2000, Compressional, fault-related folds
Cretace en Arabie Saoudite: Bulletin de la Societe and Saudi Arabia’s major hydrocarbon fields (abs.):
Geologique de France, series 8, 6, p. 1025 – 1034. GeoArabia, v. 5, p. 152 – 153.
Lucia, F. J., 1995, Rock-fabric/petrophysical classification Nicholson, P. G., 2002, A 700 million year tectonic frame-
of carbonate pore space for reservoir characterization: work for hydrocarbon exploration and production in
AAPG Bulletin, v. 79, p. 1275 – 1300. Saudi Arabia (abs.): GeoArabia, v. 7, p. 284.
Lucia, F. J., J. W. Jennings, M. Rahnis, and F. O. Meyer, Pleydell, S. M., B. Jones, F. J. Longstaffe, and H. Baadsgaard,
2001, Permeability and rock fabric from wireline logs, 1990, Dolomitization of the Oligocene – Miocene Bluff
Arab-D reservoir, Ghawar field, Saudi Arabia: GeoArabia, Formation on Grand Cayman, British West Indies:
v. 6, p. 619–646. Canadian Journal Earth Sciences, v. 27, p. 1098 –
Maiklem, W. R., D. G. Bebout, and R. P. Glaister, 1969, Clas- 1110.
sification of anhydrite; a practical approach: Bulletin Pollastro, R. M., A. S. Karshbaum, and R. J. Vigor, 1999,
of Canadian Petroleum Geology, v. 17, no. 2, p. 194 – Maps showing geology, oil and gas fields and geologic
233. provinces of the Arabian Peninsula: U.S. Geological
McGuire, M. D., G. Kompanik, M. Al-Shammery, M. Al- Survey Open-file Report, 97-470B, version 2.0, one
Amoudi, R. B. Koepnick, J. R. Markello, M. L. Stockton, CD-ROM.
and L. E. Waite, 1993, Importance of sequence strati- Powers, R. W., 1962, Arabian Upper Jurassic carbonate
graphic concepts in development of reservoir architec- reservoir rocks, in W. E. Ham, ed., Classification of car-
ture in upper Jurassic grainstones, Hadriya and Hanifa bonate rocks: AAPG Memoir 1, p. 122– 192.
reservoirs, Saudi Arabia: Proceedings of 8th Middle East Powers, R. W., 1968, Saudi Arabia: Lexique Stratigraphique
Oil Show, Society of Petroleum Engineers Paper 25578, International, 3: Paris, Centre National de la Recherche
p. 489 – 499. Scientifique, 171 p.
Meyer, F. O., 2000, Carbonate sheet slump from the Powers, R. W., L. R. Ramirez, C. D. Redmond, and E. L.
Jubaila Formation, Saudi Arabia: Slope implications: Elberg, 1966, Sedimentary geology of Saudi Arabia: Ge-
GeoArabia, v. 5, no. 1, p. 144 – 145. ology of the Arabian Peninsula: U.S. Geological Survey
Meyer, F. O., and R. C. Price, 1993, A new Arab-D deposi- Professional Paper 560-D, 150 p.
tional model, Ghawar field, Saudi Arabia: 8th Middle Radke, B. M., and R. L. Mathis, 1980, On the formation
East Oil Show and Conference Proceedings, Bahrain, and occurrence of saddle dolomite: Journal of Sedi-
p. 465 – 474. mentary Petrology, v. 50, p. 1149 – 1168.
Meyer, F. O., R. C. Price, I. A. Al-Ghamdi, I. M. Al-Goba, Read, J. F., 2004, Carbonate sequence stratigraphy, short
S. M. Al-Raimi, and J. C. Cole, 1996, Sequential stra- course notes: Saudi Aramco, Dhahran, Saudi Arabia,
tigraphy of outcropping strata equivalent to Arab-D unpaginated.
reservoir, Wadi Nisah, Saudi Arabia: GeoArabia, v. 1, Russell, S. D., M. Akbar, B. Vissapragada, and G. Walkden,
p. 435 – 456. 2002, Rock types and permeability prediction from
Meyer, F. O., G. W. Hughes, and I. Al-Ghamdi, 2000, Ju- dipmeter and image logs: AAPG Bulletin, v. 86, no. 10,
baila Formation, Tuwaiq Mountain escarpment, Saudi p. 1709 – 1732.
Arabia: Window to lower Arab-D reservoir faunal as- Russell, S. D., K. Sadler, W. H. Weihua, P. Richter, R. Y.
semblages and bedding geometry (abs.): GeoArabia, Eyvazzadeh, and E. A. Clerke, 2004, Applications of
v. 5, no. 1, p. 143. image log analyses to reservoir characterization,
Milankovitch, M., 1930, Mathematische klimalehre and Ghawar and Shaybah fields, Saudi Arabia: GEO 2004,
astronomische theorie der klimaschwankungen, in Middle East Geoscience Conference and Exhibition,
W. Koppen and R. Geiger, eds., Handbuch der Klima- abstract, p. 125.
tologie, I (A): Borntraeger, Berlin, Gebruder, 176 p. Scotese, C. R., 1998, Quick time computer animations,
Milankovitch, M., 1941, Kanon der erdbestrahlung und Paleomap Project: Department of Geology, University
seine anwendung auf das eiszeitenproblem: Akad of Texas at Arlington.
Royale Serbe, v. 133, 633 p. Sharland, P. R., R. Archer, D. M. Casey, R. B. Davies, S. H.
Mitchell, J. C., P. J. Lehmann, D. L. Cantrell, I. A. Al-Jallal, Hall, A. P. Heward, A. D. Horbury, and M. D. Simmons,
and M. A. R. Al-Thagafy, 1988, Lithofacies, diagenesis 2001, Arabian plate sequence stratigraphy: GeoArabia
and depositional sequence; Arab-D Member, Ghawar Special Publication 2, 371 p.
field, Saudi Arabia, in A. J. Lomando and P. M. Harris, Shearman, D. J., 1978, Evaporites of coastal sabkhas, in
eds., Giant oil and gas fields— A core workshop: SEPM W. E. Dean and B. C. Schreiber, eds., Marine evapo-
Core Workshop, v. 12, p. 459 – 514. rites: SEPM Short Course 4, p. 6 – 42.
Murris, R. J., 1980, Middle East stratigraphic evolution and Sibley, D. F., 1982, The origin of common dolomite fab-
oil habitat: AAPG Bulletin, v. 64, p. 597 – 618. rics: Clues from the Pliocene: Journal of Sedimentary
Newberry, B. M., L. M. Grace, and D. D. Stief, 1996, Analy- Petrology, v. 52, p. 1087 – 1100.
Ghawar Arab-D Reservoir: Widespread Porosity in Shoaling-upward Carbonate Cycles / 137

Sibley, D. F., 1991, Secular changes in the amount and U.S. Geological Survey, 1963, Geologic map of the Arabian
texture of dolomite: Geology, v. 19, p. 151 – 154. Peninsula, scale 1:2,000,000, and geologic quadrangle
Sibley, D. F., and J. M. Gregg, 1987, Classification of maps, scale 1:500,000: 1 map.
dolomite textures: Journal of Sedimentary Petrology, Wender, L. E., J. W. Bryant, M. F. Dickens, A. S. Neville,
v. 57, p. 967 – 975. and A. M. Al-Moqbel, 1998, Paleozoic (pre-Khuff ) hy-
Stegner, W., 1971, Discovery: The search for Arabian oil: drocarbon geology of the Ghawar area, eastern Saudi
Beirut, Middle East Export Press, 190 p. Arabia: GeoArabia, v. 3, p. 273 – 302.
Steineke, M., R. A. Bramkamp, and N. J. Sander, 1958, Strati- Xu, C., S. D. Russell, J. Gournay, and P. Richter, 2006,
graphic relations of Arabian Jurassic oil, in L. G. Weeks, Porosity partitioning and permeability quantifica-
ed., Habitat of oil: AAPG Symposium, p. 1294–1329. tion in vuggy carbonates using wireline logs, Permian
Swart, P. K., D. L. Cantrell, H. Westphal, C. R. Handford, Basin, West Texas: Petrophysics, v. 47, no. 1, p. 13 –
and C. G. St. C. Kendall, 2005, Origin of dolomite 22.
from Ghawar field, Saudi Arabia: Evidence from petro- Ziegler, M. A., 2001, Late Permian to Holocene paleofacies
graphic and geochemical constraints: Journal of Sedi- evolution of the Arabian plate and its hydrocarbon
mentary Petrology, v. 75, p. 476 – 491. occurrences: GeoArabia, v. 6, p. 445 – 504.
FIGURE 8. Core description of the SDGM Arab-D core. Six cores were taken, with a total cored interval of 299 ft (91 m), through the complete Arab-D reservoir.
Strohmenger, C. J., L. J. Weber, A. Ghani, K. Al-Mehsin, O. Al-Jeelani,

4
A. Al-Mansoori, T. Al-Dayyani, L. Vaughan, S. A. Khan, and J. C. Mitchell,
2006, High-resolution sequence stratigraphy and reservoir characterization
of Upper Thamama (Lower Cretaceous) Reservoirs of a giant Abu Dhabi
oil field, United Arab Emirates, in P. M. Harris and L. J. Weber, eds., Giant
hydrocarbon reservoirs of the world: From rocks to reservoir characterization
and modeling: AAPG Memoir 88/SEPM Special Publication, p. 139 – 171.

High-resolution Sequence
Stratigraphy and Reservoir
Characterization of Upper
Thamama (Lower Cretaceous)
Reservoirs of a Giant Abu Dhabi
Oil Field, United Arab Emirates
Christian J. Strohmenger, Lee Vaughan
Ahmed Ghani, Omar Al-Jeelani, ExxonMobil Exploration Company,
Abdulla Al-Mansoori, Houston, Texas, U.S.A.
and Taha Al-Dayyani,
Abu Dhabi Company for Onshore Oil Sameer A. Khan
Operations, Abu Dhabi, United Arab ExxonMobil Upstream Research Company,
Emirates Houston, Texas, U.S.A.

L. Jim Weber John C. Mitchell


ExxonMobil Exploration Company, Houston, ExxonMobil Exploration Company,
Texas, U.S.A. Houston, Texas, U.S.A.

Khalil Al-Mehsin
Abu Dhabi National Oil Company,
Abu Dhabi, United Arab Emirates

ABSTRACT

I
mportant hydrocarbon accumulations occur in platform carbonates of the
Lower Cretaceous Kharaib (Barremian and early Aptian) and Shuaiba (Aptian)
formations (upper Thamama Group) of Abu Dhabi. The Kharaib and Lower
Shuaiba formations contain three reservoir units separated by three low-porosity
and low-permeability dense zones. From base to top, the thickness of the res-
ervoir intervals range from approximately 80, 170, to 55 ft (24, 51, to 16 m),

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215876M883271

139
140 / Strohmenger et al.

respectively, for the Lower Kharaib, Upper Kharaib, and Lower Shuaiba Reservoir
Units. Core and well-log data of a giant oil field of Abu Dhabi, as well as outcrop
data from Wadi Rahabah in the Emirate of Ras Al-Khaimah were used to establish
a sequence-stratigraphic framework and a lithofacies scheme, applicable to all
three reservoir units and the three dense zones.
The Lower and Upper Kharaib Reservoir Units, as well as the lower, middle,
and upper dense zones are part of the late transgressive sequence set of a second-
order supersequence, made up of two third-order composite sequences. The over-
lying Lower Shuaiba Reservoir Unit belongs to the late transgressive sequence set
and the early highstand sequence set of this second-order supersequence and is
made up of one third-order composite sequence. The three third-order com-
posite sequences are composed of 19 fourth-order parasequence sets that show
predominantly aggradational and progradational stacking patterns, typical of
greenhouse cycles. Conventionally, composite sequence boundaries are placed
at or near the base of the three dense zones. As an alternative scenario, the pos-
sibility that the major composite sequence boundaries actually occur on top of
these dense zones is discussed.
On the basis of faunal content, texture, sedimentary structures, and litho-
logic composition, 13 reservoir lithofacies and 8 nonreservoir (dense) lithofacies
are identified from core. Similar lithofacies are identified in time-equivalent rock
exposures studied in Wadi Rahabah. Depositional environments of reservoir units
range from lower ramp to shoal crest to near-back shoal open-platform deposits.
Dense zones were deposited in an inner-ramp, restricted shallow-lagoonal setting.
Intensively bioturbated wackestone and packstone, and interbedded organic- and
siliciclastic-rich limestone, characterize the dense zones. Locally, mud cracks,
blackened grains, and rootlets are observed.
Outcrop analogs of subsurface reservoirs allow for a detailed investigation
of facies architecture and structure of carbonate bodies. Integration of subsur-
face and outcrop data (e.g., low-angle clinoforms that cannot be seen in core
data) leads to more insightful and realistic geological models of subsurface stra-
tigraphy. Geological model realizations based on core, outcrop, well-log, and
seismic data constrain fluid flow-simulation models. Results mimic known be-
havior in analogous producing fields, and the process of going from rock data to
simulation provides a useful training tool for reservoir characterization methods
and techniques.

INTRODUCTION the Lower Shuaiba Formation. The recoverable oil


reserves are estimated to be in billions of barrels.
Large hydrocarbon accumulations have been dis- More than 350 production and water-injection wells
covered and produced from platform carbonates of have been drilled to date.
the Lower Cretaceous Kharaib and Shuaiba forma- The Kharaib Formation is of Barremian and early
tions (Thamama Group) in Abu Dhabi (Alsharhan, Aptian age (Vahrenkamp, 1996; Granier, 2000; Pittet
1989; Alsharhan and Nairn, 1993, 1997). The focus of et al., 2002; Granier et al., 2003; C. Liu, T.-C. Huang,
this study is on the sequence stratigraphy and sedi- Y.-Y. Chen, 2005, personal communication) (Figure 3)
mentology of a giant Abu Dhabi oil field referred to and contains two reservoir units (Lower Kharaib Res-
herein as Field B (Figure 1). ervoir Unit and Upper Kharaib Reservoir Unit) sepa-
Field B was discovered in 1965 and went on pro- rated and encased by three zones of very low po-
duction in 1973. It is a faulted anticline (Figure 2) rosity and permeability, subsequently referred to as
with production of about 408 API oil from the Kha- dense zones (lower, middle, and upper dense zones;
raib Formation and planned future production from Figure 4). The thickness of the Lower Kharaib Reservoir
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 141

FIGURE 1. Location map showing the major oil fields of Abu Dhabi (green) and Field B (red) where core and well-log
data of the Kharaib and Lower Shuaiba formations were studied. Also shown is the location of the studied outcrops
at Wadi Rahabah, Ras Al-Khaimah.

Unit is about 80 ft (24 m), and the thickness of the because of pressure solution (compaction) and ce-
Upper Kharaib Reservoir Unit ranges between 150 ft mentation caused by the interaction of the carbon-
(45 m; downflank) and 190 ft (58 m; crestal area). ates with formation water and increasing burial
The thickness of the lower dense zone, underlying depth (Grötsch et al., 1998a). In contrast to the water
the Lower Kharaib Reservoir Unit, is approximately leg, the oil leg (mostly the crestal area) is not as af-
50 ft (15 m); the thickness of the middle dense zone, fected by pressure-dissolution processes (Burgess and
separating the Upper and Lower Kharaib Reservoir Peter, 1985). The presence of hydrocarbons seems to
Units, is approximately 45 ft (14 m), and the thick- inhibit the development of stylolites. The distribu-
ness of the upper dense zone, overlying the Up- tion of stylolites therefore shows an apparent dichot-
per Kharaib Reservoir Unit, is approximately 40 ft omy between crest (less stylolites) and flank (more
(12 m). stylolites), interpreted not to be related to differential
The thickness of the Lower Shuaiba Reservoir Unit burial effects between crest and flank, but instead, to
(overlying the upper dense zone; Figures 3, 4) is ap- the hydrocarbon infilling of the structure. Crestal
proximately 55 ft (17 m). wells (oil leg) are slightly thicker and show excellent
Highly permeable beds (>1 d) that mainly consist porosity and permeability compared to downflank
of coated-grain grainstone and rudist floatstone- wells (water leg), which exhibit reduced thickness
rudstone facies are recognized in the upper parts and reduced porosity and permeability (Grötsch et al.,
of the Upper and Lower Kharaib Reservoir Units 1998a).
(Figure 4). Generally, reservoir quality degrades from Traditionally, the subdivision of the Kharaib and
the crest to the flank throughout the field. This is Lower Shuaiba formations into reservoir units and
142 / Strohmenger et al.

FIGURE 2. Seismic line (west – east) running through the center of Field B shown in Figure 1. The seismic cross section
shows the structure of Field B and the interpreted main stratigraphic horizons from Upper Jurassic to Upper Cretaceous.
LKRU = Lower Kharaib Reservoir Unit; UKRU = Upper Kharaib Reservoir Unit; LSRU = Lower Shuaiba Reservoir Unit.

subunits is based on lithostratigraphic correlation, focused on reservoir characterization (Grötsch et al.,


using the vertical distribution of stylolites. Investi- 1998a; Melville et al., 2004), facies analyses, and se-
gation of stylolite distribution in the Kharaib For- quence stratigraphy (Grötsch et al., 1998a; Borgomano
mation shows that such an approach is not tenable et al., 2002; Pittet et al., 2002; van Buchem et al.,
and, therefore, not recommended for building geo- 2002; Hillgärtner et al., 2003; Immenhauser et al.,
logical (static) and, ultimately, reservoir (dynamic) 2004) and agree closely with the results presented in
models. The origin and the degree of stylolitization this chapter. The results of published and unpub-
depend on the heterogeneity of the sedimentary col- lished studies conducted at Field B were used to build
umn (e.g., preexisting surfaces, lithological contrasts, three-dimensional (3-D) geological and reservoir mod-
facies changes, and type of bedding; Park and Schot, els that adequately describe the reservoir (Grötsch
1968; Nelson, 1984). Moreover, some facies will be et al., 1998a; Melville et al., 2004) in its midterm pro-
less affected than others by stylolites. For example, duction state, but are based on only limited core data.
algal lithofacies will be more resistant to compaction The results presented in this chapter integrate core
during burial because of calcite cement trapping in and well-log data from Field B, as well as outcrop data
the algal structure. In addition, early diagenetic rim (Strohmenger et al., 2004a, b, c; Suwaina et al., 2004)
cement in interparticle pore space of grain-dominated from Ras Al-Khaimah (United Arab Emirates; Figure 1).
facies minimizes compaction. For the first time, detailed, sequence stratigraphy-
Previous Abu Dhabi Company for Onshore Oil keyed descriptions of approximately 15,000 ft (4580 m)
Operations (ADCO) internal and published studies of core material from 50 wells of Field B and 4 vertical
conducted on the Kharaib and Lower Shuaiba Res- outcrop sections of Wadi Rahabah (Ras Al-Khaimah)
ervoir Units of Field B and outcrop analogs of Oman have been conducted, considerably improving the
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 143

FIGURE 3. Upper Thamama Group


(Lower Cretaceous) sequence-
stratigraphic framework.

aries, parasequence set boundaries,


and parasequence boundaries (Boi-
chard et al., 1994; Rebelle et al.,
2004; Strohmenger et al., 2004b).
This can be explained by early dia-
genetic cementation occurring at
the top of the shallowing-upward
cycles, creating a contrast in rock
properties with the overlying rock
type. During burial, these hetero-
geneous intervals undergo prefer-
ential stylolitization. Calcite that
is dissolved during pressure solu-
tion reprecipitates in pores adja-
cent to the stylolites (Park and Schot,
1968; Koepnick, 1984), forming a
nonporous zone adjacent to the
chronostratigraphic boundary. If
the boundary separates grain-
dominated lithofacies (below)
from mud-dominated lithofacies
(above), cementation will be locat-
ed dominantly below the bound-
ary. In the case of grain-dominated
overlying material, cementation
may occur above and below the
chronostratigraphic boundary (Re-
belle et al., 2004; Strohmenger
et al., 2004b). Early diagenetic pro-
cesses thus follow the sequence-
stratigraphic framework and, there-
fore, can be predicted away from
well control.

preexisting facies and sequence stratigraphy models. THAMAMA GROUP


The newly developed high-resolution sequence- SEQUENCE-STRATIGRAPHIC FRAMEWORK
stratigraphic framework and facies scheme allow a
better prediction of the vertical and lateral distri- The Thamama Group can be described by two
bution of reservoir quality and reservoir continuity second-order supersequences. The older superse-
throughout Field B and have become standard for all quence corresponds to the Habshan Formation (Ber-
other ADCO fields. riasian and early Valanginian). The younger super-
Comparing the location of stylolites or denser, sequence encompasses the interval between the top
stylolite-bearing zones with the established high- Habshan and top Shuaiba formations (Sharland et al.,
resolution sequence-stratigraphic framework of the 2001, 2004; Davies et al., 2002; Droste and van Steen-
Kharaib Formation reveals that some stylolites are re- winkel, 2004; Strohmenger et al., 2004a, b, c; Haq
lated to third-, fourth-, and fifth-order sequence bound- and Al-Qahtani, 2005) (Figure 3). The transgressive
144 / Strohmenger et al.

FIGURE 4. Type well B-1,


showing the established high-
resolution sequence-stratigraphic
framework for the Kharaib and
Lower Shuaiba formations. ULRU 1 =
Upper Lekhwair Reservoir Unit 1.
See Figures 9 – 29 for color code
of lithofacies types.

Based on facies stacking pat-


terns (shallowing-upward trends),
sedimentary structures (e.g.,
Glossifungites burrows), and sur-
face morphologies (e.g., erosive
surfaces), 4 third-order compos-
ite sequence boundaries (SB),
3 composite maximum flooding
surfaces (MFS), and 13 marine
flooding surfaces (FS), which may
correspond to third- or higher
order flooding surfaces and se-
quence boundaries (flooding sur-
face/sequence boundary [FS/SB])
were identified from core mate-
rial of the Kharaib and Lower
Shuaiba formations and tied to
wire-line logs (Figure 4).
Where possible, third-order
SBs, MFSs, and parasequence
set boundaries (FS) were tied to
the stratigraphic framework
established by Sharland et al.
(2001, 2004) and Davies et al.
(2002) for the Arabian plate
(Figure 4). The nomenclature
proposed by Sharland et al.
(2001, 2004) is based on genetic
stratigraphic sequences (Gallo-
way, 1989), bounded by MFSs.
Each of their identified MFSs
is assigned a name, a number,
sequence set of this supersequence includes, from and an age, like K60 (123 Ma) and K70 (120 Ma). In
oldest to youngest, the Lekhwair (Valanginian to contrast, our sequence-stratigraphic subdivision fol-
Barremian), and the Kharaib (Barremian and early lows the sequence-stratigraphic approach estab-
Aptian) formations; including the Hawar (upper lished by ExxonMobil (Mitchum, 1977; Vail et al.,
dense zone). The Shuaiba Formation is of Aptian 1977, 1991; Vail, 1987; Van Wagoner et al., 1987, 1988;
age (Vahrenkamp, 1996; Grötsch et al., 1998b; Granier, Haq et al., 1988; Sarg, 1988; Sarg et al., 1999), where
2000; Pittet et al., 2002; Granier et al., 2003; C. Liu, depositional sequences are bounded by sequence
T.-C. Huang, Y.-Y. Chen, 2005, personal communi- boundaries. To tie our identified chronostratigraph-
cation) and is part of the late transgressive sequence ic boundaries back to the already existing strati-
set (early Aptian) and the highstand sequence set graphic framework of the Arabian plate, we had to
(Late Aptian) of this longer term depositional cycle. update the existing nomenclature proposed by
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 145

Sharland et al. (2001, 2004) and Davies et al. (2002) to as the lower dense zone (Figure 4), can be
in a way that our boundaries are assigned to the correlated for tens to hundreds of kilometers across
already existing names (e.g., K60 and K70) but also the platform. This interval is interpreted as the early
append information whether these boundaries cor- transgressive systems tract. Low- to moderate-energy,
respond to sequence boundaries (e.g., K60_SB), maxi- skeletal wackestone and packstone that overlie the
mum flooding surfaces (e.g., K60_MFS: correspond- lower dense zone are interpreted to correspond to the
ing to the K60 MFS of Sharland et al., 2001, 2004) or late transgressive systems tract (Figure 4).
flooding surfaces (e.g., K60_FS100). Overlying the transgressive systems tract (TST),
the highstand systems tract (HST) exhibits a normal
and diverse marine fauna and flora characterized by
Kharaib and Lower Shuaiba high-energy coated-grain, algal, skeletal grainstone
Sequence-stratigraphic Framework and rudstone with sparse rudist fragments (Figure 4).
The Kharaib Formation, including the Hawar The depositional environment is interpreted as a
(upper dense zone), corresponds to the late trans- tidal-influenced, high-energy bioclastic shoal envi-
gressive sequence set (TSS) of a second-order super- ronment. Water depth probably did not exceed 10 m
sequence. It is described by a second-order sequence (33 ft) even during the highest rates of relative sea
set (base lower dense zone to base upper dense zone level rise. Normal-marine conditions with good wa-
or Hawar), made up of two third-order composite ter circulation prevailed, as indicated by the generally
sequences (Figure 3). The lower third-order compos- diverse faunal content. Nutrients were sufficiently
ite sequence starts at the base of the lower dense zone abundant to sustain the development of the algal
and is capped by a pronounced sequence boundary buildups and rudists. Water turbidity was low as in-
(exposure surface) about 20 ft (6 m) below the middle dicated by the abundance of Lithocodium/Bacinella
dense zone (Figure 4). It is overlain by a third-order (Dupraz and Strasser, 1999; Immenhauser et al., 2005).
composite sequence that is bounded on top by a re- Four fourth-order parasequence sets are capped
gionally correlative sequence boundary below the by two marine flooding surfaces (K50_FS100 and
upper dense zone (Figure 4). Fourteen fourth-order K50_FS600), a third-order maximum flooding surface
parasequence sets and several fifth-order parase- (K50_MFS; Figure 5A), and a third-order sequence
quences that make up the two third-order composite boundary (K60_SB; Figure 5B). Several small-scale
sequences show predominantly aggradational and parasequences are identified in the Lower Kharaib
progradational stacking patterns, typical for green- sequence (K50_SB to K60_SB; Figure 4). These para-
house cycles (Sarg et al., 1999) (Figure 4). sequences are superimposed on a larger third-order
The Lower Shuaiba Formation corresponds to the trend. Successive parasequences show slightly differ-
late transgressive (TSS) and early highstand sequence ent characteristics. Oldest parasequences are mud-
set (HSS) of the same second-order supersequence dominated lagoonal sediments and show intense bio-
and is made up of a third-order composite sequence turbation and firmgrounds at bounding discontinuity
(Yose et al., 2004a, b, 2006) (Figure 3), subdivided by surfaces (Sattler et al., 2005). Firmgrounds formed
five fourth-order parasequence sets (Figure 4). during periods of subaerial exposure and subsequent
rapid increase in accommodation and reduced sedi-
mentation rates. Subsequent parasequences are thicker
Lower Kharaib Third-order Composite Sequence and are made up of lagoonal wackestone to packstone
Intensively burrowed and bioturbated wackestone and algal, skeletal, peloid floatstone-rudstone. Sedi-
and packstone and interbedded organic- and silici- mentation occurred in an open lagoonal environ-
clastic-rich limestone characterize the lower part ment with increased water circulation during loss of
of the Lower Kharaib sequence. Thalassinoides bur- accommodation. Upper parasequences are grain domi-
rows are abundant, and burrow fills are commonly nated (algal, skeletal, peloid floatstone-rudstone, coated-
dolomitized. Well-developed firmground surfaces grain grainstone, and coated-grain, algal, skeletal rud-
(Glossifungites burrows) cap several parasequences stone) and thinner, indicating an overall increase in
(Sattler et al., 2005). Locally, mud cracks, blackened current energy and decrease in accommodation. The
grains, and paleosols have been observed. Diversity uppermost parasequences are thin and dominated by
of fauna and flora is low. Other characteristics of this moderate- to high-energy miliolid shoals (peloid,
interval are centimeter- and decimeter-scale discon- skeletal packstone and peloid, skeletal grainstone).
tinuous beds. However, the overall interval, referred At this time, accommodation was nearly filled.
146 / Strohmenger et al.

FIGURE 5. Core photographs of chronostratigraphic boundaries. (A) Maximum flooding surface K50_MFS showing
burrowed and bored hardground (arrows), as well as patchy dolomitization (Do). (B) Sequence boundary K60_SB showing
stylolitically overprinted erosive surface, iron mineralization, and cementation below sequence boundary. (C) Maximum
flooding surface K60_MFS showing erosive contact between skeletal wackestone (SW, below) and skeletal, peloid packstone
(SPP, above), as well as patchy dolomitization (Do). (D) Sequence boundary K70_SB showing erosive and burrowed
surface (Glossifungites burrows, Gl), as well as cementation below sequence boundary. (E) Sequence boundary K80_SB
showing stylolitically overprinted erosive surface and cementation below sequence boundary. (F) Flooding surface
K70_FS100 on top of the upper dense zone (Hawar), which might also correspond to third-order sequence boundary
Ap3sb (see Figure 4). The pronounced pedogenic overprint of the carbonates below the erosive surface favors the
interpretation of a major subaerial exposure surface.

Upper Kharaib Third-Order Composite Sequence ervoir Unit (Figure 4) and is overlain by intensively
The early transgressive systems tract of the Upper burrowed and bioturbated skeletal, peloid wacke-
Kharaib sequence is composed of Orbitolina-rich, bio- stone and sparse packstone. This interval is laterally
turbated and burrowed skeletal, peloid wackestone continuous over considerable distance (tens to hun-
and packstone of the uppermost Lower Kharaib Res- dreds of kilometers) and is referred to as the middle
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 147

dense zone (Figure 4). Cycle tops are capped by firm- Lower Shuaiba Third-order Composite Sequence
grounds (Glossifungites burrows), with sediment- and The early transgressive systems tract of the third-
calcite spar-filled burrows (Sattler et al., 2005). The order Lower Shuaiba sequence corresponds to the up-
middle dense zone consists of abundant peloids, small per dense zone (Hawar), overlying the Upper Kharaib
echinoderm fragments, and some dasycladacean and sequence (Figure 4). This interval is rich in discoidal
red algae. In addition, other green algae, small mol- orbitolinids. Peloids, small echinoderm fragments, and
lusks, and sponge spicules are common. Locally, mud dasycladacean algae also occur. The upper dense zone is
cracks, blackened grains, and paleosols have been ob- organic and siliciclastic rich, and can be correlated for
served. In outcrops, the organic- and siliciclastic-rich tens to hundreds of kilometers across the platform
nature of the Kharaib middle dense zone results in a with varying thicknesses. Decimeter-scale beds are
friable, nodular weathered outcrop. marked by highly burrowed firmgrounds (Glossifungites
During the late transgressive systems tract, an open burrows). Mud cracks, blackened grains, and paleosols
lagoonal environment persisted on the platform. Dolo- are quite common, indicating frequent exposure.
mitized Thalassinoides firmgrounds indicate temporary The Lower Shuaiba Reservoir Unit, overlying the
cessation in sedimentation. Burrowed horizons are con- upper dense zone, is a shallow-to deeper water limestone
tinuous over many kilometers (Sattler et al., 2005). The consisting of algal (Lithocodium/Bacinella) boundstone
fauna is more diverse, with notable orbitolinids, echi- (Immenhauser et al., 2005) and skeletal, peloid wacke-
noderm remains, and bivalves. Carbonate rock textures stone to packstone and foraminifera, skeletal wacke-
are dominated by skeletal wackestone. At the ocean stone (Figure 4). This succession shows a deepening-
margin, rudist shoals, microbial mounds, and beaches upward trend interpreted to correspond to the late
developed at this time (van Buchem et al., 2002). transgressive systems tract. The overlying highstand
Skeletal, peloid wackestone-packstone, coated-grain systems tract is dominated by deeper marine forami-
grainstone, coated-grain, algal, skeletal grainstone nifera, skeletal mudstone to wackestone that show a
and rudstone, rudist and chondrodont floatstone and subtle shallowing-upward trend as well as an upward
rudstone, and peloid, skeletal packstone and grain- increase in grain richness (graining upward) to skel-
stone, rich in miliolids, gradually covered the entire etal (shell fragments), peloid wackestone to packstone.
platform during the highstand systems tract. The ru- The Lower Shuaiba composite sequence is bounded
dists include caprotinids (Glossomyophorus) and mono- on top by sequence boundary K80_SB (Figure 5E).
pleurids (Agriopleura) but no caprinids. Rudists, as well Five fourth-order parasequence sets are capped by
as chondrodonts, reflect deposition in shallow water. three marine flooding surfaces (K70_FS100, K70_FS200,
These grain-dominated lithofacies were deposited un- and K70_FS300), a third-order maximum flooding
der moderate- to high-energy, normal-marine condi- surface (K70_MFS), and a third-order sequence bound-
tions and exhibit considerable lateral variability. In ary (K80_SB; Figure 5E). Several small-scale parase-
the uppermost part of the sequence, miliolid-rich, quences are identified in the lower part of the Lower
skeletal, peloid packstone and grainstone with planar Shuaiba sequence (K70_SB to K80_SB; Figure 4).
bedding and low-angle cross-bedding imply moderate- The overlying Upper Shuaiba Reservoir Unit is dom-
to high-energy conditions. inated by deeper marine, planktonic foraminifera
In the Upper Kharaib sequence, shoaling did not wackestone-mudstone and intercalated low-porosity
occur until the uppermost part of the highstand sys- and low-permeability, burrowed, and bioturbated
tems tract. An increase in overall accommodation wackestone to mudstone (dense intervals).
from the Lower Kharaib to the Upper Kharaib sequence
may reflect a long-term second-order transgressive Alternative Sequence-stratigraphic Interpretation
trend in relative sea level (Figure 4). of the Kharaib Formation
Ten fourth-order parasequence sets bounded on The identified composite sequence boundaries
top by eight marine flooding surfaces (K60_FS100, K60_SB and K70_SB of the Kharaib Formation clearly
K60_FS200, K60_FS300, K60_FS600, K60_FS700, show indications of significant emersion of the Kha-
K60_FS800, K60_FS900, and K60_FS1000), a third-order raib platform carbonates (Figure 5B, D) and can re-
maximum flooding surface (K60_MFS; Figure 5C), and gionally be correlated throughout the entire Arabian
a third-order sequence boundary (K70_SB; Figure 5D), platform. However, the overall facies successions
as well as several small-scale parasequences, are from shallow-water upper ramp carbonates (reservoir
identified in the Upper Kharaib sequence (K60_SB to lithofacies types) below the sequence boundaries to
K70_SB; Figure 4). inner ramp (dense zones lithofacies types) above the
148 / Strohmenger et al.

FIGURE 6. Schematic models showing two possible sequence-stratigraphic interpretations of the dense zones, using
the upper dense zone as an example. (A) Conventional model: upper dense zone deposits correspond to the early
transgressive systems tract of the Lower Shuaiba sequence, overlying major third-order composite sequence boundary
(K70_SB) on top of the Upper Kharaib Reservoir. Minor higher order flooding surface (K70_FS100) occurs on top of the
upper dense zone. (B) Alternative model: upper dense zone deposits correspond to the late highstand systems tract of
the Upper Kharaib sequence, displaying major third-order composite sequence boundary (Ap3sb) at the top of the upper
dense zone. Minor third- or higher order sequence boundaries occur at the base (Ap2sb) and within the upper dense zone.

sequence boundaries follow a normal shallowing- ramp, restricted platform deposits, time-equivalent to
upward and progradational trend. In addition, the so- outer ramp, prograding carbonates (Figure 6). The omis-
called dense zones, especially the upper dense zone sion surfaces in the dense zones are interpreted to rep-
(Hawar), show frequent omission surfaces (firm- resent higher order sequence boundaries, indicating the
grounds), most likely corresponding to subaerial expo- frequent subaerial exposure of the inner ramp during
sure, and significant thickness variation throughout the late highstand systems tracts (reduced accommo-
the Arabian platform. For example, the upper dense dation). Most of the time represented by the dense
zone decreases in thickness from approximately 40 ft zones would actually correspond to nondeposition.
(12 m) in onshore Abu Dhabi to approximately 20 ft
(6 m) in offshore Abu Dhabi and is not present at the KHARAIB AND LOWER SHUAIBA FACIES
platform margin in Oman (van Buchem et al., 2002). AND DEPOSITIONAL ENVIRONMENT
Therefore, as an alternative sequence-stratigraphic in-
terpretation, composite sequence boundaries could On the basis of texture, grain types, sedimentary
be placed at the top of the lower (Barr4sb), middle structures, faunal content, and lithologic compo-
(Barr6sb), and upper (Ap3sb; Figure 5F) dense zones sition, 21 lithofacies types were defined for the
(Figure 4). The dense zones would thus correspond subsurface Kharaib and Lower Shuaiba formations
to late highstand systems tracts, representing inner- (Strohmenger et al., 2004a, b, c). Thirteen lithofacies
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 149

FIGURE 7. Paleobathymetrical profile showing the interpreted depositional environment of each of the thirteen
lithofacies types identified within the Lower and Upper Kharaib Reservoir Units and the Lower Shuaiba Reservoir Unit.

types (LF1–LF13) correspond to the three reservoir (Figure 7), and eight lithofacies types (LF20–LF27)
units: Lower Kharaib Reservoir Unit, Upper Kharaib occur in three dense zones: lower dense zone, middle
Reservoir Unit, and Lower Shuaiba Reservoir Unit dense zone, and upper dense zone (Figure 8). The

FIGURE 8. Paleobathymetrical profile showing the interpreted depositional environment of each of the eight lithofacies
types identified within the lower, middle, and upper dense zone.
150 / Strohmenger et al.

FIGURE 9. Reservoir lithofacies type 1: Rudist, peloid rudstone (RPR). Shown here are summary table, core photograph,
and thin-section photomicrograph (plane polarized light, porosity in blue). Thin-section photomicrograph shows large
chondrodont shells.

lithofacies types and their interpreted depositional  blocky calcite cement as common pore-filling ce-
environments are listed below. Interpretations de- ment associated with subaerial exposure surfaces
scribed here agree closely with the work of Pittet et al.  shallow subtidal, high-energy open platform
(2002) and van Buchem et al. (2002). above fair-weather-wave base
 shoal to upper-ramp rudist buildups and re-
worked buildups
Lithofacies Types of the Lower Kharaib,  high porosity and moderate to very high (>1 d)
Upper Kharaib, and Lower Shuaiba Reservoir Units matrix permeability
The analyzed 13 lithofacies types range from open  moldic, vuggy, and interparticle porosity
platform, deeper subtidal, lower ramp, to shallow sub-  molds and vugs that dominate under exposure
tidal to intertidal, upper-ramp environments (Figure 7). surfaces
A detailed description of the individual lithofacies  Lower (uncommon or not present at Field B) and
types is given below. For a summary, see Figures 9–21. Upper Kharaib Reservoir Units (see Figure 9)

Lithofacies LF1: Rudist, Peloid Rudstone (RPR) Lithofacies LF2: Rudist, Peloid Floatstone (RPF)

 rudists, chondrodonts, and other mollusks, fora-  rudists, chondrodonts and other mollusks, fora-
minifera, and echinoderms minifera, echinoderms, and sponge spicules
 peloids and coated grains (ooids, superficial ooids)  peloids
 rudstone texture  floatstone texture
 matrix: grainstone and packstone texture  matrix: packstone and wackestone texture
 moderate to well sorted  poor to moderate sorted
 bimodal grain-size distribution  bimodal grain-size distribution
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 151

FIGURE 10. Reservoir lithofacies type 2: Rudist, peloid floatstone (RPF). Shown here are summary table, core photograph,
and thin section photomicrograph (plane polarized light, porosity in blue).

FIGURE 11. Reservoir lithofacies type 3: Skeletal, peloid packstone (SPP) overlain by algal, skeletal, peloid floatstone-
rudstone (ASPF). Shown here are summary table, core photographs, and thin-section photomicrograph (plane polarized
light, stained with alizarin red-S, porosity in blue). Skeletal, peloid packstone may show fenestral structures (arrows;
left core photograph) and commonly occurs on top of shallowing-upward parasequence sets and parasequences
(Glossifungites burrows, Gl; right core photograph).
152 / Strohmenger et al.

FIGURE 12. Reservoir lithofacies type 4: Skeletal, peloid grainstone (SPG). Shown here are summary table, core
photographs, and thin-section photomicrograph (plane polarized light, stained with alizarin red-S, porosity in blue).
Core photograph (left) shows low-angle swash cross-bedding. Thin-section photomicrograph shows abundant miliolids
and small echinoderm plates.

FIGURE 13. Reservoir lithofacies type 5: Coated-grain, skeletal grainstone (CgSG). Shown here are summary table, core
photograph, and thin-section photomicrograph (plane polarized light, porosity in blue).
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 153

FIGURE 14. Reservoir lithofacies type 6: Coated-grain, algal, skeletal rudstone – floatstone (CgASR). Shown here are
summary table, core photograph, and thin-section photomicrograph (plane polarized light, porosity in blue).

FIGURE 15. Reservoir lithofacies type 7: Algal, skeletal, peloid floatstone – rudstone. Shown here are summary table, core
photographs, and thin section photomicrograph (plane polarized light, porosity in blue). Thin-section photomicrograph
shows preserved intraparticle porosity in Lithocodium/Bacinella grain and pore-filling cement (white areas).
154 / Strohmenger et al.

FIGURE 16. Reservoir lithofacies type 8: Algal, skeletal floatstone – boundstone (ASFB). Shown here are summary table,
core photograph, and thin-section photomicrograph (plane polarized light, stained with alizarin red-S, porosity in blue).
Thin-section photomicrograph shows preserved primary intraparticle (framework) porosity in Lithocodium/Bacinella
boundstone.

FIGURE 17. Reservoir lithofacies type 9: Orbitolinid, skeletal packstone (OSP). Shown here are summary table, core
photograph, and thin-section photomicrograph (plane polarized light, stained with alizarin red-S, porosity in blue).
Thin-section photomicrograph displays echinoderm plates and densely packed discoidal orbitolinids with preserved
primary intraparticle porosity.
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 155

FIGURE 18. Reservoir lithofacies type 10: Skeletal, peloid wackestone – packstone (SPWP). Shown here are summary
table, core photograph, and thin-section photomicrograph (stained with alizarin red-S, porosity in blue). Thin-section
photomicrograph shows benthonic foraminifera and small echinoderm plates.

FIGURE 19. Reservoir lithofacies type 11: Orbitolinid, skeletal wackestone (OSW). Shown here are summary table, core
photograph, and thin-section photomicrograph (plane polarized light, porosity in blue). Thin-section photomicrograph
displays discoidal orbitolinids.
156 / Strohmenger et al.

FIGURE 20. Reservoir lithofacies type 12: Skeletal wackestone (SW). Shown here are summary table, core photograph,
and thin-section photomicrograph (plane polarized light, porosity in blue).

FIGURE 21. Reservoir lithofacies type 13: Foraminifera, skeletal wackestone (FSW). Shown here are summary table, core
photograph, and thin-section photomicrograph (plane polarized light, stained with alizarin red-S, porosity in blue). Core
photograph shows bioturbation. Thin-section photomicrograph displays planktonic foraminifera and echinoderm plates.
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 157

 blocky calcite as common pore-filling cement  gradational basal contact and sharp and erosive
associated with subaerial exposure surfaces (dominantly) or gradational upper contact
 shallow subtidal, moderate-energy open platform  shallow subtidal to intertidal, high-energy open
above fair-weather-wave base platform above fair-weather-wave base
 upper ramp reworked rudist buildups  upper ramp beach, near shoal crest, and near
 low to high porosity and highly variable matrix inner shoal deposits
permeability  low to high porosity and low to high matrix
 moldic, vuggy, intraparticle, and microporosity permeability
 molds and vugs that dominate under exposure  interparticle, intraparticle, and micromoldic
surfaces porosity
 Lower (not present at Field B) and Upper Kharaib  Lower and Upper Kharaib Reservoir Units (see
Reservoir Units (see Figure 10) Figure 12)

Lithofacies LF3: Skeletal, Peloid Packstone (SPP) Lithofacies LF5: Coated Grain,
Skeletal Grainstone (CgSG)
 miliolids and other foraminifera, echinoderms,
algae, and mollusks  miliolids and other foraminifera, echinoderms, green
 miliolids that can be very abundant in the upper algae, Lithocodium/Bacinella fragments, and mollusks
part of the Upper and Lower Kharaib sequences  coated grains (ooids, superficial ooids, compos-
 peloids, aggregate grains, oncoids, and ooids ite grains, and oncoids), aggregate grains, intra-
 packstone texture with minor grainstone interlayer clasts, and peloids
 moderate to well sorted  grainstone texture with minor packstone interlayer
 unimodal grain-size distribution  moderate to well sorted
 uncommonly planar laminated and low-angle  unimodal to bimodal grain-size distribution
cross-bedded  uncommonly cross-bedded
 gradational basal contact and erosive and bur-  sharp basal contact and gradational (dominant)
rowed upper contact (Glossifungites burrows) or sharp upper contact
 caps fining- and shallowing-upward cycles  shallow subtidal, high-energy open platform
 desiccation cracks and circumgranular cracks above fair-weather-wave base
around grains, indicating subaerial exposure  upper-ramp, near-shoal crest deposits
 shallow subtidal to intertidal, moderate-energy  moderate to high porosity and low to very high
restricted and open platform above fair-weather- matrix permeability (>1 d)
wave base  interparticle, intraparticle, moldic, and vuggy
 inner shoal and upper ramp deposits porosity
 low to high porosity and low to moderate matrix  molds and vugs that dominate under exposure
permeability surfaces
 intraparticle, micromoldic, and microporosity  Lower and Upper Kharaib Reservoir Units (see
 Lower and Upper Kharaib Reservoir Units (see Figure 13)
Figure 11)
Lithofacies LF6: Coated-grain, Algal,
Lithofacies LF4: Skeletal, Peloid Grainstone (SPG) Skeletal Rudstone-Floatstone (CgASR)

 miliolids and other foraminifera and echinoderms  Lithocodium/Bacinella, miliolids and other fora-
 miliolids that can be very abundant in the upper minifera, echinoderms, green algae, rudists, and
part of the Upper and Lower Kharaib sequences other mollusks
 peloids, coated grains (ooids, superficial ooids  coated grains (composite grains, oncoids, super-
and composite grains) ficial ooids, and uncommon ooids), aggregate
 grainstone texture with minor packstone interlayer grains, intraclasts, and peloids
 well sorted  rudstone (predominant) to floatstone texture
 unimodal grain-size distribution  matrix: grainstone and mud-lean packstone texture
 planar laminated and low-angle cross-bedded indi-  poor to moderate sorted
cating tidal influence  bimodal grain-size distribution
158 / Strohmenger et al.

 sharp basal contact and gradational upper contact  moderate to high porosity and low to high ma-
 shallow subtidal, high-energy open platform above trix permeability
fair-weather-wave base  intraparticle (framework), shelter, and microporosity
 upper-ramp, near-shoal crest deposits  Lower Shuaiba Reservoir Unit (see Figure 16)
 moderate to high porosity and low to very high
(> 1 d) matrix permeability Lithofacies LF9: Orbitolinid,
 interparticle, intraparticle, moldic, and vuggy Skeletal Packstone (OSP)
porosity
 molds and vugs that dominate under exposure  orbitolinids and other foraminifera, mollusks,
surfaces echinoderms, and sponge spicules
 Lower and Upper Kharaib Reservoir Units (see  peloids
Figure 14)  packstone texture with minor wackestone inter-
layer
Lithofacies LF7: Algal, Skeletal,  moderate to well sorted
Peloid Floatstone-Rudstone (ASPF)  bimodal grain-size distribution
 bioturbation
 Lithocodium/Bacinella, foraminifera, echinoderms,  shallow subtidal, low- to moderate-energy re-
rudists and other mollusks, and sponge spicules stricted platform, and subtidal, open platform
 peloids, oncoids, aggregate grains, and uncom- near fair-weather-wave base deposits
mon coated grains (superficial ooids)  inner ramp, restricted lagoonal deposits, and
 floatstone (predominantly) to rudstone texture upper- to middle ramp deposits
 matrix: packstone to mud-lean packstone texture  low to high porosity and low matrix permeability
 poor to moderate sorted  intraparticle, micromoldic, and microporosity
 bimodal grain-size distribution  Lower and Upper Kharaib Reservoir Units (see
 sharp basal contact and gradational fining- Figure 17)
upward upper contact
 shallow subtidal, moderate- to high-energy open Lithofacies LF10: Skeletal,
platform above fair-weather-wave base Peloid Wackestone-Packstone (SPWP)
 upper-ramp algal buildups and reworked algal
buildups  foraminifera, sponge spicules, algae, and echinoderms
 moderate to high porosity and low to high  peloids and uncommon coated grains
matrix permeability  wackestone to packstone texture
 interparticle, intraparticle, and microporosity  moderately sorted
 Lower and Upper Kharaib Reservoir Units (see  unimodal grain-size distribution
Figure 15)  bioturbation
 thin dolomitized intervals common
Lithofacies LF8: Algal,  subtidal, low-energy open platform near fair-
Skeletal Floatstone-Boundstone (ASFB) weather-wave base
 upper to middle ramp deposits
 Lithocodium/Bacinella, orbitolinids and other  low to high porosity and low to moderate matrix
foraminifera, mollusks, and echinoderms permeability
 algal-binding activity forms low-relief digitate  intraparticle, micromoldic, and microporosity
growth framework and encrusting masses  Lower and Upper Kharaib Reservoir Units, and
 peloids Lower Shuaiba Reservoir Unit (see Figure 18)
 floatstone to boundstone (bindstone) texture
 matrix: packstone and wackestone texture Lithofacies LF11: Orbitolinid,
 poor to moderate sorted Skeletal Wackestone (OSW)
 bimodal grain-size distribution
 subtidal, low- to moderate energy open platform  orbitolinids and other foraminifera, mollusks,
near fair-weather-wave base echinoderms, and sponge spicules
 lower, upper-ramp algal buildups and reworked  peloids
algal buildups  wackestone texture with minor packstone interlayer
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 159

 moderate sorted  fine-grained, microporous limestone with low


 bimodal grain-size distribution matrix permeability
 bioturbation  microporosity, intraparticle, and micromoldic
 thin dolomitized layers porosity
 shallow subtidal, low-energy restricted platform,  Lower Shuaiba Reservoir Unit (see Figure 21)
and deeper subtidal, open platform below fair-
weather-wave base
 inner-ramp, restricted lagoonal deposits, and Lithofacies Types of the Lower, Middle,
middle-ramp deposits and Upper Dense Zones
 low to high porosity and low matrix permeability The analyzed eight lithofacies types represent
 microporosity, intraparticle, and micromoldic organic- and siliciclastic-rich, inner-platform, restricted
porosity lagoonal deposits (Figure 8). The frequent occurrence
 Lower and Upper Kharaib Reservoir Units (see of blackened pebbles and blackened grains suggests
Figure 19) a coastal, peritidal, oxygen-depleted environment
(Strasser and Davaud, 1983; Strasser, 1984). This is sup-
Lithofacies LF12: Skeletal ported by the elevated organic content of the carbon-
Wackestone (SW) ates, interpreted from high uranium readings of the
spectral gamma-ray log. Sedimentary structures like
 foraminifera, sponge spicules, echinoderms, and Glossifungites burrows, desiccation cracks, root marks,
algae and erosive surfaces, as well as paleosols indicate re-
 peloids peated subaerial exposures. A detailed description of
 wackestone texture the individual lithofacies types is given below. For a
 moderately sorted summary see Figures 22–29.
 well sorted
 unimodal grain-size distribution Lithofacies LF20: Wispy-laminated, Burrowed,
 bioturbation Skeletal Packstone (WBSP)
 thin dolomitized layers
 subtidal, low-energy open platform below fair-  low diversity, abundant, and small-size fauna
weather-wave base  foraminifera, mollusks, echinoderms, sponge
 middle-ramp deposits spicules, and ostracods
 fine-grained, microporous limestone with low to  peloids
moderate matrix permeability  blackened pebbles and blackened grains
 microporosity, intraparticle, and micromoldic  pyrite
porosity  siliciclastics present
 Upper Kharaib Reservoir Unit (see Figure 20)  packstone texture
 wispy- and nodular-bedded
Lithofacies LF13: Foraminifera,  burrows and bioturbation
Skeletal Wackestone (FSW)  shallow subtidal, low- to moderate-energy re-
stricted platform
 planktonic and benthonic foraminifera, sponge  inner shoal to inner ramp, restricted lagoonal
spicules, and echinoderms deposits
 peloids  low porosity and low matrix permeability
 pyrite  lower and middle (uncommon) dense zones (see
 wackestone texture with minor mudstone inter- Figure 22)
layer
 moderately sorted Lithofacies LF21: Wispy-laminated, Burrowed,
 well sorted Orbitolinid, Skeletal Packstone (WBOSP)
 unimodal grain-size distribution
 bioturbation  low diversity and high abundance of fauna
 subtidal, low-energy open platform below fair-  discoidal orbitolinids and other foraminifera,
weather-wave base green algae (dasycladacean), mollusks, echino-
 middle to lower ramp deposits derms, sponge spicules, and ostracods
160 / Strohmenger et al.

FIGURE 22. Nonreservoir lithofacies type 20: Wispy-laminated, burrowed, skeletal packstone (WBSP). Shown here
are summary table, core photograph, and thin-section photomicrograph (plane polarized light). Thin-section
photomicrograph displays abundant skeletal debris.

FIGURE 23. Nonreservoir lithofacies type 21: Wispy-laminated, burrowed, orbitolinid, skeletal packstone (WBOSP).
Shown here are summary table, core photograph, and thin-section photomicrographs (plane polarized light). Thin-
section photomicrographs show densely packed discoidal orbitolinids and wispy stylolites.
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 161

FIGURE 24. Nonreservoir lithofacies type 22: Wispy-laminated, burrowed, orbitolinid, skeletal wackestone (WBOSW).
Shown here are summary table, core photograph, and thin-section photomicrograph (plane polarized light).

FIGURE 25. Nonreservoir lithofacies type 23: Wispy-laminated, burrowed, skeletal wackestone – mudstone (WBSW).
Shown here are summary table, core photograph, and thin-section photomicrograph (plane polarized light). Thin-
section photomicrograph shows dolomite rhombs and pyrite.
162 / Strohmenger et al.

FIGURE 26. Nonreservoir lithofacies type 24: Burrowed, bioturbated, skeletal packstone (BBSP). Shown here are
summary table, core photograph, and thin-section photomicrograph (plane polarized light). Core photograph shows
intensive burrowing and bioturbation.

FIGURE 27. Nonreservoir lithofacies type 25: Burrowed, bioturbated, orbitolinid, skeletal packstone (BBOSP). Shown here
are summary table, core photograph, and thin-section photomicrographs (plane polarized light, stained with alizarin
red-S). Thin-section photomicrograph shows densely packed discoidal orbitolinids and glauconite (green grains).
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 163

FIGURE 28. Nonreservoir lithofacies type 26: Burrowed, bioturbated, orbitolinid, skeletal wackestone (BBOSW). Shown
here are summary table, core photographs, and thin-section photomicrograph (plane polarized light). Thin-section
photomicrographs show glauconite (green grains, left thin section photomicrograph), discoidal orbitolinids, and green
algae (right thin-section photomicrograph).

FIGURE 29. Nonreservoir lithofacies type 27: Burrowed, bioturbated, skeletal wackestone (BBSW). Shown here are
summary table, core photograph, and thin-section photomicrograph (plane polarized light, stained with alizarin red-S).
164 / Strohmenger et al.

 peloids  siliciclastics present


 blackened pebbles and blackened grains  wackestone to mudstone texture with minor
 dolomite rich packstone interlayer
 pyrite rich  wispy and nodular bedded
 siliciclastics present  burrows and bioturbation
 packstone texture  paleosols and caliche crusts
 wispy and nodular bedded  rhizoliths (root tubes) and desiccation cracks
 burrows and bioturbation  frequent subaerial exposure
 maroon lithoclasts, red paleosols, and caliche  shallow subtidal, low-energy restricted platform
crusts  inner-ramp, restricted lagoonal deposits
 rhizoliths (root tubes) and desiccation cracks  low porosity and low matrix permeability
 frequent subaerial exposure  lower and middle dense zones (see Figure 25)
 shallow subtidal, low- to moderate-energy re-
stricted platform Lithofacies LF24: Burrowed, Bioturbated, Skeletal
 inner-ramp, restricted lagoonal deposits Packstone (BBSP)
 low porosity and low matrix permeability
 upper dense zone (Hawar) (see Figure 23)  low diversity, abundant, and small-size fauna
 foraminifera, echinoderms, sponge spicules, and
Lithofacies LF22: Wispy-Laminated, Burrowed, ostracods
Orbitolinid, Skeletal Wackestone (WBOSW)  peloids
 blackened pebbles and blackened grains
 low diversity and high abundance of fauna  pyrite
 discoidal orbitolinids and other foraminifera,  packstone texture with minor wackestone inter-
green algae (dasycladacean), mollusks, echino- layer
derms, sponge spicules, and ostracods  burrows and bioturbation
 peloids  shallow subtidal, low- to moderate-energy re-
 blackened pebbles and blackened grains stricted platform
 dolomite rich  inner-ramp, restricted lagoonal deposits
 pyrite rich  low to moderate porosity and low matrix
 siliciclastics present permeability
 wackestone texture with minor packstone inter-  lower and middle (uncommon) dense zones (see
layer Figure 26)
 wispy and nodular bedded
 burrows and bioturbation Lithofacies LF25: Burrowed, Bioturbated,
 maroon lithoclasts, red paleosols, and caliche crusts Orbitolinid, Skeletal Packstone (BBOSP)
 rhizoliths (root tubes), and desiccation cracks
 frequent subaerial exposure  low diversity and high abundance of fauna
 shallow subtidal, low-energy restricted platform  discoidal orbitolinids and other foraminifera, green
 inner-ramp, restricted lagoonal deposits algae (dasycladacean), echinoderms, sponge spic-
 low porosity and low matrix permeability ules, and ostracods
 upper dense zone (Hawar) (see Figure 24)  peloids
 blackened pebbles and blackened grains
Lithofacies LF23: Wispy-laminated, Burrowed,  pyrite
Skeletal Wackestone-Mudstone (WBSW)  glauconite
 packstone texture with minor wackestone inter-
 low diversity, abundant, and small-size fauna layer
 foraminifera, mollusks, echinoderms, sponge  burrows and bioturbation
spicules, and ostracods  shallow subtidal, low- to moderate-energy re-
 peloids stricted platform
 blackened pebbles and blackened grains  inner ramp, restricted lagoonal deposits
 dolomite rich  low porosity and low matrix permeability
 pyrite rich  upper dense zone (Hawar) (see Figure 27)
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 165

Lithofacies LF26: Burrowed, Bioturbated, for the geological (static) and reservoir (dynamic)
Orbitolinid, Skeletal Wackestone (BBOSW) model. A generalized southwest–northeast cross sec-
tion shows the vertical and horizontal lithofacies
 low diversity and high abundance of fauna and facies association distribution of the Lower and
 discoidal orbitolinids and other foraminifera, green Upper Kharaib Reservoir Units as well as the Lower
algae (dasycladacean), echinoderms, sponge spic- Shuaiba Reservoir Unit in the established sequence-
ules, and ostracods stratigraphic framework (Figure 30). The lateral and
 blackened pebbles and blackened grains horizontal variations of the grain-dominated facies
 pyrite associations and lithofacies types in the highstand
 glauconite systems tracts of the Lower and Upper Kharaib se-
 wackestone texture with minor packstone inter- quences are evident (Figure 30).
layer
 burrows and bioturbation
 shallow subtidal, low-energy restricted platform KHARAIB FORMATION OUTCROP ANALOGS AT
 inner-ramp, restricted lagoonal deposits WADI RAHABAH, MUSANDAM PENINSULA
 low porosity and low matrix permeability (RAS AL-KHAIMAH)
 upper dense zone (Hawar) (see Figure 28)
The Musandam Peninsula consists of a series of
Lithofacies LF27: Burrowed, Bioturbated, Skeletal upthrusted blocks (Hagab thrust) of Permian to Cre-
Wackestone–Mudstone (BBSW) taceous strata. The dominant structures in the pen-
insula are north – south arcuate thrust faults, asso-
 low diversity, abundant, and small-size fauna ciated tear faults, and north –south folds. A slight
 foraminifera, mollusks, echinoderms, sponge regional northward dip of the peninsula toward Iran
spicules, and ostracods exists. The Ruus Al Jibal forms the backbone of the
 peloids peninsula, consisting of a series of upthrusted blocks
 blackened pebbles and blackened grains (Hudson, 1960). One of these upthrusted blocks is
 pyrite the 4760-ft (1450-m) high Jabal Hagab to the east of
 wackestone to mudstone with minor packstone the peninsula opposite to the city of Ras Al-Khaimah.
interlayer Rock exposures at Wadi Rahabah (north of the city
 burrows and bioturbation of Ras Al-Khaimah) are time equivalent to the Lower
 shallow subtidal, low-energy restricted platform Cretaceous upper Thamama Group, the most prolific
 inner ramp, restricted lagoonal deposits onshore oil- and gas-producing zones in Abu Dhabi.
 low porosity and low matrix permeability The Lower Cretaceous Kahmah Group (Oman) and
 lower and middle dense zones (see Figure 29) Musandam Group M4 (Musandam Peninsula) are
time equivalent to the Thamama Group (United
Arab Emirates, Saudi Arabia, and Bahrain), which con-
Sequence Stratigraphy-keyed tains (from oldest to youngest) the reservoirs of the
Lithofacies Distribution Habshan, Lekhwair, Kharaib, and Shuaiba forma-
Lithofacies distribution in the Lower and Upper tions of the subsurface of the United Arab Emirates.
Kharaib Reservoir Units is closely linked to the over- Outcrop analogs of subsurface reservoirs allow for
all sequence-stratigraphic framework. The transgres- a detailed investigation of the facies architecture
sive systems tracts of both the Lower and Upper Kha- and structure of these carbonate bodies, keyed to the
raib sequences are clearly more mud dominated (high overall sequence-stratigraphic framework (Alsharhan
porosity but low to moderate permeability), whereas and Nairn, 1993; Borgomano et al., 2002; van Buchem
the grain-dominated lithofacies types that show et al., 2002; Hillgärtner et al., 2003; Immenhauser
excellent porosity and permeability are nearly ex- et al., 2004; Strohmenger et al., 2004b; Suwaina et al.,
clusively restricted to the highstand systems tracts 2004).
(Figure 4). The identified 16 parasequence sets that The 21 lithofacies types that were defined for the
build the Lower (4 parasequence sets) and Upper subsurface Lower and Upper Kharaib and Lower
(8 parasequence sets) Kharaib Reservoir Units and the Shuaiba Reservoir Units and the lower, middle, and
Lower Shuaiba Reservoir Unit (4 parasequence sets) upper dense zones (Strohmenger et al., 2004a, b, c) are
define 16 reservoir subunits; the basic building blocks broadly applicable to the Kharaib and Lower Shuaiba
166 / Strohmenger et al.

FIGURE 30. Southwest – north-


east cross section through
Field B showing the established
sequence-stratigraphic frame-
work and the vertical and hori-
zontal distribution of facies
associations (lithofacies 1 [RPR] +
lithofacies 2 [RPF]; lithofacies 4
[SPG] + lithofacies 5 [CgSG] +
lithofacies 6 [CgASR]; and litho-
facies 9 [OSP] + lithofacies 11
[OSW]) and lithofacies types
(lithofacies 3: SPP; lithofacies 7:
ASPF; lithofacies 8: ASFB; litho-
facies 10: SPWP; lithofacies 12:
SW; and lithofacies 13: FSW).
Transgressive systems tract
(TST): green color; highstand
systems tract (HST): orange
color (see Figure 4).

thymetry and energy. The top


of a parasequence is deter-
mined by a flooding surface.
A sharp change in lithofacies
from shallow high-energy shoals
to slightly deeper water mud-
dominated lagoonal deposits
is recognized across parase-
quence boundaries. Mud-
dominated rocks grade upward
into packstone and floatstone
and subsequently into grain-
stone and rudstone with abun-
dant rudist debris. These rocks
indicate a progressive increase
in energy and shallowing of
the depositional environment.
The top of the parasequence
is marked either by well-sorted
grainstone and packstone
with abundant miliolids or
by deposits of the intershoal-
formations rock exposures. The depositional envi- protected areas that may become restricted and ex-
ronments do not occur randomly in time and space. hibit mud-dominated rocks with abundant forami-
They occur systematically in specific systems tracts nifera, green dasycladacean algae, small mollusks,
of the third-order composite sequences (Strohmenger and sparse rudists. The overlying parasequence com-
et al., 2004a, b, c; Suwaina et al., 2004; Vaughan et al., monly indicates a change to deeper and quieter de-
2004). positional environments as described, or it displays
The southeast wall of the wadi shows a carbonate a more gradual deepening with initial high-energy
succession that corresponds to the Lower and Upper deposits above a reactivation surface.
Kharaib Reservoir Units (Strohmenger et al., 2004a, Composite sequence boundary K70_SB at the
b, c; Suwaina et al., 2004) (Figure 31A). Meter-scale top of the Upper Kharaib sequence (top Upper Kha-
lithofacies stacking patterns reflect variations in ba- raib Reservoir Unit) displays pedogenic overprint of
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 167

FIGURE 31. Southeast wall of Wadi Rahabah outcrop. (A) Shown here is the Kharaib Formation measured section (base
lower dense zone to top upper dense zone; 250 ft [76 m]). (B) The top of the Upper Kharaib Reservoir (Upper Kharaib
sequence) displays pedogenic overprint of limestone beds (peds) below sequence boundary K70_SB. (C) Base of lower
dense zone showing erosive surface and mud cracks, corresponding to sequence boundary K50_SB.

limestone beds below the sequence boundary (paleo- subsurface data of Field B. Porosity measurements de-
sol: peds; Figure 31B). Composite sequence boundary rived from logs, core porosity, and core permeability
K50_SB at the base of the Lower Kharaib sequence data from an Abu Dhabi oil field (Field B) were ap-
(base lower dense zone) shows an erosive surface dis- plied to similar lithofacies observed in outcrop. These
playing desiccation cracks (Figure 31C). subsurface data were used in the outcrop 3-D model to
Outcrop analogs are a valuable source of data for more precisely reflect subsurface reservoir behavior
reservoir characterization. They are particularly use- because outcrops in the United Arab Emirates expe-
ful for visualizing interwell variability because sub- rienced a significantly different burial history relative
surface wellbores represent discrete, widely spaced to the subsurface reservoirs and, thus, have different
data. In other words, outcrops provide the continu- porosity and permeability (Strohmenger et al., 2004b;
ous, large-scale coverage of a seismic line and provide Suwaina et al., 2004). The objective of the reservoir
the fine-scale resolution of core measurements. Mod- modeling was to investigate the impact of reservoir
eling input parameters such as lateral and vertical architecture, porosity-permeability relationships, facies,
continuity (variogram range), object dimensions (as- and scale-up on the development of quantitative 3-D
pect ratios), detailed reservoir architecture (layering), geologic and fluid flow-simulation models. The use of
facies relationships, and the nature and extent of continuously exposed outcrops that are time equivalent
diagenetic features can be observed and measured. and have similar geology to the subsurface Upper
To better understand reservoir properties of the Thamama reservoirs of Abu Dhabi provide an opportu-
Upper Thamama producing zones in Abu Dhabi, a nity for improving geologic models of these reservoirs.
geologic model was developed from descriptions of For the stratigraphic interval studied in outcrop
time-equivalent, analogous outcrops in the Emirate (base lower dense zone to top upper dense zone), four
of Ras Al-Khaimah. Parasequences, maximum flood- composite reservoir facies (three permeable reservoir
ing surfaces, sequence boundaries, and lithofacies were facies and one dense facies) adequately described the
described in four measured sections and tied to the flow behavior. These reservoir facies were modeled
168 / Strohmenger et al.

FIGURE 32. Integration of outcrop data into geological, porosity, and permeability (fluid-flow) models. (A) North wall
of Wadi Rahabah showing clinoforms in the Upper Kharaib sequence. (B) Information gathered from outcrop studies
(occurrence of clinoforms) should be incorporated into geological (static) and reservoir (dynamic) models. (C) Porosity
model showing porosity distribution and the influence of clinoforms. High porosity is shown in red. No porosity is shown
in pink. (D) Permeability model showing resulting permeability distribution and the influence of clinoforms. Fluids will
encounter more baffles and travel more slowly through these dipping layers than through the flat-lying proportional
layers. High permeability is shown in red. No permeability is shown in pink.

and were scaled up into a simulation model. In addi- crop data (Figure 32A) should be used to build re-
tion to the four measured sections treated as wells, six alistic geological (static; Figure 32B) and reservoir (dy-
pseudowells were generated to help constrain the namic; Figure 32C, D) models (Strohmenger et al.,
mapping and modeling algorithms. These pseudo- 2004a, b, c; Suwaina et al., 2004; Vaughan et al., 2004).
wells represented copies of the measured sections and Porosity (Figure 32C) and permeability (Figure 32D)
contained realistic unit thickness changes based on models show that low-angle clinoforms will act as
outcrop observations. Multiple scenarios, including baffles, and fluid flow through these dipping layers
various injection patterns, scale-up methods, and lo- will therefore be slower compared to flat-lying pro-
cations of thin, high-permeability streaks, were inves- portional layers. Sector model simulation results ad-
tigated. Results mimic known behavior in analogous dress issues of scale-up, stratigraphic architecture, and
producing fields, and the process of going from rock diagenesis, as well as alternative recovery mecha-
data to simulation provide a useful training tool for nisms and field development scenarios.
reservoir characterization methods and techniques
(Vaughan et al., 2004). CONCLUSIONS
The north and south wall of the wadi show clino-
forms in the Upper Kharaib sequence. Local clusters The Kharaib Formation (Barremian and early Ap-
of fractures and offset of clinoforms by minor faults tian) contains two reservoir units (Lower and Upper
can also be observed (Figure 32A). Even if prograda- Kharaib Reservoir Unit) separated and encased by
tional patterns (clinoforms) cannot be identified in three so-called dense zones (lower, middle, and up-
core material, the information gathered from out- per dense zones). It is part of the late transgressive
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 169

sequence set of a second-order supersequence, made reservoir continuity in the subsurface (Vaughan et al.,
up of two third-order composite sequences (base lower 2004) (Figures 31, 32). Outcrops are particularly use-
dense zone to base upper dense zone or Hawar). The ful for filling the holes common in subsurface well
overlying Lower Shuaiba Reservoir Unit belongs to data sets used as input to geologic (static) and reser-
the late transgressive sequence set and the early high- voir (dynamic) models.
stand sequence set and is made up of one third-order
composite sequence. Nineteen fourth-order parase- ACKNOWLEDGMENTS
quence sets were identified in core and on well-log data.
Four parasequence sets build into the Lower Kharaib, The authors gratefully acknowledge the manage-
ten parasequence sets build into the Upper Kharaib, ments of Abu Dhabi Company for Onshore Oil Op-
and five parasequence sets build into the Lower erations (ADCO), Abu Dhabi National Oil Company
Shuaiba third-order composite sequences (Figure 4). (ADNOC), and ExxonMobil Exploration Company
These parasequence sets show predominantly aggra- for permission to publish this study. We thank Prasad
dational and progradational stacking patterns, typical Patannakara (ADCO) and his team for drafting the fig-
of greenhouse cycles (Sarg et al., 1999). The identified, ures. Appreciation is extended to Ismail Al-Mansouri,
chronostratigraphically bounded parasequence sets Majeed Haq, Mohammed Hilal, Hassan Ahmed, and
were used to subdivide the three reservoir units into Abdulsalam Khoury from ADCO core facility (Mus-
a total of 16 subunits. safah) for their help in preparing and laying out
As an alternative sequence-stratigraphic interpre- the core material and for preparing thin sections.
tation, the major third-order composite sequence Sean A. Guidry, Christine I. Gonzalez, Martin A. Herr-
boundaries (Figures 4, 6) may actually occur at the mann, Lisa A. Roehl, Chengije Liu, Susan M. Agar,
tops of the dense zones (lower, middle, and upper Jerry J. Kendall, and Mike G. Kozar (ExxonMobil), as
dense zones) and not, as conventionally interpreted well as Andrew Clark, Srikant Guruswamy, Hafez H.
(van Buchem et al., 2002) and discussed above, be- Hafez, Jorge S. Gomes, Michel J.-M. Rebelle, William L.
low the dense zones (Figure 6). Soroka, and Mohammed Ayoub (ADCO) are thanked
Thirteen lithofacies types correspond to the three for their valuable discussions. We extend special
reservoir units (Lower and Upper Kharaib Reservoir thanks to Omar Suwaina (ADNOC) for his valuable
Unit and Lower Shuaiba Reservoir Unit) and eight contributions to this study. We appreciate the help-
lithofacies types occur in the three dense zones (lower, ful comments and suggestions of reviewers José E.
middle, and upper dense zones; Figures 7–29). Mud- de Matos, Jürgen Grötsch, and Jonathan Kaufman,
dominated lithofacies types dominate the transgres- which greatly improved the manuscript.
sive systems tracts, whereas grain-dominated, highly
porous and permeable lithofacies types make up the
highstand systems tracts of the Lower and Upper Kha-
REFERENCES CITED
raib third-order composite sequences (Figure 30). Alsharhan, A. S., 1989, Petroleum geology of the United
These lithofacies types are broadly applicable to the Arab Emirates: Journal of Petroleum Geology, v. 12,
Kharaib and Shuaiba formations rock exposures (time- no. 3, p. 253 – 288.
equivalent to the Upper Thamama reservoir units) Alsharhan, A. S., and A. E. M. Nairn, 1993, Carbonate plat-
studied in Wadi Rahabah, Ras Al-Khaimah (United form models of Arabian Cretaceous reservoirs, in J. A. T.
Simo, R. W. Scott, and J.-P. Masse, eds., Cretaceous
Arab Emirates).
carbonate platforms: AAPG Memoir 56, p. 173 – 184.
Many of the so-called stylolitic intervals correspond
Alsharhan, A. S., and A. E. M. Nairn, 1997, Sedimentary
to major facies changes related to third-, fourth-, and basins and petroleum geology of the Middle East:
fifth-order sequence boundaries, parasequence set Amsterdam, Elsevier, 843 p.
boundaries, and parasequence boundaries. Early dia- Boichard, R., A. S. Al Suwaidi, and H. Karakhanian, 1994,
genetic processes follow the sequence-stratigraphic Sequence boundary types and related porosity evolu-
framework and, therefore, can be predicted away from tions: Example of the Upper Thamama Group in field
well control. ‘‘A’’ (offshore Abu Dhabi, UAE): 6th Abu Dhabi Interna-
tional Petroleum Exhibitions & Conference, Abu Dhabi
Outcrops of the Kharaib Formation at Wadi
Society of Petroleum Engineers Paper 76, Abu Dhabi,
Rahabah are of seismic scale and, thus, provide field- p. 417 – 428.
scale cross sections that help to refine sequence- Borgomano, J., J.-P. Masse, and S. Al Maskiry, 2002, The
stratigraphic and facies models and aid to the un- lower Aptian Shuaiba carbonate outcrops in Jebel
derstanding of reservoir geometry distribution and Akhdar, northern Oman: Impact on static modeling
170 / Strohmenger et al.

for Shuaiba petroleum reservoirs: AAPG Bulletin, v. 86, Hudson, R. G. S., 1960, The Permian and Triassic of the
no. 9, p. 1513 – 1529. Oman Peninsula, Arabia: Geological Magazine, v. 18,
Burgess, C. J., and C. K. Peter, 1985, Formation, distribution no. 4, p. 299 – 309.
and prediction of stylolites as permeability barriers Immenhauser, A., H. Hillgärtner, and E. Van Bentum, 2005,
in the Thamama Group, Abu Dhabi: 4th Middle East Microbial-foraminiferal episodes in the early Aptian of
Oil Show, Bahrain, Society of Petroleum Engineers the southern Tethyan margin: Ecological significance
Paper 13698, p. 165 – 174. and possible relation to oceanic anoxic event 1a: Sedi-
Davies, R. B., D. M. Casey, A. D. Horbury, P. R. Sharland, mentology, v. 52, no. 1, p. 77 – 99.
and M. D. Simmons, 2002, Early to mid-Cretaceous Immenhauser, A., et al., 2004, Barremian – lower Aptian
mixed carbonate-clastic shelfal systems: Examples, Qishn Formation, Haushi-Huqf area, Oman: A new
issues and models from the Arabian plate: GeoArabia, outcrop analogue for the Kharaib/Shu’aiba reservoirs:
v. 7, no. 3, p. 541 – 598. GeoArabia, v. 9, no. 1, p. 153 – 194.
Droste, H., and M. van Steenwinkel, 2004, Stratal geome- Koepnick, R. B., 1984, Distribution and vertical perme-
tries and patterns of platform carbonates: The Creta- ability of stylolites within a Lower Cretaceous carbon-
ceous of Oman, in G. P. Eberli, J. L. Masaferro, and J. F. ate reservoir, Abu Dhabi, United Arab Emirates —
Sarg, eds., Seismic imaging of carbonate reservoirs and Stylolites and associated phenomena: Relevance to
systems: AAPG Memoir 81, p. 185 – 206. hydrocarbon reservoirs, Abu Dhabi, U.A.E.: Abu Dhabi
Dupraz, C., and A. Strasser, 1999, Microbialites and micro- Reservoir Research Foundation Special Publications,
encrusters in shallow coral bioherms (middle – late p. 261 – 278.
Oxfordian, Swiss Jura Mountains): Facies, v. 40, p. 101 – Melville, P., O. Al Jeelani, S. Al Menhali, and J. Grötsch,
130. 2004, Three-dimensional seismic analysis in the char-
Galloway, W. E., 1989, Genetic stratigraphic sequences in acterization of a giant carbonate field, onshore Abu
basin analysis: I — Architecture and genesis of flood- Dhabi, United Arab Emirates, in G. P. Eberli, J. L. Masa-
ing-surface bounded depositional units: AAPG Bulle- ferro, and J. F. Sarg, eds., Seismic imaging of carbonate
tin, v. 73, no. 2, p. 125 – 142. reservoir systems: AAPG Memoir 81, p. 123– 148.
Granier, B., 2000, Lower Cretaceous stratigraphy of Abu Mitchum Jr., R. M., 1977, Seismic stratigraphy and global
Dhabi and the United Arab Emirates — A reappraisal: changes in sea level: Part 11 — Glossary of terms used
9th Abu Dhabi International Petroleum Exhibition & in seismic stratigraphy, in C. E. Payton, ed., Seismic
Conference Proceedings, Paper 0918, Abu Dhabi, stratigraphy — Applications to hydrocarbon explora-
p. 526 – 535. tion: AAPG Memoir 26, p. 205 – 212.
Granier, B., A. S. Al Suwaidi, R. Busnardo, S. K. Aziz, and Nelson, R. A., 1984, Geologic analysis of naturally fractured
R. Schroeder, 2003, New insight on the stratigraphy reservoirs: Contributions in petroleum geology and
of the ‘‘Upper Thamama’’ in offshore Abu Dhabi engineering: Oxford, Gulf Publishing Company, 352 p.
(U.A.E.): Carnets de Géologie, Article 2003/5, p. 1 – 17. Park, W. C., and E. H. Schot, 1968, Stylolites: Their nature
Grötsch, J., O. Al-Jeelani, and Y. Al-Mehairi, 1998a, Inte- and origin: Journal of Sedimentary Petrology, v. 38,
grated reservoir characterization of a giant Lower p. 175 – 191.
Cretaceous oil field, Abu Dhabi, U.A.E.: 8th Abu Dhabi Pittet, B., F. van Buchem, H. Hillgärtner, J. Grötsch, and
International Petroleum Exhibition & Conference H. Droste, 2002, Ecological succession, paleoenvi-
Proceedings, Society of Petroleum Engineers Paper ronmental change, and depositional sequences of
49454, Abu Dhabi, p. 77 – 86. Barremian – Aptian shallow-water carbonates in north-
Grötsch, J., I. Billing, and V. Vahrenkamp, 1998b, Carbon- ern Oman: Sedimentology, v. 49, no. 3, p. 555– 581.
isotope stratigraphy in shallow-water carbonates: Rebelle, M., C. J. Strohmenger, A. Ghani, K. Al-Mehsin, and
implications for Cretaceous black-shale deposition: A. Al-Mansouri, 2004, Lower Cretaceous Upper Tha-
Sedimentology, v. 45, no. 4, p. 623 – 634. mama reservoir high-resolution sequence stratigraphy,
Haq, B. U., and A. M. Al-Qahtani, 2005, Phanerozoic cycles United Arab Emirates (abs.): GeoArabia, v. 9, no. 1, p. 136.
of sea-level change on the Arabian platform: Geo- Sarg, J. F., 1988, Carbonate sequence stratigraphy, in C. K.
Arabia, v. 10, no. 2, p. 127 – 160. Wilgus, B. S. Hastings, C. G. St. C. Kendall, H. W.
Haq, B. U., J. Hardenbol, and P. R. Vail, 1988, Mesozoic Posamentier, C. A. Ross, and J. C. Van Wagoner, eds.,
and Cenozoic chronostratigraphy and cycles of sea- Sea-level changes: An integrated approach: SEPM
level change, in C. K. Wilgus, B. S. Hastings, C. G. St. Special Publication 42, p. 155 – 181.
C. Kendall, H. W. Posamentier, C. A. Ross, and J. C. Sarg, J. F., J. R. Markello, and L. J. Weber, 1999, The second-
Van Wagoner, eds., Sea-level changes: An integrated order cycle, carbonate-platform growth, and reservoir,
approach: SEPM Special Publication 42, p. 71 – 108. source, and trap prediction, in P. M. Harris, A. H. Saller,
Hillgärtner, H., F. S. P. van Buchem, F. Gaumet, P. Razin, and J. A. T. Simo, eds., Advances in carbonate sequence
B. Pittet, J. Grötsch, and H. Droste, 2003, The Barre- stratigraphy: Application to reservoirs, outcrops, and
mian – Aptian evolution of the eastern Arabian carbon- models: SEPM Special Publication 63, p. 11 – 34.
ate platform margin (northern Oman): Journal of Sattler, U., A. Immenhauser, H. Hillgärtner, and M. Esteban,
Sedimentary Research, v. 73, no. 5, p. 756 – 773. 2005, Characterization, lateral variability and lateral
Sequence Stratigraphy and Reservoir Characterization of Upper Thamama Reservoirs / 171

extend of discontinuity surfaces on a carbonate plat- of sea level: Part 1. Overview, in C. E. Payton, ed.,
form (Barremian to lower Aptian, Oman): Sedimen- Seismic stratigraphy — Applications to hydrocarbon
tology, v. 52, no. 2, p. 339 – 361. exploration: AAPG Memoir 26, p. 49 – 212.
Sharland, P. R., R. Archer, D. M. Casey, R. B. Davies, S. H. Vail, P. R., F. Audemard, S. A. Bowman, P. N. Eisner, and
Hall, A. P. Heward, A. D. Horbury, and M. D. Simmons, C. Perez-Cruz, 1991, The stratigraphic signatures of
2001, Arabic plate sequence stratigraphy: GeoArabia tectonics, eustasy and sedimentology — An overview,
Special Publication 2, 371 p. in G. Einsele, W. Ricken, and A. Seilacher, eds., Cycles
Sharland, P. R., R. Archer, D. M. Casey, R. B. Davies, M. D. and events in stratigraphy: Berlin, Springer, p. 617 –
Simmons, and O. E. Sutcliffe, 2004, Arabic plate se- 659.
quence stratigraphy — Revisions to SP2: GeoArabia, van Buchem, F. S. P., B. Pittet, H. Hillgärtner, J. Grötsch,
v. 9, no. 1, p. 199 – 214. A. I. Al Mansouri, I. M. Billing, H. H. J. Droste, W. H.
Strasser, A., 1984, Black-pebble occurrence and genesis in Oterdoom, and M. van Steenwinkel, 2002, High-
Holocene carbonate sediments (Florida Keys, Baha- resolution sequence stratigraphic architecture of Bar-
mas, and Tunisia): Journal of Sedimentary Petrology, remian/Aptian carbonate systems in northern Oman
v. 54, no. 4, p. 1097 – 1109. and the United Arab Emirates (Kharaib and Shuaiba
Strasser, A., and E. Davaud, 1983, Black pebbles of the formations): GeoArabia, v. 7, no. 3, p. 461 – 500.
Purbeckian (Swiss and French Jura): Lithology, geo- Van Wagoner, J. C., R. M. Mitchum Jr., H. W. Posamentier,
chemistry and origin: Eclogae Geologicae Helveticae, and P. R. Vail, 1987, Seismic stratigraphy interpreta-
v. 76, p. 551 – 580. tion using sequence stratigraphy: Part 2. Key defini-
Strohmenger, C. J., L. J. Weber, A. Ghani, K. Al-Mehsin, tions of sequence stratigraphy, in A. W. Bally, ed., Atlas
and O. Al-Jeelani, 2004a, Sequence stratigraphy and of seismic stratigraphy: AAPG Studies in Geology 27,
reservoir characterization of the Lower Cretaceous v. 1, p. 11 – 14.
Kharaib Formation, Abu Dhabi (abs.): GeoArabia, v. 9, Van Wagoner, J. C., H. W. Posamentier, R. M. Mitchum
no. 1, p. 136. Jr., P. R. Vail, J. F. Sarg, T. S. Loutit, and J. Hardenbol,
Strohmenger, C. J., L. J. Weber, A. Ghani, M. Rebelle, K. Al- 1988, An overview of the fundamentals of sequence
Mehsin, O. Al-Jeelani, A. Al-Mansoori, and O. Suwaina, stratigraphy and key definitions, in C. K. Wilgus, B. S.
2004b, High-resolution sequence stratigraphy of the Hastings, C. G. St. C. Kendall, H. W. Posamentier,
Kharaib Formation (Lower Cretaceous, U.A.E.): 11th C. A. Ross, and J. C. Van Wagoner, eds., Sea-level
Abu Dhabi International Petroleum Exhibition & changes: An integrated approach: SEPM Special
Conference Proceedings, Society of Petroleum Engi- Publication 42, p. 39 – 46.
neers Paper 88729, Abu Dhabi, 10 p. Vaughan, R. L., S. A. Khan, L. J. Weber, O. Suwaina, A. Al-
Strohmenger, C. J., L. J. Weber, A. Ghani, A. Al-Mansoori, K. Mansoori, A. Ghani, C. J. Strohmenger, M. A. Herrmann,
Al-Mehsin, O. Suwaina, O. Al-Jeelani, and L. Vaughan, and D. Hulstrand, 2004, Integrated characterization of
2004c, Sequence stratigraphy and reservoir character- UAE outcrops: From rocks to fluid flow simulation: 11th
ization of the Kharaib Formation comparing outcrop Abu Dhabi International Petroleum Exhibition & Con-
and subsurface data (Lower Cretaceous, U.A.E.) (abs.): ference Proceedings, Society of Petroleum Engineers
AAPG International Conference and Exhibition Paper 88730, Abu Dhabi, 17 p.
(Abstracts), Cancun, p. A75. Yose, L. A., S. Bachtel, L. J. Weber, C. J. Strohmenger, A.
Suwaina, O. A., L. J. Weber, C. J. Strohmenger, L. Vaughan, Al-Mansoori, and O. Suwaina, 2004a, Integrated ap-
A. Al-Mansoori, S. Khan, and A. Ghani, 2004, Sequence proaches to carbonate reservoir characterization and
stratigraphy and reservoir characterization of the prediction: Examples from the United Arab Emirates
Thamama reservoirs and outcrop equivalents: A core fields and outcrops (abs.): GeoArabia, v. 9, no. 1, p. 146.
workshop and field seminar (abs.): GeoArabia, v. 9, Yose, L. A., et al., 2004b, New frontiers in 3D seismic char-
no. 1, p. 137 – 138. acterization of carbonate reservoirs: Examples from a
Vahrenkamp, V. C., 1996, Carbon isotope stratigraphy of supergiant field in Abu Dhabi: 11th Abu Dhabi Inter-
the Upper Kharaib and Shuaiba formations: Impli- national Petroleum Exhibition & Conference Proceed-
cations for the Early Cretaceous evolution of the Ara- ings, Society of Petroleum Engineers Paper 88689, Abu
bian Gulf region: AAPG Bulletin, v. 80, no. 5, p. 647 – Dhabi, 16 p.
662. Yose, L. A., A. S. Ruf, C. J. Strohmenger, J. S. Schuelke, A.
Vail, P. R., 1987, Seismic stratigraphy interpretation using Gombos, I. Al-Hosani, S. Al-Maskary, G. Bloch, Y. Al-
sequence stratigraphy: Part 1. Seismic stratigraphy in- Mehairi, and I. G. Johnson, 2006, Three-dimensional
terpretation procedure, in A. W. Bally, ed., Atlas of characterization of a heterogeneous carbonate reser-
seismic stratigraphy, v. 1: AAPG Studies in Geology 27, voir, Lower Cretaceous, Abu Dhabi (United Arab
p. 1 – 10. Emirates), in P. M. Harris and L. J. Weber, eds., Giant
Vail, P. R., R. M. Mitchum Jr., R. G. Todd, J. M. Widmier, S. hydrocarbon reservoirs of the world: From rocks to
Thompson III, J. B. Sangree, J. N. Bubb, and W. G. reservoir characterization and modeling: AAPG Mem-
Hatlelid, 1977, Seismic stratigraphy and global changes oir 88/SEPM Special Publication, p. 173 – 212.
Yose, L. A., A. S. Ruf, C. J. Strohmenger, J. S. Schuelke, A. Gombos, I. Al-Hosani,

5
S. Al-Maskary, G. Bloch, Y. Al-Mehairi, and I. G. Johnson, 2006, Three-
dimensional characterization of a heterogeneous carbonate reservoir,
Lower Cretaceous, Abu Dhabi (United Arab Emirates), in P. M. Harris
and L. J. Weber, eds., Giant hydrocarbon reservoirs of the world: From
rocks to reservoir characterization and modeling: AAPG Memoir 88/SEPM
Special Publication, p. 173 – 212.

Three-dimensional Characterization
of a Heterogeneous Carbonate
Reservoir, Lower Cretaceous,
Abu Dhabi (United Arab Emirates)
Lyndon A. Yose Jim S. Schuelke
ExxonMobil Qatar, Inc., Doha, Qatar ExxonMobil Upstream Research Company,
Houston, Texas, U.S.A.
Amy S. Ruf
ExxonMobil Upstream Research Company, Andy Gombos
Houston, Texas, U.S.A. ExxonMobil Upstream Research Company,
Houston, Texas, U.S.A.
Christian J. Strohmenger,
Ismail Al-Hosani, Shamsa Al-Maskary, Imelda G. Johnson
Gerald Bloch, and Yousuf Al-Mehairi ExxonMobil Exploration Company, Houston,
Abu Dhabi Company for Onshore Oil Texas, U.S.A.
Operations, Abu Dhabi, United Arab
Emirates

ABSTRACT

H
igh-effort three-dimensional (3-D) seismic data collected by the Abu
Dhabi Company for Onshore Oil Operations (ADCO) are some of the
highest quality data ever collected for a carbonate field. The 3-D seismic
data were integrated with core and log data to develop a new, volume-based
framework for enhanced reservoir characterization. The Lower Cretaceous
(Aptian) reservoir is positioned over a platform-to-basin transition and records
a diverse range of depositional facies and stratal geometries. Reservoir properties
vary predictably based on position along the platform-to-basin profile and po-
sition in the sequence-stratigraphic framework.
The Aptian reservoir interval (Shuaiba Formation) records a second-order
sequence set that is divided into five depositional sequences. Sequences 1 and
2 were deposited during the transgressive phase of the sequence set. These se-
quences are retrogradational, record the initial formation of a low-relief ramp,

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215877M882562

173
174 / Yose et al.

and are dominated by algal-prone facies. Ramp interior and margin facies of
the transgressive phase are characterized by high porosity and low perme-
ability because of mud-dominated textures and development of microporos-
ity. Sequence 3 was deposited during the highstand phase of the sequence set,
is mainly aggradational, and records the proliferation or rudists across the plat-
form top. Grain-dominated platform interior and margin facies of the high-
stand phase are the highest quality reservoir facies in the Shuaiba reservoir.
Sequences 4 and 5 were deposited during the late highstand phase of the se-
quence set. These sequences are progradational and record the progressive
downstepping of the platform margin onto a low-angle (1 – 28) slope. Clino-
forms of the late highstand phase are characterized by alternations of high and
low reservoir quality developed in response to relative sea level changes.
Sequence 6 was deposited during the second-order lowstand and forms the base
of the next overlying sequence set. Sequence 6 is composed primarily of fine-
grained siliciclastics and is a nonreservoir.
Results from the study have led to an improved understanding of platform
evolution and a volume-based framework for reservoir characterization. The in-
tegrated data set provides new insights on platform paleogeography, carbonate
facies architecture, and the geometry and mechanisms of carbonate platform
progradation. In the platform interior area, 3-D seismic data reveal a complex
mosaic of tidal channels, high-energy rudist shoals, and intershoal ponds that
impact reservoir sweep and conformance. At the basin margin, the seismic data
provide high-definition images of platform-margin clinoforms that impact
reservoir architecture and well-pair connectivity.
Business applications of the volume-based reservoir framework include
(1) use of 3-D seismic visualization technology for optimizing well placement,
identifying bypassed reservoirs, and evaluating reservoir connectivity; (2) inte-
gration of quantitative, volume-based seismic information into reservoir models;
(3) maximizing recovery through full integration of all subsurface data; and
(4) enhanced communication among geoscientists and engineers, leading to
improved reservoir management practices.

INTRODUCTION voir and has been under production since the 1960s.
As the field matures, a more detailed understand-
The stratigraphic and diagenetic complexities in- ing of the reservoir is required to optimize recovery.
herent in carbonate reservoirs require accurate reser- To address this need, the Abu Dhabi Company for
voir descriptions and models to optimize recovery. Onshore Oil Operations (ADCO) acquired a high-
Three-dimensional (3-D) seismic data provide the effort, 3-D seismic survey over the field. A joint study
only continuous source of information on reservoir between ADCO and ExxonMobil was undertaken to
properties in the subsurface. With increasingly more fully integrate the 3-D seismic data with core, well,
3-D seismic data available over carbonate fields, the and production data, leading to a new volume-based
challenge is to maximize the value of the seismic data reservoir description. The integrated data set provides
for characterization of carbonate reservoir architec- new perspectives on carbonate platform evolution,
ture and rock properties (Eberli et al., 2003; Sarg and the response of carbonate systems to sea level change,
Schuelke, 2003; Masafarro et al., 2004; Yose et al., 2004). and the influence of carbonate facies and stratal ar-
This study used high-quality seismic data collected chitecture on reservoir properties and architecture.
over a supergiant onshore field in Abu Dhabi to il- The volume-based reservoir framework provides a
lustrate the value of 3-D seismic data for carbonate powerful tool for evaluating full-field production and
reservoir characterization. The field produces mainly performance issues, leading to improved reservoir
from the Lower Cretaceous (Aptian) Shuaiba reser- management.
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 175

FIGURE 1. Regional map showing major hydrocarbon accumulations in the United Arab Emirates and the locations of
Field A1 and Field A2. The fields are on separate structures, but both produce from the Shuaiba reservoir. The box
around the studied fields in the inset map highlights the area covered in Figure 2.

STUDY AREA and is one of the ten largest carbonate fields in the
world. The field has been under peripheral water in-
Field Location jection since the mid-1970s. Pattern gas and water-
The location of the study area in the United Arab injection schemes are also being implemented in
Emirates is shown in Figure 1. Onshore Abu Dhabi in- parts of the field. A more detailed understanding of
cludes five supergiant fields, each of which is located reservoir architecture and connectivity is required to
over large north-northeast–trending anticlinal struc- optimize the new injection strategy and future field
tures. The present study is focused on the Lower Cre- development to maximize recovery.
taceous Aptian reservoir interval (Shuaiba Formation)
of one of the onshore fields (Field A1 in Figure 1). As
shown in Figure 2, Field A1 occurs over a large, dou- Database
bly plunging anticline that is about 35  20 km (22  The study area provides a world-class data set and
12.5 mi) in dimension. A small field is also present presents an opportunity to develop and test high-end
just south of Field A1 (Field A2 in Figure 2) that occurs reservoir characterization technologies. Greater than
over a separate and much smaller closure. Fields A1 500 wells exist in the field, more than 100 of which
and A2 were evaluated jointly in this study because are cored. Most wells contain a standard suite of open-
the data sets overlap, and both fields are producing hole logs. More than 40 yr of production and per-
mainly from the Shuaiba reservoir. formance data are available to validate geologic in-
Field A1, discovered in 1962, is estimated to con- terpretations and calibrate reservoir models. Despite
tain more than 20 billion bbl of original oil in place the high number of wells in the field, the average well
176 / Yose et al.

FIGURE 2. Time structure map on the top of the reservoir interval at Field A1. The field is located over a large anticlinal
structure that plunges to the north and south. Field A2 is a small field on a separate closure developed just south of
Field A1. As shown, separate seismic surveys exist over each field.

spacing in the field is 1 km (0.62 mi). A 3-D seismic highlight stratigraphic and paleogeographic varia-
survey was recently acquired over Field A1, covering tions across the entire study area.
an area of nearly 2000 km2 (772 mi2). The 3-D seismic Data evaluated as part of the present study include
data are 300-fold with 25-m (82-ft) trace spacing and the 3-D seismic surveys, a strike and dip cross section
have a peak frequency of 30 –35 Hz at reservoir depth grid that intersects more than 100 wells, 25 cored
(2440–2745 m; 8000– 9000 ft). A separate, but over- wells, and available production and performance data.
lapping, seismic survey was also recently acquired
over Field A2 (Figure 1). Seismic data in the southern Regional Stratigraphy and Paleogeography
survey are 150-fold and also provide high-quality The stratal hierarchy and nomenclature used in the
imaging of the Shuaiba reservoir. Subsequent figures present study is shown in Figure 3. The Shuaiba res-
show extracts from the merged seismic surveys to ervoir interval is part of a second-order supersequence
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 177

FIGURE 3. Chronostratigraphic chart for the Lower Cretaceous of the United Arab Emirates. The Lower Cretaceous
comprises a hierarchy of second- and third-order depositional cycles. The reservoir interval at Field A1 spans most of
the Aptian and forms the second-order Aptian sequence set. The Aptian sequence set, in turn, is divided into five third-
order depositional sequences (1 – 5). A sixth depositional sequence (6) forms a composite lowstand sequence and is
part of the next overlying supersequence. OAE1a is a globally recognized oceanic anoxic event and correlates to the
maximum flooding interval of the Lower Cretaceous supersequence (coincident with the maximum flooding interval of
the Aptian sequence set). The Hardenbol et al. (1998) time sequence terminology is keyed to a nannofossil zonation.
Nannofossil dating in the present study confirms the position of the Ap 3, Ap 4, and Ap 5 time sequences (x in the time-
sequence nomenclature means unconfirmed). MFS = maximum flooding surface.

that spans the Valanginian to Aptian stages and can shown in Figure 4, Field A1 straddles the margin of
be correlated across the Arabian platform (Sharland the Bab basin. Accordingly, the Shuaiba reservoir re-
et al., 2001). Based on interpretation of facies and cords a range of platform, margin, slope, and basin
their stacking patterns in the present study, and by environments and is ideally positioned to record the
Strohmenger et al. (2006), the Lower Cretaceous full evolution of the platform-basin system. The Bab
supersequence is subdivided into three second-order basin differentiated in the early Aptian as a shallow
sequence sets (Figure 3). The Shuaiba reservoir is with- intrashelf seaway in the central Arabian platform
in the Aptian sequence set, culminating in the second- (Figure 4) (Alsharhan, 1985; Alsharhan and Nairn,
order supersequence boundary at the top. A compar- 1993). The basin is bounded by an extensive carbon-
ison of the sequence-stratigraphic – based framework ate platform, and restricted, organic-rich facies were
with the lithostratigraphic nomenclature is shown deposited within the Bab basin proper. Carbonate
in Figure 3. Lithostratigraphic nomenclature in the platform and margin facies in the Shuaiba form a
Shuaiba varies regionally, underscoring the need for regional reservoir system, and organic-rich basinal
a regional, sequence-stratigraphic framework based facies provide a potential source rock. The Bab basin is
on chronostratigraphy. filled by prograding carbonates and siliciclastics that
The Aptian paleogeography had a pronounced in- were deposited during the late Aptian sea level low-
fluence on depositional patterns in the study area. As stand. Shales and argillaceous carbonates of the Nahr
178 / Yose et al.

FIGURE 4. Aptian paleogeographic framework showing the extent of the Bab basin. Field A1 straddles the platform-basin
transition. A south-facing margin is also present in Field A2, just south of Field A1. The northern extent of the Bab
basin is uncertain. The basin may have been open to the north and connected to the Neo-Tethys open ocean, or isolated
in the Arabian platform and surrounded by a shallow, carbonate shelf (Sharland et al., 2001).

Umr Formation were deposited during the subse- the signal-to-noise ratio (compare Figure 6A, B), se-
quent transgression and form a regional top seal over lection and generation of optimum seismic attribute
the Shuaiba reservoir. volumes to image geologic features (Figure 6B), and
corendering of multiple seismic attributes to bring
WORKFLOW out more detail (Figure 6C). The seismic discontinu-
ity volume shown in Figure 6B highlights the edges
Improved recovery from carbonate reservoirs re- of structural and stratigraphic features and is one of
quires a 3-D understanding of reservoir architecture the most useful volumes for characterization of the
and properties. The 3-D seismic visualization environ- Shuaiba reservoir. The volume mathematically quan-
ment provides a powerful platform for data integra- tifies trace-to-trace variations in the seismic data,
tion and evaluation and a catalyst for multidisciplin- with areas of high continuity shown in white and
ary interactions across geoscience and engineering areas of low continuity shown in black. In the current
boundaries. A volume-based reservoir characterization example, facies geometries are resolved with clarity
workflow was applied to maximize information from comparable to aerial photographs of modern carbon-
the 3-D seismic data (Figure 5). Several key elements ate systems (Figure 6D).
of the workflow will be highlighted, including (1) post- Calibration of the enhanced seismic data with core
stack optimization of seismic data; (2) calibration of and well data is required to understand the geologic
seismic with core and log data; (3) application of vol- origin of seismic features and their impact on field
ume interpretation techniques for rapid and accurate performance. In this study, images, such as those
definition of reservoir frameworks; (4) direct detection shown in Figure 6, were used for optimum selection
and prediction of reservoir properties from seismic of wells and cores to sample the full range of seismic
data; and (5) integration of seismic data into reservoir variability. Seismic information on reservoir archi-
models. tecture and rock properties is compared with reser-
As illustrated in Figure 6, a combination of volume voir performance data to validate predictions and
types and corendering techniques are used to maxi- assess new opportunities. Seismic interpretations and
mize the information extracted from the 3-D seismic predictions are also incorporated into the reservoir
data. The approach used in this study included a com- model to guide reservoir framework development, dis-
bination of poststack filtering of the data to improve tribute reservoir features extracted from the seismic
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 179

FIGURE 5. Integrated seismic characterization workflow. Visualization technology allows for improved data integration
and evaluation during all stages of the workflow.

data, and to estimate rock properties outside of well Cretaceous reservoirs. In the study area, faults are
control. Many elements of the seismic workflow are organized into a system of grabens that are best de-
applicable to other carbonate reservoirs, although the veloped in the central and northern regions of the
details will vary on a reservoir-specific basis, depend- field. The faults are low offset, most commonly with
ing on such factors as reservoir complexity, data less than 18 ms (40 m; 130 ft) of vertical displace-
quality and availability, and the business purpose. ment, and are commonly arranged en echelon along
strike (lateral offsets). These faults are consistent
with a regional fault system that has been mapped
STRUCTURAL FRAMEWORK across onshore and offshore Abu Dhabi (Marzouk
and Sattar, 1993; Johnson et al., 2005). The regional
The impact of faults and fractures on reservoir faults are interpreted as low-offset wrench faults
performance is a key uncertainty for many large, low- that developed in response to west-northwest com-
relief structures in the Middle East, including on- pression during the Late Cretaceous ( Johnson et al.,
shore fields in Abu Dhabi. Prior to collection of 3-D 2005).
seismic data over the study area, the distribution and The impact of the faults and associated fractures
geometry of faults were not well understood (Alshar- on fluid flow and reservoir connectivity in the study
han, 1993; Marzouk and Sattar, 1993). As illustrated area is uncertain at present. The Shuaiba is a matrix-
in Figure 7, the 3-D seismic data provide a new level dominated flow system, but fracturing could enhance
of detail on the structural framework. Several dif- fluid flow in some parts of the field and may become
ferent seismic attribute volumes were evaluated as more pronounced with time. As will be discussed fur-
part of the structural analysis, including disconti- ther, premature water breakthrough along the north-
nuity, azimuthal amplitude variation with offset, and ern platform-margin trend may be in part related to
fault enhancement volumes. The seismic discontinu- enhanced fluid flow along faults. Additional work
ity volume (Figure 7) reveals a network of dominantly and data are required to further evaluate the impact
northwest-trending faults that intersect the Lower of the faults and associated fractures on fluid flow.
180 / Yose et al.

FIGURE 6. Impact of data optimization on imaging of reservoir features in the platform interior area. The images of
(A – C) are identical time slices from three different attribute volumes. (A) Time slice of the original, unfiltered amplitude
data with reservoir features poorly defined. (B) Enhanced seismic volume that incorporates ExxonMobil proprietary
filtering and seismic attribute technologies (see text for explanation). (C) Image of seismic discontinuity corendered
with the isochron of the reservoir interval. Isochron thins highlight lower porosity areas (faster velocities), and isochron
thicks highlight higher porosity areas (slower velocities). The corendered image brings out even more detail on the
platform architecture and qualitative indications of reservoir quality. (D) Aerial photograph of modern carbonates of the
Great Barrier Reef illustrating ponds and channels similar to those imaged in seismic data (photo courtesy of National
Geographic Society).

The new 3-D fault framework can be used to monitor tives on stratigraphic and paleogeographic variations
future production trends and to guide future data in the Shuaiba, including summaries of previous work,
collection. are provided in Sharland et al. (2001), van Buchem
et al. (2002), and Greselle and Pittet (2005).
SEQUENCE-STRATIGRAPHIC OVERVIEW
Sequence Age Dating
Stratigraphic variability is the primary control on Age control in the study area is based on unpub-
Shuaiba reservoir architecture in the study area. An lished work by I. Billing and E. J. Oswald (1995, per-
overview of the new sequence model for the Shuaiba sonal communication) and new age dating conducted
is summarized in Figures 3 and 8. Regional perspec- by ExxonMobil. Age dates from I. Billing and E. J.
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 181

FIGURE 7. Seismic characterization of structural framework. Image shown includes a 3-D discontinuity probe cut by
a vertical seismic amplitude profile. Most of the linear, black discontinuities on the surface of the 3-D cube are faults
(fault interpretations shown in inset). Faults are low offset, are best developed in the platform-margin area, and trend
parallel to the depositional strike of the margin.

Oswald (1995, personal communication) are based boundary is interpreted to be within the Bab Member
on a combination of macrofossil dating (mainly ru- of the Shuaiba. These results are generally consistent
dists) and carbon isotopes and provide an important with age dates reported in Azer and Toland (1993),
constraint on the position of the lower–upper Aptian Hughes (2000), Sharland et al. (2001), and van Buchem
boundary across the study area. The ExxonMobil et al. (2002). Based on these results, the time span of
age dating is based on nannofossils that established the Aptian sequence set is estimated to be on the or-
ties to the global Aptian time sequences defined in der of 7 –9 m.y.
Hardenbol et al. (1998) (Ap terminology in Figure 3)
and provided a higher resolution time-stratigraphic Stratal Hierarchy and Reservoir Impact
zonation for the lower and upper Aptian. A hierarchy of stratigraphic cyclicity is developed
Results from the age dating are summarized in within the Shuaiba that has a pronounced influence
Figure 3. Sequence 1 is of Ap 2 age, sequences 2 and 3 on reservoir properties. Stratal geometries developed
are of Ap 3 age, sequence 4 is of Ap 4 age, and se- within the Aptian sequence set record a complete
quence 5 is of probable Ap 5 age (T. C. Huang, 2004, transgressive-regressive cycle that developed in re-
personal communication). The age of sequence 6 is sponse to a second-order rise and fall in sea level
not confirmed, but is interpreted as Ap 6 time. The (Figures 3, 8). A similar depositional pattern is rec-
overlying Nahr Umr Shale is Albian in age based on ognized regionally within the Aptian (Sharland et al.,
regional age dates (Sharland et al., 2001). The position 2001; van Buchem et al., 2002; Droste, 2004). Second-
of the Aptian stage boundaries in the study area is order stacking patterns control the overall architec-
uncertain (Figure 3). The Barremian–Aptian bound- ture of the Shuaiba platform.
ary is interpreted to be within the Kharaib Forma- The Aptian sequence set is subdivided into three
tion (Strohmenger et al., 2006), and the Aptian–Albian major phases of deposition, corresponding to the
182 / Yose et al.

FIGURE 8. Sequence- and seismic-stratigraphic overview of the Aptian sequence set. See Figure 13 for explanation facies
codes. (A) Sequence-stratigraphic framework and reservoir implications. The reservoir is subdivided into four major phases
of deposition, corresponding to phases of a long-term (second-order) sea level cycle, and six depositional sequences (1– 6),
corresponding to third-order sea level changes. Wells A – D are type logs for distinct areas of the field and are shown
in Figures 17 – 20, respectively. (B) Seismic geometries displayed on 2-D vertical slice (amplitude data, flattened on base
reservoir). Reservoir top is highlighted in blue, and reservoir base is highlighted in green. Amplitude anomalies in the
platform interior correspond to channel and pond features, as shown in time slice views in Figures 6 and 10.

transgressive, early highstand, and late highstand to different phases of the third-order relative sea level
phases of the second-order sea level cycle (Figure 8). cycles. Most sequences display well-developed trans-
A subsequent lowstand phase forms the base of the gressive and highstand systems tracts (TST, HST), al-
overlying second-order sequence set. Stratal geometry, though the character of systems tracts varies signifi-
stacking patterns, and facies development vary sys- cantly based on position in the second-order cycle.
tematically between each major phase of deposition Higher frequency (fourth- or fifth-order) cycles are
(Figure 8). best developed within the HSTs of third-order se-
The Aptian sequence set is further subdivided into a quences and control the distribution and character of
series of five depositional sequences that are each inter- flow units. Lowstand systems tracts (LST) (e.g., basin-
preted to record a third-order cycle of relative sea level ward shifts of the carbonate system) are generally
rise and fall. Third-order sequences control the distri- absent in the transgressive and early highstand phases
bution of major reservoir barriers and flow compart- of the Aptian sequence set because of a combination
ments in the Shuaiba reservoir (Figure 8A). Third-order of long-term accommodation increase, and the lack of
sequences are partitioned within the second-order high-amplitude, shorter term sea level changes during
sequence set as follows: sequences 1 and 2 comprise the Cretaceous greenhouse period. Lowstand systems
the transgressive phase; sequence 3 comprises the tracts are developed during the late highstand and
highstand phase; and sequences 4 and 5 comprise the lowstand phases of the Aptian sequence set, as long-
late highstand phase. Sequence 6 forms the lowstand term sea level was falling, and are represented by
phase of the next overlying sequence set. basinward shifts in carbonate sedimentation. At the
As shown in Figure 9, third-order depositional second-order scale, sequences 4–6 are all basinally re-
sequences are divided into systems tracts that are tied stricted sequences (Figure 8). At the third-order scale,
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 183

FIGURE 9. Time-stratigraphic evolution of the Aptian sequence set, showing the sequential evolution of the carbonate
buildup. See Figure 13 for explanation of facies codes. HST = highstand systems tract; LST = lowstand systems tract;
TST = transgressive systems tract.

lowstand and TSTs that developed during the late platform interior (also see Figure 6). The width of the
highstand are difficult to distinguish because both platform interior (along depositional profile) varies
systems tracts are basinally restricted caused by long- throughout time, but is approximately 20 km (12.5 mi)
term sea level fall (see sequences 4 and 5, Figure 8). wide in the seismic time slice shown in Figure 10.

PALEOGEOGRAPHIC ELEMENTS Northern Platform Margin


The northern platform margin is a broad zone of
Paleogeographic variations across the area had a relatively transparent seismic amplitude character. The
large impact on sequence and reservoir develop- margin records the aggradation of high-energy facies,
ment (Figure 4). Paleogeographic features are well with generally high reservoir quality. The discontinu-
imaged by the seismic data (see cross sectional view ity time slice shows that the margin is characterized by
in Figure 8B and time slice view in Figure 10). An relatively high continuity, consistent with the high
introduction to the paleogeographic elements and and continuous reservoir quality trends observed in
their seismic character is provided below. log and core data. The width of the high-energy north-
ern margin averages 8 km (5 mi). Maximum relief on
Platform Interior the northern margin is estimated at 75 m (250 ft).
The central part of the field area is dominated by
stacked, shallow-water platform facies. Seismic ho- Clinoform Belt
rizons are concordant but highly discontinuous be- The northern part of the field area comprises clino-
cause of rapid lateral changes in facies and reservoir forms (sequences 4 and 5) that record more than 10 km
properties. The seismic discontinuity time slice shown (6.25 mi) of progradation. The clinoforms are charac-
in Figure 10 highlights a pond-channel complex that terized by alternations of low-angle (1–28), sigmoidal
gives rise to the complex seismic character of the seismic reflections and higher angle (3–48), oblique
184 / Yose et al.

FIGURE 10. Seismic discontinuity time slice from flattened volume (flattened on base reservoir). Seismic surveys over
Fields A1 and A2 are spliced together to create this volume. Seismic variations highlight the main paleogeographic
elements across the field area, each of which has distinct reservoir characteristics. Platform margins are recognized
north and south of the central buildup. Green line shows location of cross sections in Figure 8. Red dashes show location
of calibration wells that were evaluated in the study.

seismic reflections (Figures 8, 9). Reservoir properties the south. Based on platform-margin geometries in
and stratal geometries are highly variable in the clino- seismic data, maximum depositional relief on the
forms, resulting in significant reservoir heterogeneity. southern margin is estimated to be 30 m (100 ft).

Bab Basin Intrashelf Embayment


The Bab basin is characterized by restricted, low- The southern platform margin passes rapidly into
energy facies, reflected in a relatively thin, condensed low-energy, restricted facies interpreted to have been
interval. This interval is separated from the overlying deposited in a shallow intrashelf embayment that
clinoform sequences by a seismic downlap surface developed during the time of sequence 3. The inferred
(Figure 8). Basinal facies are characterized by highly geometry of the embayment is shown in Figure 4, but
continuous and concordant seismic reflections. is poorly constrained outside the 3-D seismic data
area. At a regional scale, it is possible that the Shuaiba
Southern Platform Margin platform margin may have evolved into a series of
The 3-D seismic data reveal a previously unrecog- low-relief, isolated banks.
nized platform margin in the southern part of the
study area. The southern margin differentiates later SEQUENCE RECOGNITION
in time than the northern platform margin (Figure 8).
The southern platform margin is aggradational, with Criteria for recognition of depositional sequences
very limited progradation. Clinoforms that devel- and surfaces in the Shuaiba vary depending on po-
oped seaward of the northern margin are absent to sition in the Aptian sequence set and position along
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 185

FIGURE 11. Paired seismic section and line drawing illustrating the major phases of platform evolution during the
deposition of the Aptian sequence set. The maximum flooding surface (MFS) shown occurs in sequence 2 and is the
second-order MFS for the Aptian sequence set.

the depositional profile (platform, margin, slope, and trated in Figure 12. Most of the surfaces are accom-
basin). Sequence-stratigraphic correlations are guid- panied by rapid facies changes above and below and
ed by a combination of seismic, log, and core data. by changes in cycle stacking patterns. Transects of
The seismic-stratigraphic architecture of the Aptian cored wells were selected to sample the full range
sequence set is summarized in Figure 11. Preliminary of stratigraphic and seismic variability observed
sequence interpretations are weighted largely on seis- across the platform (Figures 8, 10). Surface correla-
mic geometries and termination patterns (e.g., down- tions through the northern platform margin are dif-
lap, toplap, and onlap). Because seismic surfaces cor- ficult because of the thick stack of relatively contin-
respond closely to time lines, seismic interpretations uous, high-porosity rock in sequences 2 and 3.
are used as a template to guide higher resolution
correlations derived from core and log observations. Sequence Boundaries
In general, seismic surfaces are well expressed in the Because of the prevailing Cretaceous greenhouse
clinoforms of the northern field area where there are climate pattern during the Aptian, sequence bound-
significant amplitude variations and seismic geome- aries are generally not pronounced. The best devel-
tries (Figure 11). Seismic surfaces in platform interior oped sequence boundaries are those developed at the
and margin areas are difficult to constrain because of base and top of the Aptian sequence set, especially
less geometric variation and minimal impedance con- the upper sequence set boundary (Figures 8, 9). Third-
trast (Figures 8, 11). Further complexity is introduced and fourth-order sequence boundaries are best devel-
by the high-amplitude bursts in the platform interior oped in the highstand and late highstand phases of
associated with the pond and tidal-channel complex the sequence set because of a decrease in long-term
(Figure 8B). Seismic interpretation in platform inte- accommodation (Figures 8, 9).
rior and margin areas is guided by well ties and the use In updip areas of the Shuaiba (platform position),
of multiple seismic volumes (e.g., amplitude, band sequence boundaries are commonly expressed as thin,
limited impedance, and discontinuity). iron-oxide – encrusted surfaces, with pyrite mineral-
Some of the sequence-stratigraphic surfaces are ization (Figure 12A, B). Evidence for dissolution is
well expressed in core (e.g., exposure surfaces, hard- limited to within a few feet or inches of the exposure
grounds, stylolitic boundaries, shale seams), as illus- surfaces. Coarse, blocky calcite cements are developed
186 / Yose et al.

FIGURE 12. Core expression of sequence-stratigraphic surfaces. SB = sequence boundary; MFS = maximum flooding
surface; HFS = high-frequency (fourth-order) surface; HST = highstand systems tract. Surfaces in (A, B) are long-term
exposure surfaces, forming the sequence boundary of the Lower Cretaceous supersequence. Surfaces shown in (D, E)
are candidates for composite exposure and flooding surfaces, as discussed in Immenhauser et al. (2000). Characteristics
of the surfaces are discussed in the text.

locally below sequence boundaries and can penetrate Maximum Flooding Surfaces
as deep as 3 m (10 ft) below the exposure surface. The maximum flooding surface of the Aptian se-
Sequence boundaries are also marked by rapid facies quence set is interpreted at the base of the last platform-
changes above and below and are sometimes overlain margin backstep in the platform area, coincident with
by coarse skeletal and/or lithoclastic lag deposits. the maximum flooding surface of sequence 2 (Figures 8,
In downdip areas (slope and basin positions), se- 11). The equivalent surface in the slope and basin
quence boundaries become conformable and lack evi- environment is interpreted within the condensed in-
dence for exposure. However, some sequence bound- terval that separates shallow-water platform facies be-
aries record major facies shifts. For example, in the low from deeper water clinoform facies above (coinci-
clinoform area, sequence boundaries correspond to dent with the seismic downlap surface; Figures 8, 11).
a shift from grainy peloidal-skeletal facies (porous Flooding surfaces in the platform area generally oc-
zones) in the HST to burrowed, miliolid-rich wack- cur at or near a transition in facies and stacking pat-
estone (dense zones) in the LST to TST above (Figure 9). terns that record a change from deepening-upward to
In some cases, the downdip sequence boundaries are shoaling-upward characteristics. In some cases, it is
manifested by clay-rich seams that occur in the same difficult to pick an exact surface and is more appro-
position as the facies shifts (Figure 12C) and generate priate to identify a flooding interval. Flooding surfaces
a clear gamma-ray log response to the surface. (or intervals) are commonly well expressed in slope
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 187

FIGURE 13. Facies explanation and key to core log symbols used in this paper. Facies are subdivided based on the
position in the Aptian sequence set and the position along the platform to basin profile. Core photos and reservoir-quality
data for reservoir facies are illustrated in Figures 14 – 16. Core-to-log facies calibrations are shown in Figures 17 – 20. MLP =
mud-lean packstone; skel = skeletal; pel = peloidal; bndst. = boundstone; fltst. = floatstone; HST = highstand systems tract;
TST = transgressive systems tract.

and proximal basin environments and are generally Transgressive Phase


expressed as condensed intervals that include concen- The transgressive phase records the deposition dur-
trations of clay or muddy carbonate sediments, bur- ing the rising limb of the long-term sea level cycle
rowed hardgrounds, and/or stylolites (Figure 12D–F). and includes sequences 1A–1B and 2 (Figure 9). Dur-
ing this period, the rate of relative sea level rise was
SEQUENCE DESCRIPTION greater than the carbonate accumulation rate. As
a consequence, depositional relief was established
A general description of depositional sequences in over time, marking the initiation of the Bab basin
relationship to evolution of the Shuaiba platform is in the area. The carbonate buildup differentiated
provided below. The discussion is keyed to the four as a low-relief ramp that attained as much as 45 m
main phases of the Aptian accommodation cycle in- (150 ft) of depositional relief during the transgressive
troduced above. A key to the facies nomenclature and phase. Ramp interior facies are dominated by an algal-
symbols used in this report is provided in Figure 13. stromotoporoid assemblage, with some associated
Core photos of representative facies types for each corals and rudists (Figure 14). The faunal assemblage
major phase of the Aptian sequence set are displayed is interpreted to indicate relatively restricted circula-
in Figures 14 – 16 (transgressive, early highstand, and tion during the initial development of the Bab in-
late highstand phases, respectively). Type logs for each trashelf basin. Ramp interior and margin facies of
of the main areas of the buildup are provided in the transgressive phase are characterized by high po-
Figures 17 – 20 (platform interior, platform margin, rosity (10–25%) and low permeability (0.1–5 md) be-
proximal clinoform sequence, and distal clinoform cause of mud-dominated textures and development
sequence, respectively). of microporosity.
188 / Yose et al.

FIGURE 14. Examples of depositional facies and reservoir properties in the composite transgressive interval (sequences 1
and 2). Colors in the upper right of photos are keyed to Figure 13. Scale bar is in 1-cm (0.4-in.) increments. IP = intraparticle;
BP = between particle; md = millidarcys; Perm = permeability; Por = porosity; HST = highstand systems tract; TST =
transgressive systems tract.

The transgressive phase comprises as many as contrast, van Buchem et al. (2002) interpreted the
three depositional sequences (sequences 1A, 1B, and Hawar to record a regional transgression associated
2) as summarized below. with the initial formation of the Bab basin. In this
interpretation, the Hawar may be the TST of se-
Sequence 1A quence 1, as opposed to a separate depositional se-
Sequence 1A includes the Hawar shale (Figure 9) quence. More regional study is required to test these
and was not extensively evaluated in this study. The alternative interpretations.
Hawar comprises argillaceous, Orbitolina-rich pack-
stone and wackestone and contains cyclic alterna- Reservoir Quality
tions of more and less argillaceous deposits. As dis- The Hawar is of very low porosity and permeability
cussed further in Strohmenger et al. (2006), there are and forms an effective barrier (dense) between the
alternative sequence-stratigraphic interpretations of Kharaib reservoir below and Shuaiba reservoir above.
the Hawar shale. Strohmenger et al. (2006) inter-
preted that the Hawar was deposited in relatively Sequence 1B
shallow, restricted environments with episodic expo- Sequence 1B comprises the basal part of the
sure. Strohmenger et al. (2006) also interpreted that Shuaiba Formation and records a major change in
the Hawar is bounded above and below by sequence deposition relative to sequence 1A. The argillaceous
boundaries (interpretation shown in Figure 3). In content in sequence 1B drops off significantly, and
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 189

FIGURE 15. Examples of depositional facies and reservoir properties in the composite highstand interval (sequence 3).
Colors in the upper right of photos are keyed to Figure 13. Scale bar is in 1-cm (0.4-in.) increments. IP = intraparticle;
BP = between particle; md = millidarcys; Perm = permeability; Por = porosity; TST = transgressive systems tract.

there is a shift from Orbitolina-dominated facies to permeability from 0.1 to 5 md. The primary controls on
algal-dominated facies. Similar to sequence 1A, se- reservoir quality variability include (1) sparse versus
quence 1B records shallow-water deposition across the dense algal boundstone fabrics (intraparticle porosity in
study area, with no indication of depositional relief. algal boundstone is better connected if the algal compo-
The sequence is divided into three shoaling-upward nents are touching [dense]) and (2) grainy versus muddy
cycles that can be correlated across the study area matrix in the algal floatstone (grainier fabrics have higher
based on core and log character (Figures 17, 18). Fa- porosity and permeability). The chalky, fine-grained
cies stacking patterns in the high-frequency cycles facies (skeletal and peloidal wackestones and packstones)
are variable, but a general progression, from base to are characterized by the development of microporos-
top, would include (1) skeletal-peloidal wackestone, ity with relatively high porosity, but low permeability.
(2) algal boundstone-floatstone, and (3) chalky peloidal-
skeletal packstone (with microporosity). Sequence 2
Sequence 2 records the initial differentiation of
Reservoir Quality the platform margin and inception of the Bab basin
The algal boundstone and floatstone represent a in the study area (Figure 9). The sequence is differ-
range of facies that have variable reservoir quality entiated into distinct TST and HST, each with dif-
(Figures 17, 18). Porosity ranges from 10 to 20% and fering reservoir architecture and quality (Figure 9).
190 / Yose et al.

FIGURE 16. Examples of depositional facies and reservoir properties in the composite late highstand interval (sequences 4
and 5). Colors in the upper right of photos are keyed to Figure 13. Scale bar is in 1-cm (0.4-in.) increments. IP = intraparticle;
BP = between particle; md = millidarcys; Perm = permeability; Por = porosity; HST = highstand systems tract; TST =
transgressive systems tract.

The TST is characterized by a backstepping ge- densed interval corresponds with early Aptian global
ometry that records the development of the northern oceanic anoxic event (OAE1a; Figure 3) (Bralower et al.,
ramp margin. The initial margin develops via two 1994; Erba et al., 1999; Premoli Silva et al., 1999).
small-scale backsteps that, together, created on The HST records aggradation of the northern ramp
the order of 20 m (65 ft) of relief (Figure 9). Algal- margin and the inception of the southern ramp mar-
stromotoporoid boundstone is the dominant facies gin. A diverse suite of facies is represented including
in the TST (Figure 14A, B). Cycle stacking patterns shallow-water, algal-prone facies in the ramp interior
change from exposure-capped shoaling-upward and fine-grained mud-lean packstone on the ramp
cycles below (sequence 1A), to amalgamated algal- margin (Figure 14B). Faunal diversity increases slight-
stromotoporoid facies above (sequence 2, TST). The ly in the HST, with an increased abundance of corals
basin is characterized by deposition of organic-rich and rudists (Figure 14C). The increase in faunal diver-
wackestone and packstone (Figure 14F). These basinal sity may be related to more open-circulation patterns
facies immediately overlie shallow-water carbonate in the Bab intrashelf basin in response to the marine
facies of sequence 1B and record the initial platform transgression. Evidence for decreasing accommoda-
drowning (Figure 12F). The basinal deposits are in- tion during the HST phase includes a shift from re-
terpreted as the condensed interval equivalent to the trogradational to aggradational geometry along the
maximum flooding interval of the Aptian sequence northern ramp margin and increased evidence for
set. Nannofossil age dates indicate that the con- exposure. An exposure surface is present at the base
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 191

of a prominent biostromal unit in the HST interval


(Figure 8). The biostromal unit corresponds to the
mid-C coral zone of Russell et al. (2002), ranges be-
tween 3 and 8 m (10 and 25 ft) thick, and comprises a
complex assemblage of corals, stromotoporoid, and
algae, most of which are in growth position (Figure 14C).
The biostromal unit is interpreted to have nucleated
on the underlying exposure surface. Algal bound-
stone and floatstone with increasing amounts of rudist
debris overlie the biostrome interval and continue up
to the sequence boundary at the top of sequence 2.
The upper sequence boundary is not a prominent
exposure surface, but records a major shift from algal-
dominated facies below to rudist-dominated facies
above. The faunal shift is rapid, commonly occurring
in less than 1 ft (0.3 m) of vertical section. The surface
recording the faunal shift is of regional extent (Hughes,
2000; van Buchem et al., 2002; Droste, 2004) and ap-
pears to indicate a major change in circulation pat-
terns across the Bab basin.

Reservoir Quality
Vertical variations in facies and reservoir quality in
sequence 2 are shown in Figures 17 and 18. Facies
textures in the HST are generally grainier that those in
the underlying TST. The HST is dominated by algal-
boundstone and floatstone with a grain-prone ma-
trix, whereas the underling TST is dominated by algal-
boundstone and floatstone with a mud-prone matrix.
As a result, reservoir quality increases upward in se-
quence 2 (Figures 17, 18). In the TST, porosity ranges
between 15 and 25%, and permeability ranges from
1 to 20 md. In the HST, porosity ranges between 20
and 30%, and permeability ranges from 10 to 100 md.
The condensed interval deposited within the basinal
area is dense and forms a barrier between the shallow-
water platform facies of sequence 1B below and
overlying slope facies of sequences 4 and 5 above.

Early Highstand Phase


During the early highstand phase, the rate of long-
term relative sea level rise was in balance with the
carbonate sedimentation, resulting in aggradation of

FIGURE 17. Type log for the platform interior (well A in


Figure 8). Vertical scale is in 10-ft (3-m) depth increments.
Porosity values below 5% are shaded in gray. The 1-md
permeability line is highlighted in red. See Figure 13 for
key to facies codes and symbols. HST = highstand systems
tract; TST = transgressive systems tract; SB = sequence
boundary; MFS = maximum flooding surface.
192 / Yose et al.

the carbonate platform. Maximum depositional re-


lief was attained during the aggradational phase,
with more than 75 m (250 ft) of relief established
between the platform and basin (Figure 9). Facies and
faunal diversity reached a maximum during the early
highstand phase, with the proliferation of rudists
across the bank top (Figure 15A, B, D) and the de-
velopment of margins to the north and south of the
platform. The development of an open, high-energy
platform may have been triggered by changes in circu-
lation patterns in the Bab basin, in combination with
relative sea level rise. Platform interior and margin
facies deposited during the early highstand phase
are the highest reservoir quality facies in the Shuaiba.
Sequence 3 comprises the early highstand phase of
deposition and is subdivided into distinct TST and
HST, as discussed below.

Sequence 3 (TST)
The TST of sequence 3 records significant changes
in platform and margin deposition. The southern
platform margin attained maximum development at
this time, with as much as 30 m (100 ft) of relief be-
tween the platform and shallow embayment to the
south. Depositional patterns in the platform interior
changed significantly in response to changes in plat-
form paleogeography. The platform interior differen-
tiated into a complex mosaic of tidal channels, rudist
shoals, and intershoal depressions or ponds, marking
the establishment of significant cross-bank currents
and open-marine conditions (Figures 9, 10).
Platform-margin facies differ significantly between
the northern and southern margins. The northern
platform margin is characterized by a broad (5–7-km;
3.1 –4.4-mi) grainstone belt that passes northward
into deeper water slope facies. A narrow band (1 –
2 km; 0.6 – 1.2 mi) of in-situ rudist and stromotopor-
oid boundstone and bafflestone is located platform-
ward of the grainstone belt and is interpreted to form
a discontinuous band of stromotoporoid and rudist
patch reefs along the northern platform. The north-
ern margin sand belt is characterized by skeletal-to-
peloidal packstone and grainstone that are inter-
preted to have been deposited in a high- to moderate

FIGURE 18. Type log for the northern platform margin


(well B in Figures 8, 11). Vertical scale is in 10-ft (3-m)
depth increments. Porosity values below 5% are shaded
in gray. The 1-md permeability line is highlighted in red.
See Figure 13 for key to facies codes and symbols. HST =
highstand systems tract; TST = transgressive systems tract;
SB = sequence boundary; MFS = maximum flooding surface.
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 193

energy carbonate sand flat. Skeletal components are


diverse and include local concentrations of Orbitolina-
rich and miliolid-rich grains. Textures are relatively
fine grained and become coarser grained and cleaner
(less muddy) upward within the section. Textures be-
come finer and muddier toward the basin (downslope).
In contrast, the southern platform margin is charac-
terized by high-energy, rudist-dominated facies that
cover a broad region of the outer platform. Deposi-
tional facies along the southern margin include a
stack of rudist-dominated grainstone and boundstone
(Figure 15A). Little vertical variation in this facies exists,
and no internal surfaces or bedding features were iden-
tified. High-energy facies along the southern margin
pass gradationally into the mosaic of rudist shoals,
intershoal depressions, and channels characteristic
of the platform interior to the north and grade rapidly
into Orbitolina-rich packstone and grainstone of the
upper- to middle-slope environment to the south.

Sequence 3 (HST)
The HST records a significant change in platform
and margin deposition relative to the underlying TST
(Figure 9). Depositional energy levels in the platform
interior become much lower. The complex associa-
tion of channels, ponds, and shoals developed in the
TST are absent in the HST. In contrast, the HST re-
cords a major decrease in depositional energy across
the bank top and an attendant decrease in reservoir
quality. Carbonate facies are generally mud domi-
nated, with a notable increase in argillaceous con-
tent. In addition, in contrast to the amalgamated
rudist shoal facies characteristic of the TST, the HST is
characterized by the development of shoaling-up-
ward parasequences with well-developed surfaces.
Platform-margin facies and geometries also change
significantly. The width of the southern margin
narrows to 2 km (1.25 mi) and is characterized by
much lower energy facies than in the underlying
TST (Figure 9). The northern margin likewise narrows
relative to the underlying TST and evolves from a
fine-grained sand-flat environment into coarser rudist
rudstone and floatstone deposited as shoals. Facies
relationships and seismic geometries indicate that
FIGURE 19. Type log for the clinoform area (well C in the northern margin prograded approximately 3 km
Figures 8, 11). Well is in updip position of sequence 4, (1.9 mi) during the HST as the rate of sea level rise
highlighting the development of the dense TST. Vertical began to decrease (Figure 8).
scale is in 10-ft (3-m) depth increments. Porosity values The controlling factor on restriction of the plat-
below 5% are shaded in gray. The 1-md permeability
form interior during the HST may be related to ge-
line is highlighted in red. See Figure 13 for key to facies
codes and symbols. HST = highstand systems tract; ometry changes at the platform margins. In partic-
TST = transgressive systems tract; SB = sequence boundary; ular, the northern margin appears to become a more
MFS = maximum flooding surface. pronounced topographic feature, possibly in response
194 / Yose et al.

to shoaling and progradation of the margin. Progra- Reservoir Quality


dation and increased topographic expression of the Reservoir quality varies both vertically and laterally
northern platform margin may have restricted cir- within sequence 3 (Figures 17–20). Rudist shoal facies
culation in the platform interior and in the intrashelf deposited in the platform interior of the TST contain
embayment to the south. the highest reservoir quality of any interval in the
Shuaiba. Porosity ranges from 20 to 30%, and perme-
ability ranges from 10 to 500 md. The TST also marks
the period of maximum facies diversity caused by the
development of the pond-channel network in the plat-
form interior. Ponds and channels are filled with mud-
prone facies that have variable, but generally low, reser-
voir quality. Many of these features are nonreservoir.
Reservoir quality varies significantly between the
northern and southern margins (Figure 15). Rudist-
dominated facies of the southern margin have poros-
ity of 20 –30% and permeability in the 100–1000-md
range. In contrast, fine-grained grainstone and mud-
lean packstone deposited along the northern margin
has similar porosity to the southern margin, but per-
meability is much lower, in the 10– 100-md range.
The HST of sequence 3 is of generally low reservoir
quality, with the best reservoir facies corresponding
to rudist shoals developed along the northern mar-
gin. These facies range from 20 to 30% porosity, and
permeability ranges from 10 to 100 md.

Late Highstand Phase


The late highstand phase marks a fundamental
turning point in the history of the Aptian platform.
Relative sea level rise slowed to a near standstill and
then started to gradually fall. The carbonate platform
had limited space for vertical accumulation and be-
gan to move laterally, and downslope, into the Bab
basin. In response, the northern platform margin pro-
graded more than 10 km (6.2 mi) into the Bab basin.
Seismic geometries in the clinoform belt define two
major phases of progradation that correspond to two
depositional sequences (sequences 4 and 5; Figure 11).
Because of a gradual, long-term fall in sea level, each
slope sequence steps lower into the basin. Sequences 4
and 5 are thus detached from the main platform and
are slope restricted. The older Aptian platform was
exposed during most of the late hightstand phase,
and a long-term subaerial exposure surface was de-
FIGURE 20. Type log for the clinoform area (well D in veloped. The exposure surface becomes progressively
Figures 8, 11). Well is in downdip position of sequence 4, younger toward the basin in response to the down-
highlighting the development of the porous HST. Vertical stepping geometry (Figure 9).
scale is in 10-ft (3-m) depth increments. Porosity values Sequences 4 and 5 have very similar architecture,
below 5% are shaded in gray. The 1-md permeability line
facies, and reservoir quality. Each sequence includes
is highlighted in red. See Figure 13 for key to facies codes
and symbols. HST = highstand systems tract; TST = a muddy, lowstand-to-transgressive phase (low reser-
transgressive systems tract; SB = sequence boundary; voir quality) with low-angle clinoforms, followed by
MFS = maximum flooding surface. a grainy, highstand phase (moderate to high reservoir
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 195

quality) with higher angle clinoforms (Figures 10, 11). facies, to 0.1–1 md in the downdip mud-prone facies
Facies relationships in sequences 4 and 5 are sum- of the lower slope. The dominant pore type in lower-
marized in Figure 16. slope facies is microporosity. Microporosity develop-
ment may have been enhanced through meteoric dia-
Sequences 4 and 5 (TST) genesis, also in association with subaerial exposure.
The TST is dominated by muddy, nonporous facies
that create extensive barriers to fluid flow. The facies Lowstand Phase
model for the TST includes a downdip transition The maximum lowstand of sea level is marked by a
from (1) caprotinid floatstone, bafflestone, and rud- significant influx of fine-grained siliciclastics and the
stone at the margin (Figure 16D) to (2) mixed-peloidal demise of the Aptian carbonate platform. Lowstand
and skeletal packstone and grainstone in the upper- facies form a basin-restricted wedge (sequence 6) that
most slope (Figure 16E) to (3) burrowed wackestone onlaps the Aptian platform margin (Figures 10, 11)
(denses) in the upper- and lower-slope environments and forms a lateral seal to the reservoir. The lowstand
(Figure 16F). The TST is dominated by thick, mud- wedge comprises a shoaling- and cleaning-upward se-
prone facies deposited in the upper- and middle- quence that includes an upward succession from fine-
slope environments. As will be discussed below, these grained siliciclastics and carbonate mud to skeletal and
muds are interpreted to have been generated in situ oolitic carbonate grainstone. Sequence 6 is bounded at
as mud mounds. the top by an exposure surface that precedes deposition
of the overlying Nahr Umr Shale. In the study, only the
Sequence 4 and 5 (HST) updip part of sequence 6 is present. Regional relation-
The HST is dominated by grainy, porous clino- ships indicate that sequence 6, corresponding in large
forms alternating with higher frequency (fourth- part to the Bab Member of the Shuaiba, continues to
order) dense intervals. Clinoforms of the HST exhibit prograde into the Bab basin. The Nahr Umr Shale re-
higher slope angles (2 – 38) than the muddy TST cli- cords a rise in sea level that floods back across the pre-
noforms (1 – 28). A typical facies progression in the viously exposed Aptian platform. However, in contrast
HST includes rudist rudstone at the shelf margin to carbonate platform facies below, the marine envi-
(Figure 16A), skeletal grainstone in the upper slope, ronment was dominated by fine-grained siliciclastics,
peloidal packstone and grainstone in the middle with only thin intervals of argillaceous carbonates. The
slope (Figure 16B), and foraminifera-packstone in the Nahr Umr forms a top seal for the Shuaiba reservoir.
lower-slope environment (Figure 16C).
VOLUME-BASED RESERVOIR EVALUATION
Reservoir Quality
Reservoir quality varies dramatically between the The sequence-stratigraphic framework and 3-D
TST and HST components of the slope sequences and seismic data are used to guide an integrated evalua-
along the depositional profile (Figures 16, 19, 20). tion of the Aptian reservoir. Volume interpretation
The mud-dominated TST of each sequence is dense and visualization workflows are applied to further
and forms an effective and laterally continuous bar- evaluate and quantify geologic variations and their
rier. Grainier facies in the updipmost part of the TST impact on reservoir performance. A corendered im-
exhibit porosity that ranges from 10 to 20% and per- age of seismic discontinuity and seismic porosity is
meability ranging from 10 to 50 md, thus providing shown in Figure 21 and provides a seismic-scale per-
for possible updip connectivity between the slope se- spective of reservoir architecture and properties.
quences. The grain-dominated HST of each sequence Field A1 is extremely heterogeneous and can be
has much higher reservoir quality, although the in- subdivided into three main areas, each with diag-
ternal architecture of the HST clinoforms is complex, nostic geologic, seismic, and production expressions.
resulting in a high degree of heterogeneity. The fourth- These areas include the platform interior to the south,
order cyclicity generates a high-frequency alternation the northern platform margin in the central field
of flow units and flow barriers. Reservoir quality var- area, and the clinoform belt to the north. Geologic
ies systematically down the profile of each clinoform controls that drive these variations are directly related
in response to pore-type changes. Porosity values to the sequence-stratigraphic framework as discussed
are fairly consistent (15–25%) down the profile, but above and summarized in Figure 8.
permeability varies significantly, ranging from the A peripheral waterflood was implemented during
tens to hundreds of millidarcys in grain-prone updip earlier stages of field development (prior to acquisition
196 / Yose et al.

FIGURE 21. Volume-based reservoir evaluation. The volume shown is total porosity corendered with seismic discontinuity.
The volume highlights the reservoir heterogeneity resulting from sequence-stratigraphic variations and the impact on
field development. Time slice is from a flattened volume datumed on the top of sequence 1 (base of reservoir) and
represents a slice through the TST of sequence 3 in the platform interior and margin, cutting through sequences 4, 5,
and 6 in the clinoform belt and the Bab basin. The total porosity volume includes the seismic-derived porosity (band-
limited porosity) plus the low-frequency porosity trend. High-frequency porosity data from core and well data are added
during the 3-D geologic modeling stage.

of 3-D seismic data) and does not adequately accom- This interval contains between 4 and 5 billion bbl of
modate the reservoir heterogeneity that is now ap- in-place oil reserves. Thus, a small improvement in
parent. Several field management challenges have recovery will have a large economic impact. Under-
developed, including differential sweep in the platform standing the origin and distribution of reservoir
interior, early water encroachment along the platform- quality anomalies in the platform area has pre-
margin grainstone belt, and pressure support in the sented a long-standing field development challenge,
clinoforms (Figure 21). The ADCO has implemented both in terms of avoiding these anomalies with pro-
several field management strategies to further opti- ducer and injector wells and the impact of these fea-
mize field development. As demonstrated below, the tures on reservoir sweep and flood-front advance
3-D seismic data will greatly assist in guiding future (Figure 21).
well placement and field optimization strategy.
Calibration with Well and Core Data
Seismic Characterization of the The 3-D seismic data provide a blueprint of the
Platform Interior channel and pond network in the platform interior
The platform interior is characterized by a com- (Figures 6, 21) and a new tool for reservoir evaluation.
plex spatial distribution of porosity that is related to In our initial reconnaissance of the seismic data,
the channel, pond, and rudist shoal complex devel- seismic features were identified within the platform
oped during the early highstand phase (Figure 21). interior that were interpreted either as karst features
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 197

FIGURE 22. Integrated characterization of the platform interior region. (A) Reservoir isochron corendered with seismic
discontinuity (red = isochron thick; blue = isochron thin). Blue well sticks are water injector wells, and red well sticks
are selected producers. Ponds and channels correspond to isochron thins, and porous rudist shoals correspond to
isochron thicks. Prominent isochron thick to the north is the northern platform margin. Peripheral water injection is
diverted by ponds and channels resulting in differential sweep. (B) Example of high-porosity, high-permeability rudist
shoal facies developed among the ponds and channels. (C) Example of nonporous submarine pond fill (open-marine
skeletal wackestones and packstones with rudist debris). Por = porosity; Perm = permeability; md = milidarcy.

developed in association with the top Shuaiba super- thins correspond to dense, nonporous carbonate with
sequence boundary or as syndepositional features, fast seismic velocities, whereas isochron thicks corre-
such as tidal channels and submarine ponds. As sum- spond to porous carbonates with slower seismic ve-
marized in Figure 22, calibration of the seismic data locities. These observations are consistent with the
with well data supports a syndepositional origin for seismic porosity predictions shown in Figure 21, con-
the seismic features. Volume corendering of the seis- firming that the pond and channel complex is gen-
mic discontinuity volume with the reservoir isochron erally of lower reservoir quality than the surrounding,
(time thickness) enhanced the detail of stratigraphic more continuous seismic facies.
features in the platform interior (Figures 6, 22) and Cores from within the pond features comprise
guided the selection of calibration wells. Calibration largely carbonate mudstone and wackestone that
of the seismic data with log data shows that isochron was deposited in open- to restricted-marine settings
198 / Yose et al.

(Figure 22). A range of carbonate facies was observed In addition, reservoir quality tends to increase pro-
within the ponds, including open-marine carbonate gressively from the pond interior to the pond margins.
mudstone; wackestone; peloidal-skeletal packstone; Less data are available to assess facies and reservoir
and restricted-marine, organic-rich mudstone and quality variations in the large tidal channels, but seis-
algal laminites. In contrast, cores from wells located mic porosity indicates that, in general, the channels
outside the ponds comprise rudist-dominated grain- are filled with higher porosity facies than the ponds.
stone, floatstone, and rudstone that were deposited Visualization sessions provide a powerful mecha-
in high-energy carbonate shoals and patch reefs. The nism for evaluating reservoir sweep and conformance
open-marine facies in the pond fill contain rudists and for identification of bypassed reserves in the
and other sediments that are interpreted to have platform interior. Visualization of seismic volumes
been transported into the ponds from the adjacent such as those shown in Figures 21 and 22 enabled
rudist shoals. No evidence was found to indicate that reservoir engineers to rapidly resolve long-standing
karsting or dissolution was involved in the formation production anomalies. For example, many producers
of the pond and channel features. Collapse breccias, with abnormally high or low water cuts can be ex-
fractures, vadose silts, and other diagenetic features plained by their position relative to injector wells and
typical of karst environments are absent. the pond-channel complex (Figure 22). Producers
The ponds and channels appear to be time equiv- that are sheltered from injectors by ponds and chan-
alent to the surrounding rudist shoals and patch reefs. nels show less water breakthrough (higher oil cuts)
The initial distribution of the rudist shoals and ponds than producers that are not as sheltered. The 3-D
may have been controlled by antecedent relief on the seismic data provide new, detailed information on
underlying sequence boundary (top of sequence 2), the distribution of reservoir features that is being used
with shoals nucleating on subtle highs. The initial to guide reservoir development and placement of
relief was then enhanced through aggradation of the future producer and injector wells.
shoals in response to sea level rise during the trans-
gressive phase of sequence 3. This increase in accom-
modation resulted in the differentiation of the south- Seismic Characterization of the Platform Margin
ern platform margin, thus opening the platform top The northern platform-margin area is character-
up to strong cross-bank currents and development of ized by the highest and most continuous porosity in
submarine channels. As the rate of accommodation the field area (Figures 18, 21). Production data indi-
began to decrease, the rudist shoals coalesced across cate that premature water breakthrough is occurring
the bank top, thus infilling most of the pond and along the platform margin in response to peripheral
channel topography (Figure 8). The HST of sequence water injection. The seismic data reveal possible strat-
3 extends across the tidal channels, ponds, and shoals igraphic and structural controls on fluid flow in the
of the underlying TST, with little change in facies platform-margin area.
(Figures 8, 9). This relationship indicates that the
channels and ponds were developed and infilled
during the TST of sequence 3 and are not related to the Calibration with Well and Core Data
long-term exposure event on the top of the Shuaiba Seismic and sequence-stratigraphic relationships
(Lower Cretaceous supersequence boundary). of the platform-margin area are summarized in
Figure 23. The seismic character of the margin area
is generally transparent because of the stack of porous
Reservoir Implications and relatively continuous reservoir facies. Discontinu-
Seismic porosity distribution in the channel-pond ous basinward-dipping seismic reflectors (clinoforms)
network shows that, although porosity is generally are present in the HST of sequence 3. Calibration with
lower than the surrounding rudist shoal facies, there core data show that platform-margin facies is comprised
is considerable variability, and not all of the vol- of mainly skeletal and nonskeletal grainstone that is
ume is nonreservoir (Figure 21). Our calibrations remarkably fine grained and well sorted (Figures 15E,
with core and well data indicate that porosity varies F; 16B, C; 23C). The carbonate sand shoals aggraded
with respect to several factors, including the size of vertically and then, as the rate of sea level rise began
the ponds and position of wells in the ponds. Larger to decrease, prograded basinward. As a result, facies
ponds are more restricted from marine currents than transition upward from a grainy platform margin into
smaller ponds and are filled with muddier, tighter rock. rudist-dominated outer platform facies (Figure 23C).
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 199

FIGURE 23. Seismic and depositional relationships in northern platform margin area of the reservoir. Depositional sequence
shown in (C) is sequence 3 in Figure 8. The vertical stacking of facies in the platform margin records progradation, with
open-platform interior facies prograding over platform-margin facies. Reservoir quality in the margin area is generally
high and relatively continuous. (A) Seismic transect through northern platform-margin area (amplitude data). (B) Seismic
transect through northern platform-margin area (impedance data; orange = lower porosity, gray = higher porosity).
(C) Facies and reservoir-quality relationships in the platform-interior, platform-margin to upper-slope environments.
BP = between-particle porosity; IP = intraparticle porosity; micro = microporosity; Por = porosity; Perm = permeability;
HST = highstand systems tract; TST = transgressive systems tract; SB = sequence boundary. See Figure 13 for explanation
of facies codes.
200 / Yose et al.

FIGURE 24. Three-dimensional perspective on clinoform architecture, showing a time slice of seismic discontinuity
with 2-D seismic amplitude cross sections. Systematic variations in clinoform architecture are clearly imaged and are
related to the third-order sequence framework as shown. Colored well sticks show the location of wells that were
evaluated as part of the current study. Intersections of third-order sequence boundaries with the seismic time slice
are denoted by the dotted red lines.

Reservoir Implications High-porosity grainstone in the platform-margin tran-


The data indicate that the northern platform mar- sitions into the more variable reservoir quality facies
gin is a broad zone of stacked, moderate- to high- of the platform interior. The platform margin and in-
reservoir-quality facies. Seismic data show that this terior areas are thus in pressure communication, and
trend is continuous across the field area and well con- injected water moving along the strike of the platform
nected to the peripheral water injectors (Figure 21), margin can also interfinger laterally into higher per-
consistent with the observed water breakthrough pat- meability facies in the platform interior. Water break-
terns. However, as previously discussed, faults seen through has been observed in the proximal platform
in the seismic data trend parallel to the platform mar- interior area and may increase over time.
gin and may influence fluid flow and water break-
through patterns (Figure 7). Sensitivity testing can be
conducted via flow simulation to determine if ma- Seismic Characterization of the Clinoforms
trix properties can match the performance data or Slope clinoforms are characterized by complex
whether additional permeability associated with faults stratal geometries and significant variations in res-
and fractures is required. The 3-D fault framework ervoir properties (sequences 4 and 5; Figure 8). As a
derived from seismic data (Figure 7) can be leveraged consequence, peripheral water injection has not pro-
in the sensitivity testing. vided adequate pressure support or sweep to this part
The seismic and sequence-stratigraphic frame- of the reservoir (Figure 21). In response, ADCO has
works also provide an improved understanding of lat- implemented a pattern gas-injection flood to add pres-
eral facies relationships and reservoir connectivity in sure support and optimize recovery. As illustrated in
a dip direction across the platform margin (Figure 23). Figure 24, seismic data provide new 3-D information
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 201

FIGURE 25. Seismic and depositional anatomy of clinoform sequences. (A) Seismic amplitude section across the entire
clinoform belt, highlighting seismic expression of sequences 4 and 5 (TST = transgressive systems tract; HST = highstand
systems tract). (B) Close-up of sequence 4 with seismic interpretation added. Well sticks in black are featured in (C).
(SB = third-order sequence boundaries; MFS = third-order maximum flooding surface; fourth-order surfaces shown with
blue double arrows). (C) Facies schematic for sequence 4 showing impact of clinoforms on flow layering and well pair
communication.

on stratal architecture and reservoir quality variations by finer scale alternations of porous and dense facies
that can be used to assess reservoir connectivity and that result from higher order depositional cyclicity
optimize gas flood management. (Figure 25).

Calibration with Well and Core Data Reservoir Implications


The detailed anatomy of a slope-restricted deposi- Reservoir connectivity and quality variations in
tional sequence is provided in Figure 25. The seismic the clinoforms are summarized in Figure 25. Reser-
character of each sequence includes a low-angle (1 – voir implications are discussed below in the context
28), high-amplitude reflector, followed by a series of of the sequence hierarchy. At the sequence scale,
shorter, steeper (3– 48) reflectors. Cores were evalu- transgressive dense intervals at the base of each de-
ated across several slope sequence transects to cali- positional sequence form effective barriers to fluid
brate the seismic response (see Figure 24 for location flow and partition the clinoform area into separate
of calibration wells). The low-angle, high-amplitude pressure compartments (Figure 8). These dense inter-
reflector at the base of the cycle is generated by the vals are well imaged by the seismic data, and 3-D ge-
impedance contrast between more porous rock below ometries of the reservoir barriers can be mapped across
and extremely dense rock above. In the sequence mod- the northern field area (Figures 25, 26). Highstand
el, this interface is interpreted as a sequence boundary parts of sequences comprise larger scale flow com-
separating more porous highstand deposits below, partments and contain the bulk of the pore volume
from muddier, transgressive deposits above. Dense, in the clinoform region.
muddy, lowstand-to-transgressive deposits overlie the The highstand flow compartments are, in turn,
sequence boundary and form the base of each depo- subdivided into a series of alternating flow units and
sitional sequence. Higher angle reflectors in the high- discontinuous flow barriers, corresponding to the de-
stand part of each depositional sequence are generated velopment of fourth-order cyclicity. As illustrated
202 / Yose et al.

FIGURE 26. Three-dimensional visualization of clinoforms highlighting reservoir connectivity and continuity. Clinoform
geometries have a large impact on connectivity between injectors (yellow) and producers (red). Volume shown is
instantaneous phase with opacity filter applied to enhance visualization of seismic-scale flow units and flow barriers.
Basal structural surface is draped with seismic porosity and discontinuity from within the reservoir.

in Figures 25 and 26, seismic reflection data are able to the clinoforms in relation to existing producer and
resolve many of these fourth-order clinoform pack- injector wells. Such tools are used to assess 3-D res-
ages and provide a template to guide interwell cor- ervoir connectivity, to evaluate producer-injector con-
relations. Without seismic data, one may be tempted formance, to resolve specific well-pair performance
to correlate on the basis of facies and log similarities, issues, and to guide future well placement.
resulting in horizontal, instead of inclined, correlation
surfaces (i.e., connecting the colors in Figure 25C).
Such correlations crosscut seismic reflectors and geo- IMPLICATIONS FOR RESERVOIR MODELING
logic time lines and result in incorrect flow layering
and rock property distributions in the reservoir model. Three-dimensional geologic and flow-simulation
Integration of the seismic with core and well data models provide the ultimate reservoir management
provides information on subseismic variations in tools. The inherent complexity of carbonate reser-
reservoir properties. Depositional facies change sys- voirs presents significant challenges to distributing
tematically down the profile of each clinoform be- 3-D rock properties in reservoir models. The new
cause of changes in water depth and wave energy. volume-based reservoir framework provides a tem-
Grain-rich, porous facies deposited in updip posi- plate for distributing flow units and flow barriers in
tions pass gradually into finer grained, less-porous the reservoir and for distributing facies and rock prop-
facies deposited in downdip positions. These facies erties in the 3-D framework. Likewise, the 3-D seismic
changes, coupled with diagenetic effects, result in data constrain the structural mapping of the reservoir
systematic pore-type changes that control porosity and have provided a 3-D fault framework that can be
and permeability variations in individual flow units incorporated into the reservoir model framework.
(Figure 25C). An integrated static and dynamic reservoir mod-
The geometry and continuity of the flow units, eling effort that incorporates results from the pres-
barriers, and baffles will have a pronounced influ- ent study is underway, but was not concluded at the
ence on injector-producer well connectivity in the time of this publication. Some examples are provid-
pattern gas flood area. As illustrated in Figure 26, 3-D ed below to show how a volume-based workflow can
seismic visualization is used to evaluate the distribu- lead to more accurate reservoir models and can also
tion of seismic-scale flow units and flow barriers in reduce the reservoir-modeling cycle time.
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 203

FIGURE 27. Seed-detected clinoforms extracted from 3-D seismic volume. Colors correspond to connected clinoform
segments. Volume shown is the cosine of instantaneous phase, which enhances continuity of the seismic reflections.

Rapid Interpretation Techniques Volume-based Data Integration


Development of the reservoir framework is one of Most conventional 3-D modeling workflows focus
the most time-consuming steps in the construction on the integration of one-dimensional and two-
of a 3-D geologic model. The volume-interpretation dimensional (2-D) data sets, and a full 3-D represen-
techniques applied in the present study reduced the tation of the reservoir is not developed until the later
time required to construct the structural and strati- stages of the workflow, following the construction of
graphic frameworks. For example, seismic mapping the 3-D geologic model. A volume-based workflow
of individual clinoform surfaces in the northern field provides the basis for 3-D reservoir characterization
area is critical for the reservoir description, but is time throughout the workflow. Figure 28 illustrates how
consuming because of the stratigraphic complexity. the volume-based environment is used to visualize
Quantitative seed detection techniques were used to information on structure, stratal architecture, reser-
expedite the correlation process. The ExxonMobil voir properties, and static and dynamic well data in a
proprietary seed detection approach applied in this common 3-D environment. Such techniques are used
study uses a combination of amplitude and trace- to evaluate relationships between seismic and other
shape attributes to autopick the clinoform surfaces. geologic and production data and to identify key
Results are illustrated in Figure 27. The colored sur- reservoir issues early in the workflow. Volume-based
faces are autopicked clinoform surfaces, and each data integration leads to more accurate reservoir mod-
color corresponds to a connected clinoform surface. els and promotes a 3-D reservoir perspective through-
This technique is rapid (interpretations were gener- out the workflow.
ated in a few hours) and allows for quantitative as-
sessment of clinoform connectivity. The fast-track Extraction of Reservoir Features from Seismic Data
interpretations must be further evaluated and opti- As illustrated in Figure 29, seismic features in the
mized by the seismic interpreter. Ultimately, the seis- platform interior can be extracted from the seismic
mic clinoform surfaces are depth converted and used volume and incorporated directly into the 3-D geo-
to guide the correlation of higher order surfaces iden- logic model. In this workflow, geologic and seismic
tified from logs and core data. attributes are used to guide the extraction process.
204 / Yose et al.

FIGURE 28. Seismic image highlighting variations in reservoir architecture and quality at the full-field scale. Amplitude
data are shown in 2-D vertical slice. Discontinuity draped on time structure and corendered with band-limited seismic
porosity predictions shown in time slice view. Selected calibration wells are shown as yellow sticks. Major paleographic
elements are annotated.

The extraction was constrained vertically using seis- combination with well log and core information to
mic horizons tied to the sequence-stratigraphic frame- produce the final 3-D porosity distribution in the
work. Integration of the seismic with well and core geologic model.
data indicate that the pond-channel network is con-
fined to a specific sequence-stratigraphic interval Rock Property Modeling
(sequence 3, HST; Figure 9), and this knowledge is Accurate representation of rock properties in res-
used to guide the extraction of these features and ervoir models is challenging because of the different
their placement into the 3-D geologic model. The scales of information on rock properties (seismic, well
lateral extent of individual ponds and channels is tests, logs, and core) and the impact of differing pore
constrained using a combination of seismic attri- types on the porosity-permeability functions. Rock
butes, including discontinuity, isochron, and poros- property modeling requires a combined approach,
ity. Once constrained vertically and laterally, the fea- using 3-D seismic data and the sequence-stratigraphic
tures can be extracted from any seismic volume, framework to guide large-scale trends between wells,
along with the seismic properties. Figure 29B and C and high-resolution information from log and core
show 2-D and 3-D perspectives of the pond and data to characterize pore-scale relationships.
channel features extracted from the seismic porosity
volume. The extracted features, populated with seis- Integration of Seismic Porosity Data
mic porosity estimates, were converted from time to The average well spacing in Field A1 is 1 km (0.6 mi),
depth and incorporated directly into the 3-D geo- and seismic data provides the only 3-D information on
logic model. Seismic porosity predictions are used in rock properties between wells. Several seismic attributes
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 205

were calibrated to log-derived porosity data using a


neural network approach (Hampson et al., 2001) and
used to generate a full-field prediction. A more de-
tailed explanation of the seismic porosity prediction
workflow conducted in association with this study
is provided by Al-Menhali et al. (2005) and Schuelke
et al. (2005). Most of the seismic porosity informa-
tion is contained in the impedance attribute, but
adding additional attributes via the neural net ap-
proach improved the calibration. Seismic porosity
predictions are validated through blind tests, where
wells are held out of the calibration process and then
compared against the predictions to characterize
uncertainty. Visualization techniques are then used
to evaluate the seismic porosity predictions relative
to the sequence-stratigraphic framework and perfor-
mance data. Statistical error in the seismic prediction
is generally low but increases in areas where rock
properties change rapidly over short vertical or lat-
eral distances (e.g., clinoforms of sequences 4 and 5,
and tidal channels and ponds in sequence 3). As il-
lustrated in Figures 21 and 28, there is a strong tie
between seismic-based porosity predictions and the
geologic framework. The seismic porosity volume
provides a valuable reservoir evaluation tool and will
help to constrain porosity trends in the 3-D geologic
model.

Porosity and Permeability Predictions


Prediction of porosity-permeability relationships
requires understanding pore-type distributions,
which, in turn, define reservoir rock types. Reservoir
rock types reflect a combination of diagenetic and
depositional processes, and depending on the level of
diagenesis, rock types may not equate to depositional
facies. In such cases, the diagenetic component must
be modeled separately. Whereas a considerable dia-
genetic overprint does exist in the Shuaiba reservoir,
most of the diagenesis is facies selective and related
to sequence-stratigraphic surfaces (e.g., exposure

FIGURE 29. Seismic extraction of platform interior pond


and channel features for input into the 3-D geologic
model. (A) Boundaries of detected ponds and channels
shown on discontinuity volume corendered with reservoir
isochron (blue = thin). The combination of these two
seismic attributes provides the best lateral constraint on
these features. (B) Time slice view of pond and channel
features extracted from the seismic porosity volume
(blue = low porosity; red = high porosity). (C) Three-
dimensional perspective of pond and channel complex
extracted from seismic porosity volume (blue = low
porosity; red = high porosity).
206 / Yose et al.

FIGURE 30. Porosity-permeability crossplot for sequences 4 and 5. Depositional facies provide a close proxy for reservoir
rock types in this system. Porosity-permeability relationships vary as a function of sequence, systems tract, and depositional
environment and are thus predictable with the sequence framework. Dominant pore types are indicated: BP = between-
particle porosity; IP = intraparticle porosity; MO = moldic porosity; VUG = vuggy porosity.

surfaces). Thus, depositional facies, in combination in the clinoform area is how connected the clino-
with the sequence-stratigraphic framework, provide forms are in the updip area, below the top Shuaiba
a good understanding of reservoir rock-type distri- supersequence boundary (Figures 8, 31). A clear com-
bution. Figure 30 illustrates the value of using res- partmentalization exists between the TST and HST
ervoir rock types and the sequence-stratigraphic clinoform intervals downdip, but updip, relationships
framework to guide porosity-permeability model- are not clear. Only a few wells penetrate the narrow
ing in the clinoform area. Reservoir rock types and updip extension of the TST intervals, so there are
their properties vary predictably based on systems limited data. Reconciliation of this issue will require
tract (TST and HST) and position along the deposi- feedback between the geologic and flow-simulation
tional profile (updip and downdip) (Figure 30). models through integration of production and per-
Preliminary 3-D modeling results from the clino- formance data.
form area are shown in Figure 31. The flow lines in
the model correspond to sequence-stratigraphic sur-
faces and represent geologic time lines. Facies changes LEARNINGS ON CARBONATE SYSTEMS
that occur along and across these time lines have
a significant impact on rock property distribution. The volume-based approach used to characterize
Note the strong conformance between the sequence- the Shuaiba reservoir provides details on carbonate
stratigraphic framework of the clinoforms (Figure 8), system development that are sparsely observed in
the seismic porosity response (Figures 21, 28), and outcrop or subsurface studies. Some highlights and
the 3-D geologic model (Figure 31). A key uncertainty key learnings are discussed below.
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 207

FIGURE 31. Paired porosity and permeability extracts from the 3-D geologic model in the clinoform area along the
northern platform margin. HST = highstand systems tract; TST = transgressive systems tract.

Controls on Platform Architecture 4 and 5. The southern margin differentiated mainly


and Evolution during the time of sequence 3 and is characterized by
Integrated characterization of the Shuaiba reservoir high-energy rudist shoals that occupy a broad area of
reveals a very ordered system, where changes in stratal the margin and outer platform (7 –9 km; 4.4 – 5.6 mi)
geometries and facies distribution are tied closely to (Figure 9). The southern margin is mainly aggrada-
the long-term accommodation history (Figure 9). tional, and sequences 4 and 5 appear to be repre-
However, despite the order and predictability im- sented as restricted, low-energy mudstone deposited
parted by changes in relative sea level, there is vari- in the intrashelf embayment (Figure 9).
ability in platform development that may be related Carbonate platform asymmetry is common in many
to other competing factors, such as local tectonics, ancient and modern carbonate systems and may de-
sediment supply, and cross-bank energy flux. velop in response to a variety of controls, including
differential subsidence, sediment supply, physical
energy flux (e.g., currents, waves), and interference
Platform Asymmetry patterns between isolated platforms (Yose and Col-
The Shuaiba platform records a marked asymme- lins, 2002). Two primary controls on the asymmetry
try in depositional facies and stratal geometries from observed within the Shuaiba buildup are interpreted.
north to south across the study area. Despite a similar
accommodation history, the northern and southern 1) Physical energy flux: At the time of maximum
margins are very different in terms of geometry and extent of the Shuaiba platform (sequence 3), we
facies (Figure 9). The northern margin is character- interpret the southern margin as the current-
ized by the development of a broad (10-km; 6.25-mi), facing margin and the northern margin as the
grain-dominated sand belt that persisted during leeward-facing margin. The current-facing mar-
most of the time of sequences 2 and 3, followed by gin was the optimum site for proliferation of the
strong progradation during deposition of sequences rudists. Facies comprising rudist rudstone are
208 / Yose et al.

common along the southern margin, indicating the basis of seismic time slices because the shingled
strong wave and current activity. Large tidal chan- HSTs of sequences 4 and 5 are merged (Figure 24).
nels imaged by the seismic data all appear to enter These relationships demonstrate the variability
the platform from the south and then meander that can occur along the strike of carbonate platform
their way into lower energy areas of the platform margins. The changes in clinoform geometry could
interior, connecting to many of the submarine result from along-strike variations in sediment sup-
pond features (Figures 6, 9). In contrast, the lee- ply or subsidence patterns. The offstructure decrease
ward margin is characterized by the deposition of in the width of the clinoform belt indicates the pos-
fine-grained skeletal-peloidal grainstone and sibility of an early structural influence on sedimen-
packstone that grade into mud-dominated slope tation. The primary structural movement of the large
facies (Figure 8). The sands and muds along the north-northeast structures on the Arabian plate is in-
northern margin are interpreted to record offbank terpreted to have been in the Late Cretaceous ( Johnson
transport of fine-grained carbonate sediment along et al., 2005). However, even a slight increase in sub-
the leeward margin. The tidal-channel complex sidence rates offstructure would generate increased
facilitated cross-bank transport of sediments to- accommodation space and could account for the in-
ward the leeward margin. The interplay of pre- crease in slope angles and the decrease in the distance
vailing winds, currents, and paleogeography in of progradation observed in the offstructure posi-
controlling the observed patterns in cross-bank tions. Conversely, the clinoform geometries could
energy flux is not clear at present. have been produced by variations in sediment supply,
2) Position of sea level: The asymmetry developed with areas of higher progradation corresponding
during deposition of sequences 4 and 5 results to higher sediment supply. The apparent pinch and
from strong progradation along the leeward mar- swell of individual clinoforms observed in seismic
gin and virtually no progradation to the south data are interpreted to record lobate geometries re-
(Figures 9, 10). These relationships are consistent sulting from variations in erosion and sediment sup-
with the cross-bank energy flux described above. ply. These variations could also account for geometric
Another factor, however, is that the long-term variations at the larger scale.
fall in sea level may have significantly restricted
circulation patterns in the intrashelf embayment Platform Response to Sea Level Fall
to the south (Figure 4). Connections from the The long-term fall in sea level during the Aptian
larger Bab basin into the embayment may have provides an opportunity to evaluate the response of a
been restricted as sea level lowered. In contrast, the carbonate system to sea level fall, including perspec-
northern margin faced into the Bab basin, where tives on stratal architecture, sedimentation patterns,
open circulation was maintained, and rudist fring- and diagenesis.
ing shoals developed in downslope positions.
Downslope Shifts of the Carbonate Factory
The Aptian platform demonstrates the potential
Variations in Clinoform Geometry for carbonate systems to shift downslope in response
Variations in the geometry and sedimentology of to sea level fall. Sequences 4 and 5 are interpreted to
the slope clinoform sequences can be observed at a have accumulated as slope-restricted wedges that are
variety of scales and illustrate a range of possible detached from the main platform (Figure 8). Seismic
controls. At the largest scale, variations in the lateral (time) and log (depth) relationships confirm that
width of the clinoform belt show a close correspon- the clinoform sequences are downstepping into the
dence to the present-day structure. The clinoform Bab basin, tracking the second-order sea level fall. In
belt is widest (more progradation) along the crest of the literature, these relationships are referred to as
the structure and narrows offstructure (east and ‘‘forced regressions,’’ in that sedimentation is being
west) (Figure 24). The offstructure thinning of the forced basinward (downslope) in response to the sea
clinoform belt is accommodated by a combination level fall (Hunt and Tucker, 1992). Younger sequences
of increased slope angles to the east and west and a onlap the margin of older sequences, and the com-
pronounced narrowing of sequence 5 to the east, posite unconformity becomes progressively younger
compensated primarily through thinning of the TST toward the basin. Facies relationships from core in-
(Figures 9, 24). On the eastern side of the field, it be- dicate little to no erosion along the tops of the cli-
comes difficult to differentiate sequences 4 and 5 on noforms; a normal facies progression is observed up to
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 209

the sequence boundary in each clinoform sequence. phase, the rate of accommodation begins to slow,
Thus, the basinward thinning of the clinoform pack- and higher energy shoals initiate along the newly
age is not caused by erosion. The low-angle (2–38) developed platform margin and then prograde basin-
slopes flanking the buildup provided a large area for ward (Figure 25A, B).
in-situ carbonate production, and the relative de- The source of the mud in the transgressive interval
crease in accommodation during the long-term sea is enigmatic because the adjacent bank top is in-
level fall allowed for extensive progradation of the terpreted to be exposed during the deposition of the
carbonate system into the shallow basin. slope sequences. The carbonate factory was thus re-
Forced regressions, as observed within the Aptian stricted to the slope. The transgressive muds thicken
sequence set, raise questions on the placement of upslope, forming a basin-thinning wedge. Based on
sequence boundaries and on systems tract nomen- evaluation of core, we interpret that the upper- to
clature in the slope sequences. In the Shuaiba ex- middle-slope environment was an area of in-situ car-
ample, the Aptian sequence set boundary could be bonate sediment production. As the carbonate sys-
placed below sequences 4 and 5, marking the ini- tem was catching up with rising sea level after the sea
tiation of long-term exposure of the central Shuaiba level fall, the main type of sediment produced was
platform. In this interpretation, sequences 4 and 5 carbonate mud. The source of the muds may have
would be considered as part of the composite low- been sea grass and algal meadows developed in the
stand interval. Our preferred interpretation is to place upper- to middle-slope environments. The dramatic
the sequence set boundary at the base of sequence 6. upslope thickening of denses supports the concept that
In this interpretation, sequences 4 and 5 become part the carbonate muds were produced locally (Figure 9).
of the late highstand interval or, alternatively, the The presence and abundance of miliolid foraminif-
falling-stage systems tract of Plint and Nummedal era, which lived attached to sea grasses, further sup-
(2000). Sequence 6 is viewed as the lowstand phase ports the interpretation that sea grass may be an
of the next sequence set and records a major change important factor in both the baffling and production
in sedimentation patterns, including a large influx of carbonate mud. Hydrodynamic ponding of muds
of siliciclastics. Issues around nomenclature have no along the upper-slope environment may also be a
impact on the fundamental sequence interpretations factor.
or the reservoir framework. However, for regional cor-
relations, it is important to have a clear and consis- Impact of Long-term Subaerial Exposure
tent nomenclature. The supersequence boundary that developed on
the top of the Aptian sequence set resulted in long-
Mechanisms of Progradation term exposure of the carbonate platform, providing
The Shuaiba platform provides new insights on an opportunity to assess the impact of long-term ex-
mechanisms of carbonate progradation in response posure on the carbonate system. The exposure sur-
to sea level changes. Each clinoform slope sequence face is time transgressive and becomes progressively
(sequences 4 and 5) includes a low-angle, mud- younger toward the basin because of the progressive
dominated base (TST), followed by deposition of downslope shift of the carbonate system (Figure 9).
higher angle grainy clinoforms (HST) (Figures 9, 10). The time span of the exposure surface is estimated
Each sequence records on the order of 5 km (3.1 mi) of to range from 3 – 4 m.y. in the platform interior to
progradation, resulting in more than 10 km (6.2 mi) 1 – 2 m.y. in sequence 5.
of total progradation (Figures 24, 25). In the central Based on observations from core and seismic data,
part of the clinoform belt, partitioning of prograda- there is no evidence for karsting of the platform in
tion between the TST and HST intervals is roughly response to the long-term sea level fall. Exposure sur-
equal, but the mechanisms of progradation are dif- faces are marked by thin (2–6-cm; 0.8–2.4-in.) crusts
ferent. A two-phase style of progradation is proposed. of iron mineralization, with only minor evidence for
The TST phase includes aggradation of carbonate dissolution (Figure 12A, B). Although evidence for
mud banks that nucleate downslope in response to karsting is absent, long-term exposure of the carbon-
sea level fall and then aggrade vertically during the ate platform is interpreted to have had a significant
subsequent transgression. Aggradation of the mud impact on porosity development near sequence bound-
banks creates new depositional relief and effectively aries and in some sequences. Blocky calcite cements
steps the margin 2 – 3 km (1.2 – 1.9 mi) outboard rela- are common below surfaces and can extend down for
tive to the previous highstand. During the HST several feet, creating tight intervals immediately below
210 / Yose et al.

sequence boundaries. Subaerial exposure of grain- tion is a useful tool to facilitate integration and
prone facies results in the development of abundant evaluation of a wide range of subsurface data and
vuggy and moldic porosity through dissolution of their relationships.
skeletal and nonskeletal grains. Selective dissolution 3) Use of multiple seismic attributes for structural
is observed to penetrate tens of feet below sequence and stratigraphic interpretations: A range of seis-
boundaries into the underlying sequence. Dissolu- mic attribute volumes should be evaluated to
tion enhances porosity in facies that have abundant facilitate interpretation of the seismic data. Dif-
primary interparticle porosity, leading to high poros- ferent attribute volumes provide different infor-
ity and permeability (Figure 16A). Abundant micro- mation and perspectives on the structural and
porosity is observed in algal platform facies of se- stratigraphic frameworks and distribution of rock
quence 1B (Figure 14A, B, E) and in the middle- to properties. Attribute volumes that were most use-
lower-slope facies of sequences 3–5 (Figures 15F, 16C). ful in this study included amplitude, discontinu-
Development of microporosity may have been en- ity, impedance, isochron, porosity, dip, instanta-
hanced by meteoric fluids. Mircroporosity results in neous phase, and quadrature. Attribute volumes
relatively high porosity but low permeability. can also be corendered in different combinations
to provide additional detail and information.
CONCLUSIONS 4) Evaluation of seismic data within an integrated
geologic framework: Application of high-end seis-
High-quality 3-D seismic data acquired over an mic technologies must be conducted within in-
onshore field in Abu Dhabi demonstrate the value of tegrated structural and stratigraphic frameworks.
seismic data for integrated reservoir characterization This is a continuous process because seismic data
and field optimization. The seismic data quality, abun- is used to develop geologic frameworks, and the
dance of core, well and production data, and the frameworks, in turn, are used to guide more de-
reservoir heterogeneity all combine to produce an tailed seismic attribute analyses, such as porosity
ideal data set to test the limits of high-end seismic prediction and fault and fracture analysis.
technologies in carbonates and to demonstrate the 5) Multidisciplinary approach: Accurate reservoir
impact on reservoir characterization. Seismic data, in characterization and evaluation requires a multi-
combination with sequence-stratigraphic concepts, disciplinary approach. Three-dimensional visuali-
provide a valuable framework for reservoir evalua- zation of seismic and other subsurface data fa-
tion and for incorporating flow layering and rock cilitates interactions among geoscientists and
property variations into reservoir models. engineers, leading to a collective understanding of
Key elements of and insights gained from the the reservoir and improved reservoir management.
workflows applied in the present study are summa- 6) Volume-based reservoir optimization: As dem-
rized below. onstrated in the present study, calibrated 3-D
seismic data can provide a powerful foundation
1) Optimizing the seismic data quality: The impact for reservoir evaluation and optimization. Visual-
of high-end seismic technologies increases as ization of seismic data, along with well and pro-
the data quality improves. Efforts on seismic duction data, can help to resolve longstanding
data optimization, including poststack filtering, reservoir performance issues, identify new op-
should occur early in the seismic workflow. The portunities, and guide future field development
present study demonstrates that poststack filter- strategies.
ing can improve the signal-to-noise ratio and
the interpretability of 3-D seismic surveys, even
those with high-effort seismic data acquisition ACKNOWLEDGMENTS
and prestack processing.
2) Calibration of seismic data to core, well, and pro- The authors thank the Abu Dhabi National Oil
duction data: Seismic data must be calibrated with Company, the Abu Dhabi Company for Onshore Oil
other subsurface data to groundtruth the seismic Operations, the ExxonMobil Exploration Company,
response to variations in geology and reservoir and the ExxonMobil Upstream Research Company
properties. A fundamental understanding of the for supporting this collaborative study and for per-
underlying geology is critical for carbonate res- mission to publish these results. We also thank our
ervoir characterization. Volume-based visualiza- colleagues for their contributions and valuable
Three-dimensional Characterization of a Heterogeneous Carbonate Reservoir, Abu Dhabi / 211

discussions, including Jim Anderson, Jim Markello, Hampson, D., J. S. Schuelke, and J. Quirein, 2001, Use of
Peter Holterhoff, Imelda Johnson, Po Tai, Nat Stevens, multi-attribute transforms to predict log properties
from seismic data: Geophysics, v. 66, no. 1, p. 220 –
Steve Bachtel, Brian Coffey, and Linda Corwin. The
236.
manuscript was improved by reviews from Jim
Hardenbol, J., J. Thierry, M. B. Farley, T. Jacquin, P. De
Weber, Scott Tinker, Tony Simo, Sherry Becker, and Graciansky, and P. R. Vail, 1998, Mesozoic and Ceno-
Mike Kozar. zoic sequence chronostratigraphic framework of Euro-
pean basins, in P. De Graciansky, J. J. Hardenbol, T.
Jacquin, and P. R. Vail, eds., Mesozoic and Cenozoic
REFERENCES CITED sequence stratigraphy of European basins: SEPM Spe-
cial Publication 60, p. 3 – 14.
Al-Menhali, S. S., W. L. S. Abu, and J. S. Schuelke, 2005, Hughes, G. W., 2000, Bioecostratigraphy of the Shuaiba
Rock property prediction using multiple seismic and Formation, Shaybah field, Saudi Arabia: GeoArabia,
geologic attributes provides insight to field develop- v. 5, p. 545 – 578.
ment for a large U.A.E. field: International Petroleum Hunt, D., and M. E. Tucker, 1992, Stranded parasequences
Technology Conference (Qatar) Paper 10595, CD-ROM. and the forced regressive wedge systems tract: Depo-
Alsharhan, A. S., 1985, Depositional environment, res- sition during base-level fall: Sedimentary Geology,
ervoir units evolution, and hydrocarbon habitat of v. 81, p. 1 – 9.
Shuaiba Formation, Lower Cretaceous, Abu Dhabi, Immenhauser, A., A. Cresusen, M. Esteban, and H. B.
United Arab Emirates: AAPG Bulletin, v. 69, p. 899 – Vonhof, 2000, Recognition and interpretation of poly-
912. genic discontinuity surfaces in the middle Cretaceous
Alsharhan, A. S., 1993, Bu Hasa field — United Arab Shuaiba, Nahr Umr, and Natih formations of northern
Emirates: Rub’ al Khali basin, Abu Dhabi, in N. H. Oman: GeoArabia, v. 5, no. 2, p. 299 – 322.
Foster and E. A. Beaumont, compilers, Structural traps Johnson, C. A., T. Hauge, S. Al-Mehhali, S. B Sumaidaa,
VIII: AAPG Treatise of Petroleum Geology, Atlas of Oil B. Sabin, and B. West, 2005, Structural styles and tec-
and Gas Fields, v. A-26, p. 99 – 127. tonic evolution of onshore and offshore Abu Dhabi,
Alsharhan, A. S., and A. E. M., Nairn, 1993, Carbonate U.A.E.: International Petroleum Technology Confer-
platform models of Arabian Cretaceous reservoirs, in ence (Qatar) Paper 10646, CD-ROM.
J. A. Simo, R. W. Scott, and J. Masse, eds., Cretaceous Marzouk, I. M., and M. A. Sattar, 1993, Implications of
carbonate platforms: AAPG Memoir 56, p. 173 – 184. wrench tectonics on hydrocarbon reservoirs, Abu
Azer, S. R., and C. Toland, 1993, Sea level changes in the Dhabi, U.A.E.: Proceedings of 8th Middle East Oil Show
Aptian and Barremian (upper Thamama) of offshore and Conference, Bahrain, Society of Petroleum Engi-
Abu Dhabi, U.A.E.: Society of Petroleum Engineers neers Paper 25608, p. 119 – 130.
Middle East Oil Technical Conference and Exhibition, Masafarro, J. L., R. Bourne, and J. C. Jauffred, 2004, Three-
Bahrain, SPE Paper 25610, p. 141 – 154. dimensional seismic volume visualization of carbon-
Bralower, T. J., M. A. Arthur, R. M. Leckie, W. V. Sliter, D. J. ate reservoirs and structures, in G. P. Eberli, J. L.
Allard, and S. O. Schlanger, 1994, Timing and paleo- Masaferro, and J. F. Sarg, eds., Seismic imaging of
ceanography of oceanic dysoxia/anoxia in the late carbonate reservoirs and systems: AAPG Memoir 81,
Barremian to early Aptian: Palaios, v. 9, p. 335 – 369. p. 11 – 41.
Droste, H. J., 2004, Regional controls on reservoir Plint, A. G., and D. Nummedal, 2000, The falling stage
properties in the Shuaiba Formation of North Oman systems tract: Recognition and importance in se-
(abs.): 6th Middle East Geoscience Conference and quence stratigraphic analysis, in D. Hunt and R. L.
Exhibition, Bahrain, CD-ROM. Gawthorpe, eds., Sedimentary responses to forced
Eberli, G. P., G. T. Baechle, F. S. Anselmetti, M. L. Incze, regressions: Geological Society (London) Special Pub-
2003, Factors controlling elastic properties in carbon- lication 172, p. 1 – 18.
ate sediments and rocks: The Leading Edge, v. 22, Premoli Silva, I., E. Erba, G. Salvini, C. Locatelli, and D. Verga,
no. 7, p. 654 – 660. 1999, Biotic changes in Cretaceous oceanic anoxic
Erba, E., J. E. T. Channell, M. Claps, C. Jones, R. Larson, events of the Tethys: Journal of Foraminiferal Research,
B. Opdyke, I. Premoli-Silva, A. Riva, G. Salvini and v. 29, p. 352– 370.
S. Torricelli, 1999, Integrated stratigraphy of the Cismon Russell, S. D., M. Akbar, B. Vissapragada, and G. M. Walkden,
APTICORE (southern Alps, Italy): A ‘‘reference section’’ 2002, Rock types and permeability prediction from
for the Barremian – Aptian interval at low latitudes: dipmeter and image logs: Shuaiba reservoir (Aptian),
Journal of Foraminiferal Research, v. 29, p. 371 – 391. Abu Dhabi: AAPG Bulletin, v. 86, p. 1709 – 1732.
Greselle, B., and B. Pittet, 2005, Fringing carbonate plat- Sarg, J. F., and J. S. Schuelke, 2003, Integrated seismic
forms at the Arabian plate margin in northern Oman analysis of carbonate reservoirs: From the framework
during the late Aptian – middle Albian: Evidence for to the volume attributes: The Leading Edge, v. 22,
high-amplitude sea-level changes: Sedimentary Geol- no. 7, p. 640 – 645.
ogy, v. 175, p. 367 – 390. Schuelke, J. S., L. A. Yose, S. Al-Menhali, and W. Soroka,
212 / Yose et al.

2005, Seismic rock property predictions provide insight voirs of the world: From rocks to reservoir characteri-
to field development (abs.): AAPG Annual Meeting zation and modeling: AAPG Memoir 88/SEPM Special
Program, v. 14, p. A125. Publication, p. 139– 171.
Sharland, P. R., R. Archer, D. M. Casey, R. B. Davies, S. H. van Buchem, F. S. P., B. Pittet, H. Hillgartner, J. Grotsch,
Hall, A. P. Heward, A. D Horbury, and M. D. Simmons, A. I. Al Mansouri, I. M. Billig, H. H. J. Droste, W. H.
2001, Arabian plate sequence stratigraphy: GeoArabia Oterdoom, and M. van Steenwinkel, 2002, High-
Special Publication 2, 371 p. resolution sequence stratigraphic architecture of
Sharland, P. R., D. M. Casey, R. G. Davies, M. D. Simmons, Barremian/Aptian carbonate systems in northern
and O. E. Sutcliffe, 2004, Arabian plate sequence Oman and the United Arab Emirates (Kharaib and
stratigraphy — Revisions to SP2: GeoArabia, v. 9, no. 1, Shuaiba formations): GeoArabia, v. 7, p. 461 – 500.
p. 199 – 214. Yose, L. A., and J. F. Collins, 2002, Windward-leeward
Strohmenger, C. J., L. J. Weber, A. Ghani, K. Al-Mehsin, models for carbonate platforms revisited (abs.); AAPG
O. Al-Jeelani, A. Al-Mansoori, T. Al-Dayyani, L. Vaughan, Annual Meeting Program, v. 11, p. A196.
S. A. Khan, and J. C. Mitchell, 2006, High-resolution Yose, L. A., et al., 2004, New frontiers in 3-D seismic char-
sequence stratigraphy and reservoir characterization of acterization of carbonate reservoirs: Example from a
upper Thamama (Lower Cretaceous) reservoirs of a supergiant field in Abu Dhabi: 11th Abu Dhabi In-
giant Abu Dhabi oil field, United Arab Emirates, in P. M. ternational Petroleum Exhibition and Conference,
Harris and L. J. Weber, eds., Giant hydrocarbon reser- Society of Petroleum Engineers, SPE Paper 88689, 16 p.
Strohmenger, C. J., P. E. Patterson, G. Al-Sahlan, J. C. Mitchell, H. R. Feldman,

6
T. M. Demko, R. W. Wellner, P. J. Lehmann, G. G. McCrimmon, R. W.
Broomhall, and N. Al-Ajmi, 2006, Sequence stratigraphy and reservoir
architecture of the Burgan and Mauddud formations (Lower Cretaceous),
Kuwait, in P. M. Harris and L. J. Weber, eds., Giant hydrocarbon reservoirs
of the world: From rocks to reservoir characterization and modeling:
AAPG Memoir 88/SEPM Special Publication, p. 213 – 245.

Sequence Stratigraphy and


Reservoir Architecture of the
Burgan and Mauddud Formations
(Lower Cretaceous), Kuwait
Christian J. Strohmenger,1 Timothy M. Demko
John C. Mitchell, Howard R. Feldman, University of Minnesota Duluth, Duluth,
Patrick J. Lehmann, and Minnesota, U.S.A.
Robert W. Broomhall
ExxonMobil Exploration Company, Houston, Robert W. Wellner
Texas, U.S.A. ExxonMobil Upstream Research Company,
Houston, Texas, U.S.A.
Penny E. Patterson
ExxonMobil Upstream Research Company, G. Glen McCrimmon
Houston, Texas, U.S.A. Hibernia Management and Development
Company, St. John’s, Newfoundland
Ghaida Al-Sahlan and Labrador, Canada
Kuwait Oil Company, Ahmadi, Kuwait
Neama Al-Ajmi
Kuwait Oil Company, Ahmadi, Kuwait

ABSTRACT

A
new sequence-stratigraphic framework is proposed for the Burgan and
Mauddud formations (Albian) of Kuwait. This framework is based on the
integration of core, well-log, and biostratigraphic data, as well as seismic
interpretation from giant oil fields of Kuwait.
The Lower Cretaceous Burgan and Mauddud formations form two third-
order composite sequences, the older of which constitutes the lowstand, trans-
gressive, and highstand sequence sets of the Burgan Formation. This composite
sequence is subdivided into 14 high-frequency, depositional sequences that are
characterized by tidal-influenced, marginal-marine deposits in northeast Kuwait
that grade into fluvial-dominated, continental deposits to the southwest.

1
Present address: Abu Dhabi Company for Onshore Oil Operations, Abu Dhabi, United Arab Emirates.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215878M883271

213
214 / Strohmenger et al.

The younger composite sequence consists of the lowstand sequence set of


the uppermost Burgan Formation and transgressive and highstand sequence sets
of the overlying Mauddud Formation. This composite sequence is sand prone
and mud prone in southern and southwestern Kuwait and is carbonate prone
in northern and northeastern Kuwait. The lowstand sequence set deposits of
the Burgan Formation are subdivided into five high-frequency depositional se-
quences, which are composed of tidal-influenced, marginal-marine deposits in
northeastern Kuwait that change facies to fluvial-dominated deposits in south-
western Kuwait. The transgressive and highstand sequence sets of the Mauddud
Formation are subdivided into eight high-frequency, depositional sequences. The
Mauddud transgressive sequence set displays a lateral change in lithology from
limestone in northern Kuwait to siliciclastic deposits in southern and south-
western Kuwait. The traditional lithostratigraphic Burgan –Mauddud contact is
time transgressive. The Mauddud highstand sequence set is carbonate prone and
thins south- and southwestward because of depositional thinning. Significant
postdepositional erosion occurs at the contact with the overlying Cenomanian
Wara Shale.
The proposed sequence-stratigraphic framework and the incorporation of a
depositional facies scheme tied to the sequence-stratigraphic architecture allow
for an improved prediction of reservoir and seal distribution, as well as reservoir
quality away from well control.

INTRODUCTION Al-Eidan et al., 2001) (Figures 1, 2). The recoverable


oil reserves are estimated to be in the billions of bar-
A regional sequence-stratigraphic analysis of the rels (Christian, 1997).
Lower Cretaceous Burgan and Mauddud formations Sabiriyah field was discovered in 1956. It is an
was undertaken through a joint study conducted by elongated faulted anticline with production of 28 –
ExxonMobil Exploration Company and Kuwait Oil 328 API oil from the Burgan and Mauddud forma-
Company. The study focused on the stratigraphic ar- tions (Milton and Davies, 1965; Adasani, 1967; Al-
chitecture of selected major oil fields throughout Ku- Rawi, 1981; Brennan, 1990a; Carman, 1996; Kaufman
wait, including the supergiant Greater Burgan field et al., 1997; Al-Eidan et al., 2001) (Figures 1, 2). The
in southeastern Kuwait, and Raudhatain and Sabiri- recoverable oil reserves are estimated to be in the bil-
yah fields in northern Kuwait (Figure 1). lions of barrels (Christian, 1997).
Three culminations constitute the supergiant Great- The Burgan and Mauddud formations are part of
er Burgan field: Burgan, Magwa, and Ahmadi. These the Wasia Group that overlies the Lower Cretaceous
three culminations are located near the crest of the Thamama Group of the Arabian plate (Alsharhan and
north–south-trending Kuwait arch (Fox, 1961; Ada- Nairn, 1993, 1997). The lower to middle Albian Bur-
sani, 1965; Brennan, 1990b; Carman, 1996) (Figure 1). gan Formation is the major oil-bearing sandstone res-
The first well on these structures was drilled in 1938, ervoir throughout the Greater Burgan field, as well as
followed by wells in 1951 and 1952. Production of at Raudhatain and Sabiriyah fields in northern Kuwait.
28 – 368 API oil comes from the Burgan (the major The thickness ranges from approximately 1250 ft
oil-producing reservoir), the Mauddud (a minor oil- (380 m) at the Greater Burgan field area to approxi-
producing reservoir), and the Wara formations (Kauf- mately 900 ft (275 m) at Raudhatain and Sabriya fields
man et al., 1997). The recoverable oil reserves are esti- area (Bou-Rabee, 1996). The overlying upper Albian
mated to be in the tens of billions of barrels (Christian, Mauddud Formation is a major oil-bearing carbonate
1997). reservoir in northern Kuwait (Al-Anzi, 1995). Thick-
Raudhatain field was discovered in 1955. It is a ness of the Mauddud Formation ranges from only a
faulted anticlinal dome with production of 28 – 408 few feet at the Greater Burgan field and Minagish field
API oil from the Ratawi, Zubair, Burgan, and Maud- areas (south and southwest Kuwait) to approximately
dud formations (Milton and Davies, 1965; Adasani, 450 ft (140 m) at the Abdali, Raudhatain, and Sabi-
1967; Al-Rawi, 1981; Brennan, 1990a; Carman, 1996; riyah fields in northern Kuwait (Bou-Rabee, 1996).
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 215

FIGURE 1. Location map showing the major oil fields of Kuwait (green) and the wells (black dots) studied.

Methodology graphic interpretations were integrated into this


The emphasis of this study involved the interpreta- study.
tion of the sequence-stratigraphic architecture based on The identified chronostratigraphic surfaces (SB, TS,
correlation of regional stratal surfaces (sequence bound- MFS, and FS) were assigned to high-frequency se-
aries [SB], transgressive surfaces [TS], maximum flooding quences (HFS) and numbered sequentially from the
surfaces [MFS], and flooding surfaces [FS]) through- top down; the underlying sequence boundary giv-
out Kuwait, as well as on the sequence-stratigraphy- ing name to the overlying high-frequency sequence
keyed facies analyses of the Burgan and Mauddud (Figure 3). The HFSs were subsequently grouped into the
formations. Approximately 9600 ft (2930 m) of con- sequence sets of two third-order composite sequences
ventional core from 30 wells penetrating the Burgan (Mitchum, 1977; Mitchum et al., 1977; Vail et al., 1977,
Formation and approximately 5000 ft (1520 m) of con- 1991; Haq et al., 1987, 1988; Vail, 1987; Van Wagoner
ventional core from 35 wells penetrating the Mauddud et al., 1987, 1988; Sarg, 1988; Haq, 1991; Mitchum
Formation were described sedimentologically. The de- and Van Wagoner, 1991; Sarg et al., 1999) (Figure 3).
positional environments interpreted from core were
correlated to well-log signatures and used to develop High-frequency Sequences
the regional sequence-keyed sequence-stratigraphic High-frequency sequences in the siliciclastic Bur-
framework. More than 100 wells were correlated with- gan Formation (Figure 3) are defined by the stacking
in this sequence-stratigraphic context. In addition, patterns and facies distributions. These sequences
the results from seismic-stratigraphic and biostrati- may contain lowstand, transgressive, and highstand
216 / Strohmenger et al.

FIGURE 2. Seismic cross section oriented northwest – southeast showing Raudhatain and Sabiriyah structures and
the interpreted main stratigraphic horizons. Seismic line runs through the center of Raudhatain and Sabiriyah fields
shown in Figure 1.

systems tracts (LST, TST, and HST). Each systems tract nitic sandstones in the northern Kuwait Raudhatain
exhibits distinct facies trends, thickness distribu- and Sabiriyah fields area and grade into siliciclastics
tions, and reservoir quality (Strohmenger et al., 2002; toward the south at Greater Burgan field and toward
Demko et al., 2003). the southwest at Minagish field areas. Highstand sys-
The LSTs consist of incised-valley fills (IVF). These tems tracts typically show an upward increase in
valleys become thinner, as well as more laterally grain richness (graining upward) as well as porosity
discontinuous and tidal influenced downdip to the (Strohmenger et al., 2002; Demko et al., 2003).
northeast. Transgressive systems tracts correspond to
retrogradational successions of coarsening-upward, Composite Sequences
marginal-marine mudstones and sandstones that grade High-frequency sequences are grouped into se-
into marine carbonates downdip and mudstone-prone quence sets of two third-order composite sequences,
coastal-alluvial plain deposits updip. Highstand sys- based on the stacking patterns and facies distribu-
tems tracts correspond to progadational successions tions (Figure 3). Each composite sequence consists
of coarsening-upward, marginal-marine mudstones of a lowstand sequence set (LSS), transgressive se-
and sandstones that grade updip into mudstone- quence set (TSS), and a highstand sequence set (HSS)
prone coastal-alluvial plain deposits (Strohmenger (Strohmenger et al., 2002; Demko et al., 2003).
et al., 2002; Demko et al., 2003). In the siliciclastic Burgan Formation of Kuwait,
High-frequency sequences in the carbonate- LSSs are characterized by an aggradational stacking
dominated Mauddud Formation (Figure 3) are de- of HFSs dominated by braided fluvial deposits. The
fined by parasequence (PS) stacking patterns, facies TSS exhibits an overall retrogradational stacking
distributions, and microkarst or exposure surfaces. pattern, dominated by nonmarine facies at the base,
High-frequency sequences of the Mauddud Forma- whereas the uppermost HFSs contain increasing
tion may contain TST and HST (Strohmenger et al., marginal-marine components. The HSS forms an overall
2002; Demko et al., 2003). progradational succession dominated by marginal-
Transgressive systems tracts are generally more mud marine facies, especially in northern Kuwait (Stroh-
dominated with intercalated sandstones and glauco- menger et al., 2002; Demko et al., 2003).
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 217

FIGURE 3. Mauddud – Burgan sequence-stratigraphic framework showing the lower and upper third-order composite
sequences, as well as the interpreted high-frequency depositional sequences (Raudhatain field type well RA-G). GR =
gamma-ray log; MD = measured depth (feet); RES = resistivity log; NEU = neutron porosity log; DENS = density log.

The Mauddud TSS shows a lateral change in lithol- menger et al., 2002; Demko et al., 2003). To define
ogy from limestone in northern Kuwait to siliciclas- coeval facies successions in both the Burgan and
tics in southern and southwestern Kuwait. An overall Mauddud formations, a chronostratigraphically sig-
shoaling-upward or progradational signature char- nificant regional flooding surface (B100_TS) was de-
acterizes the HSS. Most of the upper HSS is removed fined as the Burgan-Mauddud contact (Figure 3).
by erosion throughout much of southern and south- Within this chronostratigaphic framework, the up-
western Kuwait (Strohmenger et al., 2002; Demko permost Burgan and overlying Mauddud formations
et al., 2003). form a composite sequence (Figure 3) that becomes
more siliciclastic prone to the southwest and carbon-
ate prone to the northeast. A second composite se-
BURGAN FORMATION quence encompasses the rest of the underlying Burgan
Formation (Figure 3). This composite sequence is dom-
A regional sequence-stratigraphic analysis of the inated by marginal-marine deposits to the northeast
Burgan and Mauddud formations reveals that the and nonmarine deposits to the southwest.
traditional lithostratigraphic Burgan-Mauddud con- The Burgan Formation, as defined in this study,
tact is a time-transgressive facies boundary (Stroh- comprises 19 HFSs (Figure 3). Each of these HFSs
218 / Strohmenger et al.

FIGURE 4. Sequence-keyed depositional facies models for the Burgan Formation. In these models, the lowstand systems
tract consists of incised-valley deposits, whereas the trangressive and highstand systems tracts are composed of wave-
dominated shoreface depositional systems. GR = gamma-ray log.

contains LSTs, TSTs, and HSTs, each with distinct facies The primary reservoirs in the Burgan Formation
trends, thickness distributions, and reservoir quality. are fluvial and tidal deposits that formed within the
The LSTs of these sequences consist of IVF. These incised valleys (Figure 4). Shoreline sandstones in the
valleys become thinner, as well as more laterally dis- TSTs and HSTs are also potential reservoirs (Figure 4).
continuous and tidal influenced downdip to the However, these marginal-marine sandstones have low-
northeast. The TSTs of the Burgan sequences display a er porosity and permeabilities because of their finer
systematic downdip to updip change from marine grained and increased clay matrix due to bioturba-
carbonates to marginal-marine mudstones and sand- tion. In summary, the sequence-stratigraphic analy-
stones to mudstone-prone alluvial- and coastal-plain sis of the Burgan Formation provides an improved
deposits. The HSTs are dominated by marginal-marine understanding of the spatial and temporal distribution
sandstones and mudstones downdip and alluvial- of reservoirs that can be used to address exploration-
and coastal-plain mudstones and sandstones updip. scale to production-scale issues.
The shorelines in these highstands trend northwest – In general, the Burgan Formation is a classic re-
southeast and are best developed in the northern gressive-transgressive-regressive package. It is domi-
Raudhatain and Sabiriyah fields area. nated by sandstone-prone fluvial deposits at its base
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 219

and top and marginal-marine mudstones, sandstones, cross-bedding and current-ripple cross-lamination.
and limestones in the middle parts to the northeast. Large plant fossils, including compressed branches
The Burgan Formation is bounded at its top by a re- and leaves, are common. Mudstone drapes and bio-
gional flooding surface, referred to as the Burgan trans- turbation are uncommon. Grain size trends are typi-
gressive surface B100_TS (Figure 3). In general, this cally fining upward. Small-scale fining-upward beds
boundary is marked by a change from blocky, high are 0.2–2 ft (0.06–0.6 m) thick and typically asso-
net/gross, fluvial-dominated sandstones of the Bur- ciated with individual trough cross-bed sets. Larger-
gan Formation below to low net/gross marine mud- scale, fining-upward trends range from 20 to 50 ft
stones, sandstones, and limestones of the Mauddud (6 to 15 m) of thickness.
Formation above. Within this context, fine-grained The cross-bedding and grain-size trends all indicate
siliciclastics, traditionally included at the top of the deposition in a low-sinuosity fluvial setting. Some
Burgan Formation to the south, are assigned to the fining-upward trends suggest sedimentation in sandy
clastic member of the Mauddud Formation. This Maud- point bars. This lithofacies occurs primarily in incised
dud clastic member occurs above the Burgan transgres- valleys and in updip braid-plain deposits.
sive surface B100_TS, a regional flooding surface, de-
fined as the top of the Burgan Formation and beneath
the carbonate strata traditionally included within the Lithofacies 1B: Trough Cross-bedded Sandstone
Mauddud Formation (Mauddud carbonate member). with Minor Clay Drapes
Lithofacies 1B (Figure 5B) consists of poorly to mod-
erately sorted, fine-grained to coarse-grained sand-
Lithofacies and Depositional Environments stones that possess sparse to common clay drapes and
Within the Burgan Formation, 22 lithofacies were thick-thin couplets. Sedimentary structures are domi-
identified and define 6 distinct facies. The physical nated by trough- and current-ripple cross-bedding.
criteria used to delineate the individual lithofacies in Small mudstone and siderite clasts may be locally
this study include grain size, composition, sorting, abundant. Small-scale, grain-size trends are typically
grading, physical and biogenic sedimentary struc- fining upward. Large-scale, grain-size trends are ei-
tures, stratal boundaries, presence or absence of clay ther coarsening upward or fining upward. Large plant
drapes, organic-rich drapes, organic debris, and dia- fossils, such as compressed sticks and leaves, may be
genetic features. For ease of description, each litho- common. Bioturbation is sparse to moderate. Common
facies was classified into six categories based on mud burrow types are horizontal sand-filled tubes with a
content and apparent reservoir quality. The six cate- circular cross section, approximately 1 cm (0.4 in.) in
gories range from well-sorted, very clean sand (litho- diameter (Planolites).
facies 1: excellent reservoir quality) to mudstone (litho- The clay drapes and thick-thin couplets indicate
facies 5: very poor reservoir quality) and, ultimately, that some tidal influence occurred during deposition.
to coal (lithofacies 6). In addition, lithofacies were This lithofacies occurs at the fluvial to tidal transi-
assigned an alphabetic qualifier to account for subtle tion in IVF and at the top of coarsening-upward tidal
differences in grain size, mud content, dominant sedi- bars.
mentary structures, and type and extent of bioturba-
tion. Trace fossil assemblages were used to infer en-
vironments of deposition, which are based on models Lithofacies 1C: Current-rippled,
proposed by Pemberton et al. (1992a, b). In addition, Cross-laminated Sandstone
our regional database includes two lithofacies that This lithofacies (Figure 5C) consists of well-sorted,
were not observed in the Burgan strata. They are litho- very fine-grained to fine-grained sandstone. Sedimen-
facies 3A, coarsening-upward, bioturbated mudstone tary structures are characterized by current-ripple
to sandstone; and lithofacies 4C, bioturbated to pla- cross-lamination with minor small-scale trough cross-
nar laminated mudstone to sandstone. bedding. Clay drapes are sparse to absent. This facies
occurs in beds typically a few feet thick with no grain-
size trends. Bioturbation is uncommon. Small plant
Lithofacies 1A: Trough Cross-bedded Sandstone fossils, such as small leaves and fine plant fragments,
Lithofacies 1A (Figure 5A) consists of poorly to are common.
moderately sorted, fine-grained to coarse-grained sand- Lithofacies 1C occurs in a range of environments.
stone. The dominant sedimentary structures are trough Where it is associated with coastal-plain deposits, it
220 / Strohmenger et al.

FIGURE 5. Siliciclastic lithofacies 1, slabbed core photographs. (A) Lithofacies 1A: trough cross-bedded sandstone.
(B) Lithofacies 1B: trough cross-bedded sandstone with minor clay drapes. (C) Lithofacies 1C: current-rippled, cross-
laminated sandstone, which is interbedded with small-scale, trough cross-beds.

represents splays and channel levees. Within distal of a low diversity of burrows, including Planolites,
incised valleys, this lithofacies represents sedimen- Skolithos, and Arenicolites. Fine plant debris and amber
tation in tidal channels. flakes are common.
This lithofacies represents deposition in tidal creeks
or tidal flats.
Lithofacies 2A: Clay-draped, Current-rippled to
Laminated Sandstone
Lithofacies 2A (Figure 6A) consists of well-sorted, Lithofacies 2B: Clay-draped, Current-rippled to
very fine- to fine-grained sandstone. Sedimentary Trough Cross-bedded Sandstone
structures include current-ripple cross-laminations Lithofacies 2B (Figure 6B) consists of very fine-
and minor horizontal laminations, with abundant to fine-grained sandstone. The dominant sedimen-
thin (<1-mm; <0.04-in.) clay and organic drapes, com- tary structures are current-ripple cross-laminations
monly on ripple foresets. Some laminations exhibit to small-scale, trough– cross-bedding with abundant
centimeter-scale cyclicity in mud-sand couplets. Sider- thin (1-mm) clay and organic drapes, commonly
ite cement is common. Bedsets typically fine upward. on ripple and trough foresets. Bedsets may distinctly
Bioturbation is uncommon to moderate and consists coarsen upward. Bioturbation is sparse to moderate
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 221

FIGURE 6. Siliciclastic lithofacies 2, slabbed core photographs. (A) Lithofacies 2A: clay-draped, current-rippled to
laminated sandstone. (B) Lithofacies 2B: clay-draped, current-rippled to trough cross-bedded sandstone. (C) Lithofacies
2C: clay-draped, horizontally laminated sandstone. (D) Lithofacies 2D: glauconitic sandstone that has been extensively
bioturbated by Teichichnus (Te) and Planolites (Pl) burrows.

and is characterized by the low diversity of burrows than those of the detrital quartz). Bioturbation is mod-
(mostly Planolites). Fine plant debris and amber flakes erate to extensive and is characterized by Thalassi-
are locally abundant. noides, Asterosoma, Planolites, Teichichnus, Palaeophycus,
This lithofacies is inferred to have been deposited and Scolicia. Siderite concretions are common, par-
as subtidal bars in estuaries. ticularly associated with Thalassinoides burrows. This
facies is almost always cemented by calcite and grades
Lithofacies 2C: Clay-draped, Horizontally into overlying limestone beds. Mollusk shells and
Laminated Sandstone shell fragments are common.
Lithofacies 2C (Figure 6C) is dominated by hori- Lithofacies 2D represents deposition in the distal
zontal laminations, minor current ripples with abun- parts of low-energy, marine shorelines.
dant thin (1-mm) clay and organic drapes that de-
fine horizontal laminations. Most laminations show Lithofacies 3B: Bioturbated, Muddy Sandstone
some centimeter-scale cyclicity of mud-sand couplets. This lithofacies (Figure 7A) consists of poorly to
Lamina and bedsets typically fine upward. Biotur- moderately well-sorted, upper fine-grained to lower
bation is sparse to moderate and is characterized by medium-grained muddy sandstone. Bioturbation is
Planolites, Palaeophycus, Arenicolites, Skolithos, and typically extensive and is characterized by multiple
Cylindrichnus. Fine plant debris and amber flakes and tiers of Thalassinoides, Teichichnus, Asterosoma, Pla-
pebbles may be locally abundant. nolites, Palaeophycus, Skolithos, and Cylindrichnus.
This lithofacies is interpreted to have been depos- This lithofacies represents deposition along brack-
ited on tidal flats and in associated low-energy, tidal ish, estuarine, and marine low-energy shorelines.
point bars.
Lithofacies 3C: Hummocky
Lithofacies 2D: Glauconitic Sandstone Cross-bedded Sandstone
This lithofacies (Figure 6D) consists of fine-grained Lithofacies 3C (Figure 7B), which is not often ob-
to medium-grained sandstone with abundant grains served in the Burgan Formation, consists of mod-
of glauconite (which are invariably coarser grained erately well- to very well-sorted, very fine-grained to
222 / Strohmenger et al.

FIGURE 7. Siliciclastic lithofacies 3, slabbed core photographs. (A) Lithofacies 3B: bioturbated, muddy sandstone
containing shell fragments. (B) Lithofacies 3C: hummocky cross-bedded sandstone. (C) Lithofacies 3D: interlaminated
sandstone, silt, and shale. (D) Lithofacies 3E: current-rippled, interbedded sandstone and mudstone.

lower fine-grained sandstone. The dominant sedi- lithos, and Cylindrichnus. Plant debris, amber, and
mentary structures are hummocky, cross-stratified siderite-cemented bands are common.
beds with wave-rippled and bioturbated upper sur- This lithofacies represents deposition on proximal
faces. Mudstone beds as much as several centimeters (sandy) to distal (muddy) tidal flats, and the centimeter-
in thickness are only rarely preserved between hum- scale cycles are interpreted as neap-spring tidal cycles.
mocky bedsets. Bioturbation is typically absent to
sparse, but may increase upward with moderate bur- Lithofacies 3E: Current-rippled, Interbedded
rowing at the top of some bed sets. Trace fossils rec- Sandstone and Mudstone
ognized include Planolites, Palaeophycus, and uncom- Lithofacies 3E (Figure 7D) consists of interbedded,
mon Ophiomorpha. lower medium-grained sandstone and mudstone.
This lithofacies was most likely deposited in ma- Sedimentary structures are dominated by horizontal
rine, proximal lower shoreface settings. lamination and current-ripple cross-lamination, with
uncommon syneresis cracks. Bioturbation is sparse
Lithofacies 3D: Interlaminated Sandstone, to moderate and is characterized mostly by Planolites,
Silt, and Shale Palaeophycus, Skolithos, and Cylindrichnus. Plant debris
Lithofacies 3D (Figure 7C) consists of interlami- and amber are locally abundant as well as siderite-
nated, very fine-grained sandstone, silt, and shale. cemented bands.
Sedimentary structures include horizontal bedding, This lithofacies is interpreted to have been depos-
starved current-ripple cross-lamination, and uncom- ited in tidal flats and distal tidal bars.
mon syneresis cracks. Depositional strata range from
mostly clay to mostly sand, but are characterized by Lithofacies 3F: Calcareous, Bioturbated Sandstone
sand and mud couplets, thick and thin couplets, and This lithofacies (Figure 8A) consists of fine-grained
centimeter-scale cycles of thickening and thinning to lower medium-grained sandstone. Bioturbation
sand laminae. Bioturbation is sparse to moderate and is typically extensive (churned). Mollusk shells, shell
include trace fossils of Planolites, Palaeophycus, Sko- fragments, and foraminifera are locally abundant.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 223

FIGURE 8. Siliciclastic lithofacies 3, slabbed core photographs. (A) Lithofacies 3F: calcareous, bioturbated sandstone
containing shell fragments. (B) Lithofacies 3G: wave-rippled sandstone. (C) Lithofacies 3H: carbonaceous, muddy
sandstone.

This facies is tightly cemented by calcite and grades common to slight and comprises Planolites burrows
into limestone beds above or below. and rootworking. Large plant fossils and amber are
This lithofacies represents deposition in shallow- very abundant.
marine shoreline or shoreface settings adjacent to This lithofacies represents deposition in flood-
subtidal carbonate facies. plain, crevasse splay, or levee settings.

Lithofacies 3G: Wave-rippled Sandstone Lithofacies 4A: Heterolithic Siltstone to Mudstone


Lithofacies 3G (Figure 8B) consists of well-sorted, This lithofacies (Figure 9A) consists mostly of mud-
very fine-grained to fine-grained sandstone. Sedimen- stone with siltstone to very fine-grained sandstone
tary structures consist of wave-ripple cross-laminations. stringers (>80% mudstone). Sedimentary structures
Clay drapes are common between ripple laminae and include parallel-lamination, silt stringers, uncom-
laminae sets. Bioturbation is sparse to moderate and mon syneresis cracks, and starved current-ripple cross-
includes Astersoma, Teichichnus, and Planolites burrows. lamination. Bioturbation is uncommon to slight and
This lithofacies is interpreted to have formed in is characterized by Planolites burrows. Plant fossils
distal, lower shoreface settings. and amber are common to abundant.
Lithofacies 4A is interpreted to represent depo-
sition in flood-plain and abandoned channel-fill
Lithofacies 3H: Carbonaceous, Muddy Sandstone settings.
Lithofacies 3H (Figure 8C) consists of very fine- to
fine-grained sandstone with abundant clay and or-
ganic drapes ranging from 5 to 10 mm (0.2 to 0.4 in.) Lithofacies 4B: Bioturbated to Wave-rippled
in thickness. Organic layers range from 1 mm (0.04 in.) Sandstone to Mudstone
to a few centimeters in thickness. Sedimentary struc- This lithofacies (Figure 9B) is composed of lower
tures are dominated by horizontal lamination and very fine-grained sandstone to siltstone to clay. Sedi-
current-ripple cross-lamination. Bioturbation is un- mentary structures include isolated to amalgamated
224 / Strohmenger et al.

FIGURE 9. Siliciclastic lithofacies 4, slabbed core photographs. (A) Lithofacies 4A: heterolithic siltstone to mudstone.
(B) Lithofacies 4B: bioturbated to wave-rippled sandstone to mudstone. This core interval has been extensively bio-
turbated by Teichichnus (Te), Asterosoma (As), and Planolites (Pl) burrows. (C) Lithofacies 4D: bioturbated mudstone. The
upper interval of this core has been extensively bioturbated by Teichichnus (Te) burrows. (D) Lithofacies 4E: laminated
siltstone.

wave-ripple cross-laminations. Bioturbation is slight Lithofacies 5A: Laminated Gray Shale


to extensive (churned) and is characterized by Aste- Lithofacies 5A (Figure 10A) is composed of dark-
rosoma, Teichichnus, Thalassinoides, Planolites, Palaeo- gray, horizontally laminated shale. Bioturbation is
phycus, Scolicia, and Skolithos. Fine plant fragments slight to moderate and is characterized by Planolites,
and mollusk shells are uncommon. Teichichnus, sand-filled Chondrites, and sparse Zoo-
This lithofacies is interpreted as deposition in ma- phycos burrows. Bivalve and gastropod shells range
rine, distal lower shoreface settings. from sparse to common, and small plant fragments
are uncommon.
Lithofacies 4D: Bioturbated Mudstone This lithofacies is interpreted to have been deposited
Lithofacies 4D (Figure 9C) is mostly mudstone, in offshore marine settings, below storm-wave base.
but may grade vertically into siltstone. Bioturbation
is moderate to extensive (churned) and is charac- Lithofacies 5B: Carbonaceous Mudstone
terized by mostly indistinct burrows with some dis- Lithofacies 5B (Figure 10B) is composed of car-
cernible Planolites, Teichichnus, and Thalassinoides bonaceous, horizontally laminated mudstone, which
burrows. commonly displays postdepositional, compactional
This lithofacies is inferred to have been deposited slickensides. Bioturbation is slight to moderate and
in offshore to lower shoreface settings. is characterized mostly by rootworking (Figure 11A).
Large leaves and amber (Figure 11B, C) are very
Lithofacies 4E: Laminated Siltstone abundant.
This lithofacies (Figure 9D) is composed of hori- This lithofacies represents deposition in clastic
zontally laminated to wavy laminated siltstone. Bio- swamp and abandoned-channel settings.
turbation is uncommon to slight and is character-
ized by small (<5-mm; <0.2-in.) indistinct horizontal Lithofacies 5C: Laminated Dark Gray Shale
burrows. Lithofacies 5C (Figure 10C) is composed of very
Lithofacies 4E represents deposition in lacustrine, dark-gray, horizontally laminated shale. Bioturbation
flood-plain, pond, and abandoned-channel settings. is uncommon to slight and is characterized by small
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 225

FIGURE 10. Siliciclastic lithofacies 5 and 6, slabbed core photographs. (A) Lithofacies 5A: laminated gray shale.
(B) Lithofacies 5B: carbonaceous mudstone. (C) Lithofacies 5C: laminated dark-gray shale. (D) Lithofacies 6: coal.

Planolites burrows. Large leaves, sticks,


and amber are very abundant.
This lithofacies is interpreted to
represent deposition in lacustrine,
abandoned channel, and flood-
plain settings.

Lithofacies 6: Coal
This lithofacies (Figure 10D) is com-
posed exclusively of coal.
Some coals are rooted and repre-
sent peat swamps. Other coals are
composed of allochthonous plant
detritus and formed in abandoned
channels to estuarine settings with
low clastic input.

Facies, Facies Associations,


Depositional Environments, and
Reservoir Quality
The 22 lithofacies are grouped into
6 distinct facies. These groupings are
either spatially reoccurring groups
FIGURE 11. Slabbed core photographs of (A) rooted horizon, (B) amber, and of different lithofacies or thick oc-
(C) fragment of amber. currences of the same lithofacies. In
226 / Strohmenger et al.

turn, the six facies are grouped into two unique facies valley system (sensu Van Wagoner et al., 1990). The
associations. The facies associated with each are mu- typical idealized Burgan updip to downdip facies
tually exclusive (Figure 4). succession and associated lithofacies are

Facies Association I  cross-stratified sandstone facies (lithofacies 1A


Facies association I (Figure 4) consists of four and 1C): fluvial-dominated valley fill
unique facies, which are interpreted as the normal,  heterolithic tidal facies (lithofacies 1B, 2A, 2B,
shoaling-upward facies succession of a low-energy, 2C, 3D, 3E, and 4A): tidal-dominated valley fill
wave-dominated, shoreline (sensu Walker and Plint,
1992). This facies association is the basic building The incised valleys, identified in this study, typi-
block of highstand and transgressive deposits in the cally trend southwest to northeast across Kuwait and
Burgan Formation. The typical idealized vertical are perpendicular to somewhat oblique to the shore-
facies succession in this facies association from top line trend of the underlying highstand and overlying
to base is transgressive deposits. Deposits of the incised valley
are characterized by fluvial-dominated facies in up-
 carbonaceous facies (lithofacies 1C, 3H, 5B, and dip regions to the southwest and tidal-dominated
6): coastal-plain and backshore facies in the distal regions to the northeast. The
 bioturbated to stratified sandstone facies (litho- fluvial facies are interpreted as low-sinuosity, braided-
facies 3B and 3C): proximal lower shoreface stream deposits based on the prevalence of fining-
 bioturbated to interstratified mudstones and sand- upward, fine-grained to coarse-grained, trough–cross-
stones (lithofacies 2D, 3F, 3G, 4B, and 4D): distal beds and bedsets. The fluvial strata change facies
lower shoreface downdip to tidal-influenced deposits that possess more
 bioturbated mudstone facies (lithofacies 5A): heterolithic lithofacies, indicative of fluctuation-energy
offshore conditions.
Reservoir quality is excellent in the braided fluvial
Lithofacies association I is interpreted to represent strata and in the updip parts of the mixed fluvial-tidal
the deposits of a low-energy shoreline environment. systems. However, reservoir quality rapidly diminishes
Wave-dominated shoreface sandstones are oriented in the seaward direction as the valley systems thin
along a northwest-southeast belt and primarily occur and become more mudstone prone.
in northeastern Kuwait. Shoreface deposits change
facies to marginal-marine sandstones and mudstones High-frequency Sequences
and offshore mudstones in a northeast transect. They Nineteen high-frequency depositional sequences
change facies updip to coastal-plain deposits, which were interpreted within the Burgan Formation, as
dominate the southwestern region of Kuwait. The defined in this study (Figure 3). From bottom to top,
shoreface deposits are interpreted to represent a low- these are B900, B850, B800, B750, B725, B700, B650,
energy environment based on the paucity of high- B600, B550, B500, B450, B400, B350, B300, B250,
energy stratification and the presence of strata that B200, B150, B125, and B100 (LST).
have been moderately to intensely bioturbated by In the nonmarine to marine intervals of the
trace fossil assemblages indicative of open-marine Burgan Formation, high-frequency sequence bound-
conditions. aries are interpreted as abrupt vertical changes in
In general, reservoir quality ranges from moderate stratal stacking pattern or abrupt basinward shift in
in deposits of the proximal lower shoreface to poor in environments of deposition. In the downdip position,
the distal lower shoreface and offshore strata. Al- high-frequency sequence boundaries are evident as
though sandstone reservoirs may exist within small abrupt changes from coarsening-upward marginal-
(<0.5-km [<0.31-mi]-wide) fluvial channels on the marine parasequences to blocky fluvial-tidal deposits.
coastal-plain, for the most part, the coastal-plain part Updip, these sequence boundaries can be traced into
of this succession is mud prone and nonprospective. an abrupt change from interpreted low net/gross
coastal-alluvial plain (transgressive or highstand) de-
Facies Association II posits (below) into blocky fluvial (lowstand) deposits
Facies association II (Figure 4) consists of two (above; Figure 12). Downdip and laterally, the se-
unique facies, which are interpreted as the normal quence boundaries and transgressive surface become
updip to downdip facies variation in an incised- coincident. However, in the most distal sequences in
FIGURE 12. Regional cross section showing facies distribution and sequence-stratigraphic framework of the Burgan Formation. Note that the fluvial-dominated
sandstones of the incised-valley fills, which constitute the lowstand systems tracts, thicken toward the south in the vicinity of the Greater Burgan field area.
The overlying Mauddud clastic member and Mauddud carbonate member also display thickness variations along the depositional transect. The detailed
sequence-stratigraphic framework for the Mauddud Formation is shown in Figures 22 and 23. GR = gamma-ray log; MD = measured depth (feet); RES = resistivity
log; DT = sonic log; NEU = neutron porosity log; DENS = density log.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 227
228 / Strohmenger et al.

dominated strata in the updip regions. (B) Incised valleys become thinner and more discontinuous downdip. IVF = incised-valley fills; HST = highstand systems tract.
FIGURE 13. Depositional environment and isopach maps of a lowstand systems tract in the Burgan Formation. These maps are interpreted from the sequence-

upper, third-order composite sequence. (A) The lowstand deposits fill incised valleys and are composed of tidal-influenced strata in the downdip positions and fluvial-
the Burgan Formation, an onlap-

stratigraphic architecture of the B100 high-frequency, depositional sequence, which is the uppermost lowstand systems tract in the lowstand sequence set of the
ping succession of carbonates ap-
pears to mark a depositional shelf
break and downdip limit of clas-
tic progradation in each sequence.
In more marine-dominated
parts of the Burgan Formation,
the TSs are interpreted at the ver-
tical change from prograding
to retrograding marine parase-
quences. In more medial settings,
these surfaces coincide with the
vertical facies change, from blocky,
fluvial-tidal deposits below to
retrograding marine parase-
quences above. In more proxi-
mal (updip) settings, the TSs are
placed at the vertical change from
blocky fluvial-tidal below to low
net/gross coastal- and alluvial-
plain deposits above (Figure 12).
The MFSs of these sequences
are interpreted at major marine
incursions marked by a change
from retrograding to prograd-
ing stacking of marine parase-
quences. In the downdip parts
of the most distal sequences, ma-
rine limestones occur beneath in-
terpreted MFSs (Figure 12).
Lowstand systems tracts pri-
marily consist of incised-valley
systems that are filled by tidal-
influenced sandstones and mud-
stones in the downdip regions
and braided fluvial sandstones
in updip regions (Figure 13A). In
the low-relief, HFSs identified
within the Burgan Formation, on-
lapping lowstand clastic wedges,
as well as basin-floor lowstand
fans are absent. Clastic lowstand
deposition appears limited to
IVF. The valleys become thinner
and more discontinuous down-
dip (Figure 13B) and do not ap-
pear to extend to the maximum
downdip progradational limit of
underlying sandstone-prone
highstands. These relationships
strongly suggest that transgres-
sive erosion has modified the
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 229

FIGURE 14. Paleogeographic map showing the depositional environments of a highstand systems tract in the Burgan
Formation. This paleogeographic map is interpreted from the sequence-stratigraphic architecture of the B600 high-
frequency, depositional sequence, which is a highstand systems tract in the transgressive sequence set of the lower,
third-order composite sequence.

primary distribution and thickness patterns in the sandstones updip (Figure 14). The shorelines in these
most distal (downdip) parts of the Burgan incised- highstands trend northwest– southeast and are best
valley systems. developed in northern Kuwait.
The TSTs correspond to retrogradational succes-
sions of coarsening-upward strata and display a Composite Sequences
systematic downdip to updip change from marine The 19 HFSs identified within the Burgan Forma-
carbonates, to marginal-marine mudstones and sand- tion define four distinct sequence sets that form parts
stones, to mudstone-prone alluvial- and coastal-plain of two composite sequences. The basal four HFSs
deposits. (B900, B850, B800, and B750) stack in an overall ag-
Highstand systems tracts correspond to progada- gradational succession and form the LSS of the
tional successions of coarsening-upward, dominantly lower Burgan composite sequence (Figure 3). This
marginal-marine sandstones and mudstones down- aggradational sequence set is dominated by braided
dip and alluvial- and coastal-plain mudstones and fluvial deposits that contain few mudstone breaks
230 / Strohmenger et al.

(Figure 12). The basal aggradational sequence set is which encompass lithofacies 1B, 2A, 2B, 2C, and 2F,
overlain by six HFSs (B725, B700, B650, B600, B550, exhibit moderately good reservoir-quality properties.
and B500) that stack in an overall retrogradational They have average porosity and permeability values of
pattern, which constitutes the TSS of the lower 23% and 270 md, respectively. The marginal-marine
Burgan composite sequence (Figure 3). In the Raud- deposits, which include lithofacies 2A, 2E, and 3B,
hatain and Sabiriyah field area in northern Kuwait, generally display poorer reservoir quality as a result
the basal HFSs in this retrogradational succession are of variable extents of bioturbation. Average porosity
dominated by nonmarine facies, whereas the upper- and permeability values for these lithofacies are 19%
most sequences contain increasing marginal-marine and 10 md, respectively.
components (Figure 12). The next four HFSs (B450,
B400, B350, and B300) stack in an overall prograda- Reservoir Quality Distribution
tional succession and comprise the HSS of the lower Stratal distributions, thickness variations, and re-
Burgan composite sequence (Figure 3). In the north- gional facies architecture of the Burgan systems tracts
ern fields area, these sequences are dominated by can be interpreted from isopach and paleogeograph-
marginal-marine highstand deposits (Figure 12). The ic maps of these intervals (Figures 13, 14). High-
uppermost five sequences (B250, B200, B150, B125, frequency lowstands of the Burgan Formation consist
and the LST of B100) stack in an aggradational pattern of fluvial and tidal deposits that filled incised valleys.
and form the LSS of the upper Burgan and Mauddud In general, the incised valleys become thinner, nar-
upper composite sequence (Figure 3). In the northern rower, and more tidal influenced downdip to the
fields area, these HFSs consist of alternating non- northeast (Figure 13). Conversely, the valleys become
marine LSTs and marginal-marine TSTs and HSTs thicker, more widespread, and more fluvially influ-
(Figure 12). enced updip to the south and southwest (Figure 13).
Based on this delineation of sequence sets, a com- High-frequency highstands of the Burgan consist of
posite sequence boundary B900_SB is placed at the shoreline deposits, which, in general, trend north-
base of the Burgan Formation (Figure 3). In general, west–southeast, with more marine facies to the north-
the overlying aggradational, retrogradational, and east and more nonmarine facies toward the south-
progradational sequence sets correspond to the LSS, west (Figure 14).
TSS, and HSS of the lower composite sequence. It The sequence-stratigraphic framework presented
should be noted, however, that the high-frequency in this paper provides an improved understanding of
transgressive surface B725_TS and the high-frequency the distribution of sandstone-prone lowstands and
maximum flooding surface B500_MFS are used as marginal-marine highstands in each of the 19 HFSs
composite TS and MFS, respectively, for the lower identified within the Burgan Formation (Figure 12).
composite sequence (Figures 3, 12). The uppermost Furthermore, some additional new play concepts were
aggradational sequence set is interpreted as a second identified. Within the context of the new Mauddud–
LSS. The composite transgressive surface B100_TS is Burgan boundary, distinct isolated incised-valley sys-
interpreted as the TS of this composite sequence, with tems in the clastic member of the Mauddud Forma-
strata in the overlying Mauddud Formation, form- tion can be defined. Within the Burgan proper, the
ing the TSS and the HSS of the younger composite possibility exists that marginal-marine shorelines in
sequence. the TSSs may also contain hydrocarbons off the flank
In general, the various sequence sets are thicker of structures. This combined structural-stratigraphic
and more sandstone prone in the southwest area of trap would depend on mudstone-prone coastal-plain
Kuwait and thinner and more mudstone prone in the deposits in each sequence acting as the updip lateral
northeast (Figures 12 – 14). seal and marine shales in the overlying sequence
acting as the top seal.
Reservoir Quality
Reservoir quality of the Burgan Formation is closely
related to the interpreted depositional environ- MAUDDUD FORMATION
ments. Fluvial-dominated sandstones, which include
lithofacies 1A and 1C, possess the best reservoir- A sequence-stratigraphic framework for the Maud-
quality attributes. These lithofacies types have av- dud Formation has been established using all avail-
erage porosity and permeability values of 25% and able well and core data from Kuwait. The Mauddud
1600 md, respectively. Tidal-dominated sandstones, Formation can be described by the TSS and the HSS
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 231

FIGURE 15. Schematic facies model of the Mauddud transgressive sequence set (upper third-order composite
sequence).

as a third-order composite sequence that is composed (Raudhatain and Sabiriyah fields area). In southern
of eight high-frequency, depositional sequences. The (Greater Burgan field area) and southwestern Kuwait
top (MAU100_TS) and base (B100_TS) of the Maud- (Minagish field area), glauconite-rich, deltaic sand-
dud Formation are marked by TSS (Strohmenger et al., stones are more common.
2002; Demko et al., 2003) (Figure 3).
The lower Mauddud (Mauddud TSS) shows a lat- Lithofacies Types and Depositional Model
eral change in lithology from limestone in northern Sediments of the Mauddud Formation were de-
Kuwait at the Raudhatain and Sabiriyah fields area to posited along a gently northward- and northeast-
siliciclastics in southern (Greater Burgan field area; ward-dipping homoclinal ramp (Strohmenger et al.,
Figure 15) and southeastern Kuwait (Minagish field 2002; Demko et al., 2003).
area). The upper Mauddud (Mauddud HSS) is mostly Lithofacies and lithology of the Mauddud Forma-
eroded in southern and southwestern Kuwait (Stroh- tion of Kuwait vary by geographic area and time. In
menger et al., 2002; Demko et al., 2003) (Figure 16). the Mauddud transgressive sequence set, siliciclastic
The Mauddud Formation in Kuwait comprises deposits of southern (Greater Burgan field area) and
eight carbonate lithofacies and three siliciclastic litho- southwestern Kuwait (Minagish field area) grade into
facies. Carbonate lithofacies were deposited in inner carbonates toward the north and northeast (Abdali,
to lower ramp, normal to slightly restricted environ- Raudhatain, and Sabiriyah fields area; Figure 15).
ments. Clastic lithofacies were deposited in inner- The Mauddud Formation can be described by
ramp (deeper lagoon), nearshore marine, and offshore means of 11 lithofacies types (F1 – F11). The facies
marine environments (Strohmenger et al., 2002; Demko scheme follows, in great parts, the one established by
et al., 2003) (Figures 15, 16), similar to those de- I. Goodall, N. Cross, and D. Payne (1996, personal
scribed for the Burgan Formation. communication), C. Hollis, N. Cross, T. Needham, and
The Cenomanian Wara Formation overlies the B. Jones (1998, personal communication), and N. Cross,
Mauddud Formation and provides a regional top seal R. Heath, and G. Paintal (1999, personal communi-
(Figure 3). It characteristically is a slightly calcareous cation), with some modifications. Lithofacies types
to noncalcareous marine shale in northern Kuwait range from moderate- to high-energy upper-ramp
232 / Strohmenger et al.

FIGURE 16. Schematic facies model of the Mauddud highstand sequence set (upper third-order composite sequence).

deposits (F1, F2, and F3) through medium-energy, F5). The grain composition (occurrence of miliolids),
upper- to middle-ramp (F4), and lagoonal deposits as well as the facies-stacking pattern, suggests the
(F4, F5 and F6) to low-energy, lower-ramp (F7, F8, rudist floatstone-rudstone represents a ramp-crest to
and F11), and deeper lagoonal deposits (F7, F8, F9, inner-ramp, moderate to high-water-energy sediment,
F10, and F11; Figures 15, 16). deposited laterally or mostly lagoonward of lithofa-
The thickness of the Mauddud Formation increases cies F2 (skeletal-peloidal grainstone).
toward the north and northeast and decreases to- Relatively thick accumulations (as much as 60 ft
ward the south and southwest. The thinning is the [18 m]) of rudist floatstone-rudstone occur within
result of reduced accommodation during the lower Mauddud sequences MAU450 and MAU200. Espe-
Mauddud TSS (facies change from carbonates to silici- cially within Mauddud sequence MAU450, rudist
clastics; Figure 15) and pronounced erosion of the floatstone-rudstone forms very porous intervals that
upper Mauddud HSS in southern (Greater Burgan are easily identified by low gamma-ray- and high-
field area) and southwestern Kuwait (Minagish field resistivity-log responses.
area) (Figure 16). Cementation is minor. Dominant cement type is
blocky calcite cement, partly to completely filling the
Lithofacies 1 (F1): Rudist Floatstone to Rudstone molds of dissolved rudist shells.
Lithofacies type F1 (Figure 17A) is rich in rudists or Dominant porosity types are moldic and vuggy
rudist fragments. Miliolids, conical orbitolinids, porosity.
other benthic foraminifera, skeletal fragments, and The average porosity is 18%, and the typical per-
echinoderms are common. Discoidal orbitolinids are meability is 5 md.
uncommon. Nonskeletal grains are peloids. Biotur-
bation is moderate. Lithofacies 2 (L2): Skeletal and Peloidal Grainstone
This lithofacies type commonly grades into the Lithofacies type F2 (Figure 17B) is rich in conical
high-energy skeletal-peloidal grainstone (lithofacies orbitolinids, skeletal fragments, and echinoderms.
F2), as well as into the restricted lagoon nodular, Other benthic foraminifera, red algae, and green algae
miliolid-bearing packstone-wackestone (lithofacies are common. Discoidal orbitolinids and gastropods
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 233

FIGURE 17. Carbonate lithofacies types, slabbed core photographs. (A) Lithofacies F1: rudist floatstone-rudstone.
(B) Lithofacies F2: skeletal-peloidal grainstone showing Glossifungites (Gl) burrows as well as fenestral structures
(keystone vugs). (C) Lithofacies F3: skeletal-peloidal mud-lean packstone showing bioturbation. (D) Lithofacies F4:
bioturbated skeletal-peloidal packstone.

are sparse. Nonskeletal grains are coated grains and trix and the overall facies-stacking pattern suggest
peloids. Bioturbation is low. the skeletal-peloidal grainstone to represent a high-
This lithofacies type frequently shows Glossifun- energy, shoal (ramp-crest) to upper-ramp deposits.
gites burrows and/or, less common, desiccation cracks Cementation varies between minor and extensive.
(Figures 17B, 18A, B), as well as, also more uncom- Typical cement types are neomorphosed isopachous
monly, karstification (Figure 18C). Sedimentary rim cement, blocky calcite cement, and syntaxial cal-
structures indicating intertidal conditions, such as cite cement. Glossifungites burrows as well as karst
trough – cross-bedding, fenestral structures (keystone fillings are commonly dolomitized.
vugs and sheet cracks; Figure 17B), geopedal fillings Dominant porosity type is microporosity. Inter-
(vadose silt), and circumgranular cracks, are general- particle, intraparticle, and moldic porosity are minor
ly limited to lithofacies type F2. porosity types.
This lithofacies generally occurs at the top of the The average porosity is 15%, and the typical per-
Mauddud parasequences and HFSs. The lack of ma- meability is 2 md.
234 / Strohmenger et al.

FIGURE 18. Sedimentary struc-


tures, slabbed core photographs.
(A) Erosional surface with Glossi-
fungites burrows. (B) Glossifungites
burrows and/or desiccation cracks
with younger sediment infill.
(C) Microkarst with sediment
infill.

Cementation generally is mi-


nor. Cement types are blocky cal-
cite and syntaxial calcite cements.
Dominant porosity type is mi-
croporosity. Intraparticle and
moldic porosity are minor po-
rosity types.
The average porosity is 15%,
and the typical permeability is
2 md.

Lithofacies 4 (F4): Bioturbated


Skeletal and Peloidal Packstone
Lithofacies type F4 (Figure 17D)
is rich in conical and discoidal
orbitolinids and echinoderms.
Skeletal fragments are common.
Other benthic foraminifera, ru-
dist fragments, and thin-shelled
bivalves are sparse. Nonskeletal
grains are peloids. Bioturbation
is high.
This lithofacies typically un-
derlies lithofacies type F3 (skeletal-
peloidal mud-lean packstone)
Lithofacies 3 (F3): Skeletal and Peloidal and, less commonly, lithofacies type F5 (nodular,
Mud-lean Packstone miliolid-bearing packstone-wackestone). It commonly
Lithofacies type F3 (Figure 17C) is rich in conical overlies lithofacies type F2 (skeletal-peloidal grain-
orbitolinids. Skeletal fragments and echinoderms are stone) and lithofacies type F3 (skeletal-peloidal mud-
common. Discoidal orbitolinids, other benthic fora- lean packstone). Grain composition, texture, and the
minifera, and gastropods are sparse. Nonskeletal grains overall facies-stacking pattern suggest that the bio-
are peloids. Bioturbation varies from moderate to turbated skeletal-peloidal packstone represents a low-
high. to moderate-energy, upper- to middle-ramp deposit. It
This lithofacies type commonly alternates with may also be present behind the high-energy bar de-
lithofacies type F2 (skeletal-peloidal grainstone). The posits (lithofacies F1, F2, and F3) in an inner-ramp,
relatively low mud content and the overall facies- protected lagoonal environment.
stacking pattern suggests the skeletal-peloidal mud- Cementation is minor to extensive. Dominant ce-
lean packstone to represent a moderate-energy, upper- ment type is blocky calcite cement.
ramp deposit. Dominant porosity type is microporosity. Intrapar-
Like lithofacies type F2 (skeletal-peloidal grain- ticle and moldic porosity are minor porosity types.
stone), the skeletal-peloidal mud-lean packstone lo- The average porosity is 12%, and the typical per-
cally is karstified. meability is 0.5 md.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 235

FIGURE 19. Carbonate lithofacies types, slabbed core photographs. (A) Lithofacies F5: nodular, miliolid-bearing packstone-
wackestone. (B) Lithofacies F6: skeletal wackestone-mudstone (C) Lithofacies F7: clay-rich, bioturbated skeletal wacke-
stone showing Thalassinoides (Th) and Planolites (Pl) burrows. (D) Lithofacies F8: bioturbated glauconitic packstone.

Lithofacies 5 (F5): Nodular, Miliolid-bearing of miliolids), as well as the facies-stacking pattern,


Packstone to Wackestone suggests the nodular, miliolid-bearing packstone-
Lithofacies type F5 (Figure 19A) commonly con- wackestone to represent a low- to moderate-energy,
tains miliolids, conical orbitolinids, other benthic fo- inner-ramp, restricted lagoonal deposit.
raminifera, rudist fragments, skeletal fragments, and Cementation is minor, and the dominant cement
thin-shelled bivalves. Sponge spicules are common type is blocky calcite cement.
and only occur within this lithofacies type. Discoidal The dominant porosity type is microporosity. Intra-
orbitolinids and echinoderms are uncommon. Non- particle and moldic porosity are minor porosity types.
skeletal grains are peloids. Bioturbation and burrowing The average porosity is 19%, and the typical per-
is high, most probably causing its nodular appearance. meability is 0.5 md.
This lithofacies type is restricted to the upper
Mauddud (Mauddud HSS; Figure 16) and commonly Lithofacies 6 (F6): Skeletal Wackestone to Mudstone
over- and underlies the rudist floatstone-rudstone Lithofacies type F6 (Figure 19B) is mud rich, light
(lithofacies F1), predominantly within Mauddud se- gray, with sparse skeletal fragments and peloids. Biotur-
quence MAU200. The grain composition (occurrence bation is very low. Pyrite framboids are quite frequent.
236 / Strohmenger et al.

This lithofacies type is present only in the upper lean packstone). We interpret this lithofacies type as
Mauddud HSS (Mauddud sequence MAU300; Figure 16). representing periods of low sedimentation rates during
We interpret this lithofacies type to represent a low- rapid sea-level rise. It dominantly occurs within the
energy, inner-ramp, protected lagoonal deposit. Mauddud TSS. The depositional environment ranges
Cementation (fine-crystalline, blocky calcite filling from inner ramp, lagoonal (Mauddud TSS; Figure 15)
the pore space of dissolved skeletal fragments) is minor. to middle to lower ramp (Mauddud HSS; Figure 16).
The dominant porosity type is microporosity, and Cementation is minor to extensive. Dominant ce-
intercrystalline porosity is present. ment types are blocky calcite, ferroan dolomite, and
Average porosity and typical permeability (only four ferroan calcite cement. Ferroan cements are most com-
samples analyzed) is within the range of lithofacies F5 mon in southern Kuwait (Greater Burgan field area).
(nodular, miliolid-bearing packstone-wackestone). Dominant porosity type is microporosity. Intrapar-
ticle and moldic porosity are minor porosity types.
Lithofacies 7 (F7): Clay-rich, Bioturbated The average porosity is 7%, and the typical per-
Skeletal Wackestone meability is 0.05 md.
Lithofacies type F7 (Figure 19C) is rich in discoidal
orbitolinids and echinoderms, as well as in trace fossils, Lithofacies 9 (F9): Bioturbated
including Thalassinoides and Planolites. Skeletal frag- Glauconitic Sandstone
ments, thin-shelled bivalves, and pelagic foraminif- Lithofacies type F9 (Figure 20A) is quartz rich and
era are common. Nonskeletal grains are peloids. Bio- contains sparse discoidal orbitolinids and skeletal frag-
turbation is very high (solution-seam rich). Pyrite is ments. It is rich in trace fossils, including Teichichnus,
common. Asterosoma and Terebellina (uncommon). Grain types
This lithofacies typically underlies lithofacies type are detrital quartz, glauconite, pyrite, and siderite. Bio-
F4 (bioturbated skeletal-peloidal packstone) or over- turbation is very high. This lithofacies is similar to
lies lithofacies type F2 (skeletal-peloidal grainstone) lithofacies 2D (glauconitic sandstone) of the Burgan
and lithofacies type F3 (skeletal-peloidal mud-lean Formation but has a higher glauconite content.
packstone). Grain composition, texture, and the over- Like lithofacies type F8 (bioturbated glauconitic
all facies-stacking pattern suggest the clay-rich, biotur- packstone), this lithofacies typically overlies sequence
bated skeletal wackestone to represent a low-energy, boundaries on top of lithofacies types F2 (skeletal-
middle- to lower-ramp deposit. It may, however, also peloidal grainstone), F3 (skeletal-peloidal mud-lean
be present behind the high-energy bar deposits (litho- packstone), and F10 (bioturbated mud-rich sand-
facies F1, F2, and F3) in the inner-ramp, protected stone). It dominantly occurs within the Mauddud
lagoon environment. TSS (Figure 15). The bioturbated glauconitic sandstone
Cementation is minor. The dominant cement type is interpreted to represent an inner-ramp, nearshore-
is fine-crystalline, blocky calcite cement. Dolomiti- marine, lower-shoreface deposit.
zation occurs along solution seams. Cementation is minor to extensive. Dominant ce-
The dominant porosity type is microporosity, and ment types are blocky calcite, ferroan dolomite, and
intraparticle and intercrystalline porosity are minor ferroan calcite cement.
porosity types. Predominant porosity type is microporosity, and in-
The average porosity is 4%, and the typical per- traparticle and intergranular porosity types are present.
meability is 0.01 md. The average porosity is 10%, and the typical per-
meability is 0.3 md.
Lithofacies 8 (F8): Bioturbated
Glauconitic Packstone Lithofacies 10 (F10): Bioturbated
Lithofacies type F8 (Figure 19D) commonly con- Mud-rich Sandstone
tains discoidal orbitolinids, echinoderms, and skeletal Lithofacies type F10 (Figure 20B) is intensely bio-
fragments. Conical orbitolinids, thin-shelled bivalves, turbated and rich in trace fossils, including Teichichnus,
and rudist fragments are uncommon. Nonskeletal Asterosoma, Thalassinoides, Planolites, and Chondrites.
grains are glauconite (glauconitized peloids), peloids, Lithofacies type F10 occurs within the Mauddud
quartz, and pyrite. Bioturbation is high. TSS (Figure 15). Relatively thick intercalations (>60 ft;
This lithofacies type typically overlies sequence >18 m) occur within the Mauddud sequence MAU500.
boundaries on top of lithofacies types F2 (skeletal- This lithofacies is similar to lithofacies 3F (calcareous,
peloidal grainstone) and F3 (skeletal-peloidal mud- bioturbated sandstone) of the Burgan Formation.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 237

FIGURE 20. Siliciclastic lithofa-


cies types, slabbed core pho-
tographs. (A) Lithofacies F9:
bioturbated glauconitic sand-
stone showing Teichichnus (Te)
burrows. (B) Lithofacies F10:
bioturbated mud-rich sand-
stone showing Teichichnus (Te)
and Thalassinoides (Th) bur-
rows. (C) Lithofacies F11: dark-
gray shale and mudstone.

(bioturbated glauconitic sand-


stone; Figure 15). The fact that
it is not highly bioturbated as
it would be expected for off-
shore shales underlying low-
er shoreface deposits (litho-
facies F9 and F10) indicates
a high-stress environment.
The dark-gray shale and mud-
stone are interpreted to rep-
resent a low-energy, deeper
marine, lower-ramp to basin-
al deposit during the Maud-
dud HSS (Figure 16) and the
Wara Formation.
Predominant porosity type
is microporosity.
The bioturbated mud-rich sandstone is interpreted to The average porosity is 5%, and typical perme-
represent an inner-ramp, nearshore-marine, lower- ability is 0.03 md.
shoreface deposit.
Cementation is minor; dominant porosity type is
intergranular. High-frequency Sequences
The average porosity is 12%, and typical perme- Eight high-frequency, depositional sequences were
ability is 11 md. interpreted within the Mauddud Formation. From bot-
tom to top, these are B100 (TST and HST), MAU600,
MAU500, MAU450, MAU400, MAU350, MAU300,
Lithofacies 11 (F11): Dark-gray Shale and MAU200 (Figure 3).
and Mudstone Eight SBs, two TSs, three MFSs, and two FSs were
Lithofacies F11 (Figure 20C) shows only sparse identified. All surfaces, except for one flooding sur-
skeletal fragments, thin-shelled bivalves, planktonic face (MAU600_FS), can be correlated throughout all
foraminifera, and nannofossils. This lithofacies is sim- of Kuwait.
ilar to lithofacies 5A (laminated gray shale) of the Sequence boundaries typically are characterized by
Burgan Formation. Millimeter-scale laminae indi- an abrupt change in texture and lithofacies. Glossi-
cate low bioturbation caused by stressed (lagoon) or fungites burrows and/or desiccation cracks are common
deeper marine, offshore (lower-ramp) environment in the upper part of each high-frequency sequence
of deposition. (Figures 18A, B; 21). Karstification is seen locally
Dominant lithofacies type of the Mauddud trans- (Figure 18C). Evidence of at least periodic exposure,
gressive sequence set (Figure 15); deposited largely in such as fenestral structures (keystone vugs and sheet
an inner-ramp, deeper lagoonal setting, juxtaposed cracks; Figure 17B) and circumgranular cracks, is also
to lithofacies types F7 (clay-rich, bioturbated wacke- present in the upper part of each high-frequency se-
stone), F8 (bioturbated glauconitic packstone), and F9 quence. Sandstone or packstone, rich in glauconite,
238 / Strohmenger et al.

FIGURE 21. Mauddud Formation idealized carbonate parasequence (upper part of a high-frequency sequence)
showing shallowing-upward trend of facies from base to top: bioturbated skeletal-peloidal packstone (F4, blue
color), skeletal-peloidal mud-lean packstone (F3, red color), and skeletal-peloidal grainstone (F2, orange color). Flooding
surface/sequence boundary (FS/SB) is interpreted by Glossifugites burrows or, uncommonly, karstification (microkarst)
and is overlain by bioturbated glauconitic packstone (F8) or bioturbated glauconitic sandstone (F9, green color).

typically immediately overlies each high-frequency Burgan field area; Figure 23), and dominantly coastal-
sequence boundary (Figure 21), recording periods of plain deposits occur in southwestern Kuwait (Min-
rapid flooding and resulting in low sedimentation rate. agish field area; Figure 23). The sequence is partly
The identified high-frequency sequence boundaries eroded by Mauddud sequence boundary MAU600_SB
correspond to combined flooding surface and se- in southern Kuwait (Figure 23).
quence boundaries (FS/SB, Figure 21). The sequence consists of lithofacies F11 (dark-gray
Maximum flooding surfaces are interpreted at major shale and mudstone) and lithofacies F9 (bioturbated
marine incursions marked by a change from retrograd- glauconitic sandstone) grading upward into lithofa-
ing to prograding stacking of marine parasequences. cies F7 (clay-rich, bioturbated skeletal wackestone),
Flooding surfaces are interpreted on top of shallowing- lithofacies F8 (bioturbated glauconitic packstone),
upward parasequences, commonly corresponding to and lithofacies F4 (bioturbated skeletal-peloidal pack-
Glossifungites surfaces (Figure 21). stone) in northern Kuwait (Raudhatain and Sabiriyah
fields area; Figures 22, 23).
Mauddud Sequence B100
(B100_SB to MAU600_SB) Mauddud Sequence MAU600
In northern Kuwait (Raudhatain and Sabiriyah (MAU600_SB to MAU500_SB)
fields area; Figure 22), this sequence is shale and This sequence is composed of mud-dominated car-
mud rich at the base of the TST, grading upward, to bonates in northern Kuwait (Raudhatain and Sabir-
grain rich at the top of the HST. Dominantly tidal- iyah fields area; Figure 22), dominantly coastal-plain
flat deposits occur in southern Kuwait (Greater and tidal-influenced deposits in southern Kuwait
FIGURE 22. Field-scale cross section (northwest – southeast) showing facies distribution and sequence-stratigraphic framework of the Mauddud Formation
throughout the northern Kuwait Raudhatain and Sabiriyah fields. Mauddud transgressive sequence set: B100_TS to MAU400_MFS. Mauddud highstand sequence
set: MAU400_MFS to MAU100_SB. Mauddud lowstand systems tract: MAU100_SB to MAU100_TS. GR = gamma-ray log; FAC = Mauddud carbonate and siliciclastic facies
and Burgan deposits (based on core); MD = measured depth (feet); DT = sonic log; NEU = neutron porosity log; DENS = density log.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 239
240 / Strohmenger et al.

FIGURE 23. Regional cross section oriented north – south and east – west showing facies distribution and sequence-stratigraphic framework of the Mauddud
Formation. Note that chronostratigraphic boundaries (time lines) crosscut the lithostratigraphic boundary between the Mauddud carbonate member (Mauddud
Formation) and the Mauddud clastic member (Burgan Formation, time equivalent to Mauddud Formation). The proposed sequence-stratigraphic correlation
is supported by age dating and palynofacies analyses (T. D. Davies and T. C. Huang, 2000, personal communication). Stratigraphy-diagnostic palynofacies
assemblages are shown as colored dots. Blue dots: only found above Mauddud transgressive surface MAU100_TS (in Wara Formation). Yellow dot: only found between
Mauddud sequence boundary MAU100_SB and Mauddud trangressive surface MAU100_TS (Mauddud lowstand systems tract, onlapping on Mauddud sequence
boundary MAU100_SB in southern and southwestern Kuwait). Green dots: only found between Burgan transgressive surface B100_TS and Mauddud maximum
flooding surface MAU400_MFS (Mauddud transgressive sequence set: lower part of Mauddud carbonate member in northern Kuwait and Mauddud clastic member in
southern and southwestern Kuwait). Red dots: only found below Burgan transgressive surface B100_TS (Burgan lowstand sequence set). No stratigraphy-diagnostic
palynofacies assemblage was found in grain-dominated, shallow-water carbonates between Mauddud maximum flooding surface MAU400_MFS and Mauddud
sequence boundary MAU100_SB (Mauddud highstand sequence set). Color codes for Mauddud carbonate and siliciclastic facies as well as for Burgan siliciclastic
deposits are shown in Figure 22. GR = gamma-ray log; FAC = Mauddud carbonate and siliciclastic facies and Burgan deposits (based on core); MD = measured depth
(feet); DT = sonic log; NEU = neutron porosity log; DENS = density log.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 241

(Greater Burgan field area; Figure 23), and dominantly Mauddud sequence boundary MAU400_SB in south-
coastal-plain and tidal-flat deposits in southwestern ern Kuwait (Greater Burgan field area; Figure 23).
Kuwait (Minagish field area; Figure 23). Lithofacies F9 (bioturbated glauconitic sandstone)
Dominantly lithofacies F11 (dark gray shale and overlies Mauddud sequence boundary MAU400_SB in
mudstone) and minor lithofacies F8 (bioturbated southwestern Kuwait (Minagish field area; Figure 23).
glauconitic sandstone) and lithofacies F7 (clay-rich, This sequence is dominated by lithofacies F4 (bio-
bioturbated skeletal wackestone) occur in the north- turbated skeletal-peloidal packstone) in southern
ern fields area. The succession may show upward in- Kuwait (Greater Burgan field area; Figure 23) and
crease of porosity and permeability and shallowing- dominated by marginal-marine siliciclastics with
upward lithofacies trend, as well as coarsening-upward thin, karstified lithofacies F3 (skeletal-peloidal mud-
texture in the northern fields area (Figures 21, 22). lean packstone) at the top in southwestern Kuwait
(Minagish field area; Figure 23). It is partly erod-
Mauddud Sequence MAU500 ed in southwestern Kuwait (Minagish field area;
(MAU500_SB to MAU450_SB) Figure 23).
Marginal-marine (distal lower shoreface) siliciclas- The sequence is composed predominantly of litho-
tics occur in northern Kuwait (Raudhatain and Sabi- facies F2 (skeletal-peloidal grainstone) and litho-
riyah fields area) and were not eroded by sequence facies F3 (skeletal-peloidal mud-lean packstone), as
boundary MAU450_SB (Figure 22). Dominantly coastal- well as minor lithofacies F4 (bioturbated skeletal-
plain deposits occur in southern Kuwait (Greater Burgan peloidal packstone) and lithofacies F7 (clay-rich, bio-
field area; Figure 23) and dominantly tidal-influenced turbated skeletal wackestone) in the northern fields
to fluvial deposits occur in southwestern Kuwait (Mi- area (Figures 22, 23).
nagish field area; Figure 23).
Mauddud Sequence MAU350
Mauddud Sequence MAU450 (MAU350_SB to MAU300_SB)
(MAU450_SB to MAU400_SB) This sequence is composed predominantly of litho-
This sequence is composed of grain-dominated facies F2 (skeletal-peloidal grainstone) and litho-
carbonates in northern Kuwait (Raudhatain and facies F3 (skeletal-peloidal mud-lean packstone), as
Sabiriyah fields area; Figure 22) and dominated by well as minor lithofacies F4 (bioturbated skeletal-
marginal-marine (distal lower shoreface) deposits in peloidal packstone) and intercalations of lithofacies F1
southern (Greater Burgan field area; Figure 23) and (rudist floatstone-rudstone) in northern Kuwait (Raud-
southwestern Kuwait (Minagish field area; Figure 23). hatain and Sabiriyah fields area; Figure 22). It is partly
The sequence is composed predominantly of litho- eroded in the Greater Burgan field area (lithofacies F4:
facies F2 (skeletal-peloidal grainstone), with minor skeletal-peloidal packstone; Figure 23) and eroded
lithofacies F3 (skeletal-peloidal mud-lean packstone) (not present) in southwestern Kuwait (Minagish field
and lithofacies F4 (bioturbated skeletal-peloidal pack- area; Figure 23).
stone). Relatively thick accumulations (as much as This sequence may show an upward increase of
60 ft [18 m]) of highly porous lithofacies F1 (rudist porosity and permeability and shallowing-upward
floatstone-rudstone) occur as aerially restricted inter- lithofacies trend as well as coarsening-upward tex-
calations throughout the northern fields area. This ture in the northern fields area (Figures 21, 22).
sequence may show an upward increase of porosity
and permeability and a shallowing-upward lithofacies Mauddud Sequence MAU300
trend, as well as coarsening-upward texture (northern (MAU300_SB to MAU200_SB)
fields area; Figures 21, 22). This sequence is composed predominantly of litho-
facies F2 (skeletal-peloidal grainstone) and lithofa-
Mauddud Sequence MAU400 cies F3 (skeletal-peloidal mud-lean packstone), as
(MAU400_SB to MAU350_SB) well as minor lithofacies F4 (bioturbated skeletal-
This sequence is composed of grain-dominated car- peloidal packstone) and lithofacies F1 (rudist float-
bonates in northern Kuwait (Raudhatain and Sabiri- stone-rudstone) in northern Kuwait (Raudhatain and
yah fields area), with lithofacies 8 (bioturbated glau- Sabiriyah fields area; Figure 22). The sequence is eroded
conitic packstone) commonly overlying Mauddud (not present) in southern (Greater Burgan field area;
sequence boundary MAU400_SB (Figure 22). Litho- Figure 23) and southwestern Kuwait (Minagish field
facies F8 (bioturbated glauconitic packstone) overlies area; Figure 23).
242 / Strohmenger et al.

Mauddud Sequence MAU200 trast, the sequence-stratigraphic boundary between the


(MAU200_SB to MAU100_SB) Mauddud Formation and the underlying Burgan For-
This sequence is composed predominantly of mation occurs approximately 100 ft (30 m) deeper in
lithofacies F5 (nodular, miliolid-bearing packstone- the clastic section of the Mauddud Formation. The
wackestone) and lithofacies F3 (skeletal-peloidal lithostratigraphic boundary between the Mauddud
mud-lean packstone), as well as minor lithofacies F4 carbonate member and the Mauddud clastic member
(bioturbated skeletal-peloidal packstone) and litho- is highly diachronous (Figure 23). The thickness rela-
facies F1 (rudist floatstone-rudstone) in northern Ku- tionships of the two Mauddud Members are recipro-
wait (Raudhatain and Sabiriyah fields area; Figure 22). cal. The Mauddud carbonate member becomes very
The sequence is eroded (not present) in southern thin in locations where the Mauddud clastic member
(Greater Burgan field area; Figure 23) and southwest- is thick. Toward the north (Raudhatain and Sabiriyah
ern Kuwait (Minagish field area; Figure 23). fields area), the Mauddud clastic member becomes
very thin to absent, whereas the Mauddud carbonate
Mauddud Interval MAU100_SB to MAU100_TS member is thinning toward southern Kuwait (Great-
This interval is composed predominantly of litho- er Burgan field area; Figure 23) and southwestern
facies F11 (gray shale and mudstones) and lithofacies Kuwait (Minagish field area; Figure 23).
F7 (clay-rich, bioturbated skeletal wackestone). The sequence-stratigraphic correlation of the
The interval was deposited during the third-order Upper Burgan and the Mauddud presented here is
relative sea-level rise above Mauddud sequence bound- supported by biostratigraphic and palynofacies data
ary MAU100_SB (Figures 22, 23) and is not present (T. D. Davies and T.-C. Huang, 2000, personal com-
in southern (Greater Burgan field area; Figure 23) and munication) (Figure 23).
southwestern Kuwait (Minagish field area; Figure 23).
Reservoir Quality Distribution
Composite Sequence In general, most Mauddud lithofacies have mod-
The TSS of the Mauddud composite sequence starts erate porosity and overall low permeability. Litho-
at the base of Mauddud transgressive surface B100_TS facies types F2, F3, F4, F5, and F6 have essentially the
and is bounded on the top by Mauddud maximum same porosity-permeability distributions. Only litho-
flooding surface MAU400_MFS of the Mauddud high- facies type F1 (rudist floatstone-rudstone) and litho-
frequency sequence MAU400. The overlying HSS facies F10 (bioturbated, mud-rich sandstone) show
is bounded on top by Mauddud sequence boundary higher porosity and/or permeability values. Intervals
MAU100_SB. The top of the Mauddud Formation is that show enhanced reservoir quality are related to
marked by the transgressive surface MAU100_TS fracturing and faulting (enhanced permeability), the
(Figures 22, 23) that merges with the Mauddud se- occurrence of microkarst (Figure 18C), the occurrence
quence boundary MAU100_SB toward southern and of rudist floatstone-rudstone (lithofacies F1: Maud-
southwestern Kuwait (Figure 23). dud high-frequency sequence MAU450; Figure 17A),
The Mauddud TSS shows a lateral change in litho- or the occurrence of the bioturbated, mud-rich sand-
logy from limestone in northern Kuwait (Raudhatain stone (lithofacies F10: Mauddud high-frequency se-
and Sabiriyah fields area; Figure 22) to siliciclastics in quence MAU500; Figure 20B).
southern (Greater Burgan field area; Figure 23) and
southwestern Kuwait (Minagish field area; Figure 23). CONCLUSIONS
An overall shoaling-upward or progradational sig-
nature characterizes the Mauddud HSS. The HSS is Sequence-stratigraphic and biostratigraphic analy-
interpreted to be eroded down to and below sequence ses indicate that the traditional lithostratigraphic
boundary MAU350_SB in southern (Greater Burgan Burgan-Mauddud contact is actually a time-transgres-
field area; Figure 23) and southwestern Kuwait (Mi- sive boundary (Figure 23). The uppermost Burgan and
nagish field area; Figure 23). overlying Mauddud formations form a third-order
Our sequence-stratigraphic interpretation of the composite sequence that becomes more siliciclastic
Mauddud Formation contrasts with the previously prone toward the south (in the Greater Burgan field
used lithostratigraphic correlation. Lithostratigraphic area) and toward the southwest (in the Minagish field
correlation places the boundary between the Maud- area; Figure 23) and carbonate prone to the north
dud Formation and the underlying Burgan Forma- and northeast of Kuwait near the Raudhatain and
tion at the base of the deepest limestone bed. In con- Sabiriyah fields (Figures 22, 23). A lower, third-order
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 243

composite sequence encompasses the remainder of The Mauddud highstand sequence set is carbonate
the underlying Burgan Formation (Figure 12). It com- prone and is mostly eroded in southern and south-
prises marginal-marine deposits to the northeast western Kuwait, resulting in a significant southward
and nonmarine deposits to the southwest of Kuwait and southwestward thinning (Figures 16, 23).
(Figure 12). A total of eight carbonate lithofacies and three
Within the Burgan Formation, 19 high-frequency, siliciclastic lithofacies types were identified within
depositional sequences are recognized and stack to the Mauddud Formation. Carbonate lithofacies were
form one third-order composite sequence and the deposited in inner- to lower-ramp, normal- to slightly
LSS of a younger third-order composite sequence restricted-marine environments. Siliclastic lithofa-
(Figures 3, 12). Lowstand systems tracts consist pri- cies were deposited in a fairly wide range of environ-
marily of fluvial, incised-valley-fill deposits in southern ments, including deeper lagoon, nearshore-marine,
and western Kuwait that become thinner, laterally and offshore-marine (Mauddud carbonate member),
more discontinuous, and more tidal-influenced in a as well as coastal-plain, fluvial, and tidal, (Mauddud
downdip direction to the northeast (Figure 13). The clastic member) environments.
TSTs display a systematic updip to downdip change In general, most Mauddud lithofacies have mod-
from mud-prone alluvial- and coastal-plain strata to erate porosity and low permeability, with micropo-
marginal-marine mudstones and sandstones to ma- rosity as the dominant pore type. Intervals with en-
rine carbonates. The HSTs are dominated by updip hanced reservoir quality can be related to fracturing
alluvial- and coastal-plain mudstones and sandstones and faulting, the occurrence of microkarst (Figure 18C),
and marginal-marine sandstones and mudstones the presence of rudist floatstone-rudstone (Figure 17A),
downdip (Figure 14). and the distribution of intercalated mud-rich sand-
Facies analysis of the Burgan Formation reveals the stone (Figure 18A). Within the context of the new
presence of 22 lithofacies, which are delineated based Mauddud – Burgan chronostratigraphic boundary,
on unique physical criteria, including grain size, com- distinct isolated incised-valley systems can be defined
position, sorting, grading, and physical and biogenic within the Mauddud clastic member. These IVF are
sedimentary structures. These lithofacies describe six expected to have reservoir characteristics similar to
distinct facies, which form the basic building blocks those of the underlying Burgan Formation.
of two facies associations (Figure 4). The proposed sequence-stratigraphic framework
The primary reservoirs in the Burgan Formation and the sequence-stratigraphy-keyed facies scheme
are fluvial and tidal deposits of the incised-valley-fill result in a predictable distribution of reservoir and
facies association (Figure 4). Minor potential reser- seal facies and allow for a better prediction of the
voirs are shoreline sandstones of the TSTs and HSTs vertical and lateral distribution of reservoir quality
(Figure 4). In addition, the possibility exists that and reservoir continuity at both field scale and re-
marginal-marine shorelines in the TSSs may contain gional scale.
hydrocarbons off the flanks of present-day structures.
This combination of structural and stratigraphic trap ACKNOWLEDGMENTS
would be dependent on mud-prone coastal-plain sedi-
ments in each sequence acting as the updip lateral The authors gratefully acknowledge the manage-
seal and marine shales in the overlying sequence ments of ExxonMobil Exploration Company, Exxon-
serving as a (very high-risk) top seal. Mobil Upstream Research Company, and Kuwait Oil
The Mauddud Formation is sequence stratigraphi- Company (KOC) for their permission to publish this
cally subdivided into eight high-frequency deposi- paper. For valuable discussions, we thank Linda W.
tional sequences (Figures 3, 22, 23). The identified Corwin (ExxonMobil), Daniel H. Cassiani (Exxon-
high-frequency SBs, TSs, MFSs, and FS can be cor- Mobil), Kathleen M. McManus (ExxonMobil), Menahi
related regionally throughout Kuwait. The Mauddud Al-Anzi (KOC), Ahmed Al-Eidan (KOC), and Moham-
transgressive sequence set displays a lateral change in med Al-Ajmi (KOC). We extend our thanks to David
lithology from limestone in northern Kuwait (Raud- Awwiller (ExxonMobil) for his detailed reservoir-
hatain and Sabiriyah fields area; Figure 22) to silici- quality analyses and to Tom D. Davies and Ting-
clastics in southern (Greater Burgan field area) and Chang Huang for performing the biostratigraphic
southwestern Kuwait (Minagish field area; Figures 15, and palynofacies analyses. Dolores A. Claxton (Exxon-
23), interfingering with what has traditionally been Mobil) is thanked for drafting the figures. We extend
recognized as the Burgan Formation (Figures 3, 23). special thanks to Arthur D. Donovan (BP; formerly
244 / Strohmenger et al.

with ExxonMobil) for his valuable contributions to donald, ed., Sedimentation, tectonics and eustasy: Sea-
this study. The authors greatly appreciate the thor- level changes at active margins: International Asso-
ciation of Sedimentologists Special Publication 12,
ough and thoughtful reviews of J. R. (Rick) Sarg and
p. 3 – 39.
James McGovney.
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology
of fluctuating sea levels since the Triassic: Science,
REFERENCES CITED v. 235, p. 1156 – 1167.
Haq, B. U., J. Hardenbol, and P. R. Vail, 1988, Mesozoic
Adasani, M., 1965, The Greater Burgan field: Fifth Arab and Cenozoic chronostratigraphy and cycles of sea-
Petroleum Congress, Cairo, p. 7 – 27. level change, in C. K. Wilgus, B. S. Hastings, C. G. St. C.
Adasani, M., 1967, The northern Kuwait oil fields: Sixth Kendall, H. W. Posamentier, C. A. Ross, and J. C. Van
Arab Petroleum Congress, Baghdad, p. 1 – 39. Wagoner, eds., Sea-level changes: An integrated
Al-Anzi, M., 1995, Stratigraphy and structure of the Bahra approach: SEPM Special Publication 42, p. 71 – 108.
field Kuwait, in M. I. Al-Husseini, ed., The Middle East Kaufman, R. L., H. Dashti, C. S. Kabir, J. M. Pederson, M. S.
petroleum geosciences (Geo94): Selected Middle East Moon, R. Quttainah, and H. Al-Wael, 1997, Charac-
papers from the Middle East Geoscience Conference: terizing the Greater Burgan field: Use of geochemistry
Bahrain, Gulf PetroLink, v. 1, p. 53 – 64. and oil fingerprinting: Society of Petroleum Engineers
Al-Eidan, A. J., W. B. Wethington, and R. B. Davies, 2001, 10th SPE Middle East Oil Show and Conference,
Upper Burgan reservoir distribution, northern Kuwait: Bahrain, SPE Paper 37803, p. 385 – 394.
Impact on reservoir development: GeoArabia, v. 6, Milton, D. I., and C. C. S. Davies, 1965, Exploration and
no. 2, p. 179 – 208. development of the Raudhatain field: Journal of the
Al-Rawi, M. M., 1981, Geological interpretation of the oil Institute of Petroleum, v. 51, no. 493, p. 17 – 28.
entrapment in the Zubair Formation, Raudhatain field: Mitchum Jr., R. M., 1977, Seismic stratigraphy and global
Society of Petroleum Engineers, Middle East Oil Tech- changes in sea level: Part 11 — Glossary of terms used
nical Conference, Bahrain, SPE Paper 9591, p. 149–159. in seismic stratigraphy, in C. E. Payton, ed., Seismic
Alsharhan, A. S., and A. E. M. Nairn, 1993, Carbonate plat- stratigraphy — Applications to hydrocarbon explora-
form models of Arabian Cretaceous reservoirs, in J. A. T. tion: AAPG Memoir 26, p. 205 – 212.
Simo, R. W. Scott, and J.-P. Masse, eds., Cretaceous Mitchum Jr., R. M., and J. C. Van Wagoner, 1991, High-
carbonate platforms: AAPG Memoir 56, p. 173 – 184. frequency sequences and their stacking patterns:
Alsharhan, A. S., and A. E. M. Nairn, 1997, Sedimentary Sequence-stratigraphic evidence of high-frequency
basins and petroleum geology of the Middle East: eustatic cycles, in K. T. Biddle and W. Schlager, eds.,
Amsterdam, Elsevier, 843 p. The record of sea-level fluctuations: Sedimentary Ge-
Bou-Rabee, F., 1996, Geologic and tectonic history of ology, v. 70, p. 131 – 160.
Kuwait as inferred from seismic data: Journal of Pe- Mitchum Jr., R. M., P. R. Vail, and S. Thompson III, 1977,
troleum Science and Engineering, v. 16, p. 151 – 167. Seismic stratigraphy and global changes of sea level:
Brennan, P., 1990a, Raudhatain field, Kuwait, Arabian Part 2 — The depositional sequence as a basic unit for
basin, in E. A. Beaumont and N. H. Foster, eds., AAPG stratigraphic analysis, in C. E. Payton, ed., Seismic
Treatise of Petroleum Geology, Atlas of Oil and Gas stratigraphy — Applications to hydrocarbon explora-
Fields: Structural Traps 1, p. 187 – 210. tion: AAPG Memoir 26, p. 53 – 62.
Brennan, P., 1990b, Greater Burgan field, Kuwait, Arabian Pemberton, G. S., J. A. MacEachern, and R. W. Frey, 1992a,
basin, in E. A. Beaumont and N. H. Foster, eds., AAPG Trace fossil models; environmental and allostrati-
Treatise of Petroleum Geology, Atlas of Oil and Gas graphic significance, in R. G. Walker and J. P. Noel,
Fields: Structural Traps 1, p. 103 – 128. eds., Facies models; response to sea level change: Geo-
Carman, G. J., 1996, Structural elements of onshore Kuwait: logical Association of Canada, p. 47 – 72.
GeoArabia, v. 1, no. 2, p. 239 – 266. Pemberton, G. S., J. C. Van Wagoner, and G. D. Wach,
Christian, L., 1997, Cretaceous subsurface geology of the 1992b, Ichnofacies of a wave-dominated shoreline:
Middle East region: GeoArabia, v. 2, no. 3, p. 239 – 256. SEPM Core Workshop 17, p. 339 – 382.
Demko, T. M., P. E. Patterson, H. R. Feldman, C. J. Stroh- Sarg, J. F., 1988, Carbonate sequence stratigraphy, in C. K.
menger, J. C. Mitchell, P. J. Lehmann, G. Alsahlan, Wilgus, B. S. Hastings, C. G. St. C. Kendall, H. W.
H. Al-Enezi, and M. Al-Anezi, 2003, Sequence stratig- Posamentier, C. A. Ross, and J. C. Van Wagoner, eds.,
raphy and reservoir architecture of the Burgan and Sea-level changes: An integrated approach: SEPM
Mauddud formations (Lower Cretaceous), Kuwait (abs.): Special Publication 42, p. 155 – 181.
AAPG Annual Meeting Program, v. 5, p. A39. Sarg, J. F., J. R. Markello, and L. J. Weber, 1999, The second-
Fox, A. F., 1961, The development of the southeastern order cycle, carbonate-platform growth, and reservoir,
Kuwait oil fields: Institute of Petroleum Review, v. 15, source, and trap prediction, in P. M. Harris, A. H. Saller,
no. 180, p. 373 – 379. and J. A. T. Simo, eds., Advances in carbonate sequence
Haq, B. U., 1991, Sequence stratigraphy, sea-level change, stratigraphy: Application to reservoirs, outcrops and
and significance for the deep sea, in D. I. M. Mac- models: SEPM Special Publication 63, p. 11 – 34.
Sequence Stratigraphy and Reservoir Architecture of Burgan and Mauddud Formations / 245

Strohmenger, C. J., T. M. Demko, J. C. Mitchell, P. J. Van Wagoner, J. C., R. M. Mitchum Jr., H. W. Posamentier,
Lehmann, H. R. Feldman, A. Douban, A. J. Al-Eidan, and P. R. Vail, 1987, Seismic stratigraphy interpreta-
G. Alsahlan, and H. Al-Enezi, 2002, Regional sequence tion using sequence stratigraphy: Part 2 — Key defini-
stratigraphic framework for the Burgan and Mauddud tions of sequence stratigraphy, in A. W. Bally, ed., Atlas
formations (Lower Cretaceous, Kuwait): Implications of seismic stratigraphy, v. 1: AAPG Studies in Geology 27,
for reservoir distribution and quality (abs.): GeoArabia, p. 11–14.
v. 7 , no. 2, p. 304. Van Wagoner, J. C., H. W. Posamentier, R. M. Mitchum Jr.,
Vail, P. R., 1987, Seismic stratigraphy interpretation using P. R. Vail, J. F. Sarg, T. S. Loutit, and J. Hardenbol, 1988,
sequence stratigraphy: Part 1 — Seismic stratigraphy An overview of the fundamentals of sequence stratig-
interpretation procedure, in A. W. Bally, ed., Atlas of raphy and key definitions, in C. K. Wilgus, B. S. Hastings,
seismic stratigraphy, v. 1: AAPG Studies in Geology 27, C. G. St. C. Kendall, H. W. Posamentier, C. A. Ross,
p. 1 – 10. and J. C. Van Wagoner, eds., Sea-level changes: An in-
Vail, P. R., R. M. Mitchum Jr., R. G. Todd, J. M. Widmier, tegrated approach: SEPM Special Publication 42, p. 39 –
S. Thompson III, J. B. Sangree, J. N. Bubb, and W. G. 46.
Hatlelid, 1977, Seismic stratigraphy and global changes Van Wagoner, J. C., R. M. Mitchum, K. M. Campion, and
of sea level: Part 1— Overview, in C. E. Payton, ed., V. D. Rahmanian, 1990, Siliciclastic sequence stratig-
Seismic stratigraphy— Applications to hydrocarbon ex- raphy in well logs, cores, and outcrop: AAPG Methods
ploration: AAPG Memoir 26, p. 49– 212. in Exploration Series 7, 55 p.
Vail, P. R., F. Audemard, S. A. Bowman, P. N. Eisner, and Walker, R. G., and A. G. Plint, 1992, Wave- and storm-
C. Perez-Cruz, 1991, The stratigraphic signatures of tec- dominated shallow marine systems, in R. G. Walker
tonics, eustasy and sedimentology — An overview, in and N. P. James, eds., Facies models — Response to
G. Einsele, W. Ricken, and A. Seilacher, eds., Cycles and sea-level change: Geological Association of Canada,
events in stratigraphy: Berlin, Springer, p. 617 – 659. p. 219 – 238.
7
Dull, D. W., R. A. Garber, and W. S. Meddaugh, 2006, The sequence
stratigraphy of the Maastrichtian (Upper Cretaceous) reservoir at Wafra
field, Partitioned Neutral Zone, Saudi Arabia and Kuwait: Key to reservoir
modeling and assessment, in P. M. Harris and L. J. Weber, eds., Giant
hydrocarbon reservoirs of the world: From rocks to reservoir characteriza-
tion and modeling: AAPG Memoir 88/SEPM Special Publication, p. 247 –279.

The Sequence Stratigraphy of the


Maastrichtian (Upper Cretaceous)
Reservoir at Wafra Field, Partitioned
Neutral Zone, Saudi Arabia and
Kuwait: Key to Reservoir Modeling
and Assessment
Dennis W. Dull
Chevron Energy Technology Company, Houston, Texas, U.S.A.

Raymond A. Garber
Chevron Energy Technology Company, Houston, Texas, U.S.A.

W. Scott Meddaugh
Chevron Energy Technology Company, Houston, Texas, U.S.A.

ABSTRACT

T
he Maastrichtian (Upper Cretaceous) reservoir is one of five prolific oil res-
ervoirs in the giant Wafra oil field. Although discovered and first produced
in 1959, the reservoir is currently in early development because of its low
but variable oil gravity, high sulfur content, relatively high water cut, and appar-
ent compartmentalization. This made it a much less attractive resource than other
productive intervals at Wafra field. Less than 1% of the original oil in place in the
Maastrichtian has been produced.
The Maastrichtian oil production is largely from subtidal dolomites at an
average depth of 760 m (2500 ft). Carbonate deposition occurred on a very gently
dipping, shallow, arid, and restricted ramp setting that transitioned between
normal-marine conditions to restricted lagoonal environments. The average po-
rosity of the reservoir interval is about 15%, although productive zones have po-
rosity values as much as 30–45%. The average permeability of the reservoir in-
terval is about 30 md; individual core plugs have measured permeability as much

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215879M883030

247
248 / Dull et al.

as 1200 md. This study was undertaken to (1) determine reservoir volumetrics,
(2) understand the areal and stratigraphic distribution of intervals likely to yield
higher volumes of better quality oil, and (3) provide a reservoir property model for
use in fluid-flow simulation.
The key to modeling the reservoir was the construction of an appropriately
detailed sequence-stratigraphic framework for use in building a geostatistical res-
ervoir model with high-quality descriptions from five cored wells in the reservoir.
The geostatistical model of the Maastrichtian reservoir demonstrates the
layered and compartmentalized nature of the reservoir and clearly shows that
the location of the reservoir facies in the Maastrichtian is controlled by the
original depositional fabric and subsequent dolomitization, both of which have
been influenced by the paleotopography. Such understanding is critical to ef-
ficiently develop the 1.5 billion bbl of Maastrichtian oil at Wafra field.

INTRODUCTION cated south and west of Wafra field in March 1959. The
first Maastrichtian production at Wafra field occurred
The Wafra field is located in the Partitioned Neutral in November 1959. The field has produced about
Zone (PNZ) between Kuwait and Saudi Arabia. Figure 1 20 million bbl of oil primarily from two wells. Figure 3
shows the location of the major oil fields in the PNZ. provides a structure map for the Maastrichtian show-
Figure 2 provides a generalized stratigraphic column ing the location of the Maastrichtian producing wells,
for the PNZ showing the five producing horizons at five wells with core data used to build the sequence-
Wafra field. Oil was discovered in the Upper Cretaceous stratigraphic interpretation, and the 123 wells used
Maastrichtian in the PNZ at the South Fuwaris field lo- to generate the geostatistical reservoir model.

FIGURE 1. Location of the Maastrichtian reservoir, Wafra field, Partitioned Neutral Zone (PNZ). S = south. Also shown are
the other producing fields, S. Umm Gudair, S. Fuwaris, and Humma in the PNZ.
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 249

 the integration of the


cores with the well logs
for extrapolating the high-
frequency sequence (HFS)
correlations beyond the
cores in constructing the
reservoir model
 the impact of the data
integration in a three-
dimensional (3-D) Earth
model to provide insight
into possible varying oil-
water contacts and reser-
voir compartmentalization
 the importance of se-
quence stratigraphy and
3-D modeling in identify-
ing the reservoir potential
of the Wafra field Maas-
trichtian reservoir

REGIONAL
STRATIGRAPHY

The Maastrichtian inter-


val in the Wafra field con-
sists of the Upper Cretaceous
Tayarat Formation, which is
the uppermost part of the
Aruma Group and was de-
posited on a stable platform
(Figure 4). The geological set-
ting for the Maastrichtian
carbonates of the Arabian
shield includes (1) the Arabo-
Nubian shield, situated in
Saudi Arabia, covering sev-
eral hundreds of kilometers
in width; (2) a zone of clastic
sediments that rim the east-
ern margin of the shield con-
FIGURE 2. Stratigraphic column for the PNZ showing the five producing reservoirs of
taining both continental and
Wafra field.
shallow-marine siliciclastics,
sourced from the shield; and
The objectives of this chapter are to demonstrate (3) a carbonate shelf lying seaward of the clastic belt,
which was the site of shallow shelf carbonate sedi-
 the importance of core in defining the sequence- mentation (Harris et al., 1984). This shelf contains
stratigraphic framework and depositional intrashelf basins, which were filled with shales and
environment deep-water limestones.
 the methodology used to define the sequence- The Upper Cretaceous sequence was deposited dur-
stratigraphic framework ing a period of mild tectonic activity and is variable in
250 / Dull et al.

FIGURE 3. Structure map on top of high-frequency sequence (HFS) M00 (first Maastrichtian shale) constructed from
3-D seismic data. The contour interval is 50 ft (15 m). The map shows production from 35 wells at an average depth of
2700 ft (822 m). Map shows well locations used in the modeling, Maastrichtian producing wells, and cored well
locations.

regional thickness and facies. Most of the tectonism evaporite solution (Harris et al., 1984). A lithostrati-
affected areas to the east of the PNZ, particularly in the graphic correlation chart is found in Figure 5.
northern Oman Mountains (Harris et al., 1984). A re- In outcrop, the Aruma is treated as a formation
gional unconformity is observed at the top of the Cre- and is divided into three members from base to top:
taceous (upper Maastrichtian), and moldic porosity is Khanasir Limestone Member, Hajajah Limestone
commonly developed at this contact with the overly- Member, and Lina Shale Member (El-Nakhal and El-
ing Tertiary. In some places, the upper Maastrichtian Naggar, 1994; Philip et al., 2002). The lower two mem-
is characterized by large collapse breccias formed by bers are believed to be of Maastrichtian age, whereas
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 251

FIGURE 4. Map of the Arabian


Peninsula showing location of
the Partitioned Neutral Zone
(PNZ) in relation to major
structural features.

Gulf (El-Nakhal and El-Naggar,


1994) (Figure 6). In wells in
the Wafra field, the Tayarat
Formation is about 400 m
(1300 ft) thick, whereas in out-
crop, the Tayarat equivalent
is about 160 m (524 ft) thick.
The Aruma Formation un-
conformably overlies sand-
stones of the Wasia Forma-
tion and is overlain by dense
argillaceous limestone and
thin pyritic shales of the Umm
er Radhuma Formation.
As the Aruma succession
thickens from outcrop to the
subsurface to the east around
the Persian Gulf, the type lo-
the uppermost member is believed to be of Paleocene cality is not complete (Alsharhan and Nairn, 1990).
age based on biostratigraphic analysis. In the subsur- In addition, different formation names have been
face, the Aruma is given group rank, and in Kuwait, it used, and it has been subdivided differently in dif-
is divided into three formations (Gudair, Bahrah, and ferent places. Consequently, the Aruma is typically
Tayarat) (El-Nakhal and El-Naggar, 1994). The Aruma referred to as a group and may include anywhere from
increases in thickness from the outcrop areas in the two to six formations (Alsharhan and Nairn, 1990). As
west into the subsurface at the edges of the Persian noted earlier, the Maastrichtian interval in the Wafra

FIGURE 5. Lithostratigraphic correlation of the Upper Cretaceous in the Arabian Peninsula (Alsharhan and Nairn, 1990).
252 / Dull et al.

FIGURE 6. Schematic cross section showing Aruma surface to subsurface relationships (El-Nakhal and El-Naggar, 1994).

field in the PNZ is found within the Tayarat Forma- intercrystalline porosity in the dolomites (Alsharhan
tion, a part of the Aruma Group. and Kendall, 1995).
In Abu Dhabi, the Shah field produces minor oil During the Maastrichtian, the Arabian Peninsula
and gas-condensate from the Upper Cretaceous Aruma was located about 128 north of the equator on a rel-
Group, which consists of Coniacian to Maastrichtian atively stable continental margin on the African plate
lithologies (Alsharhan and Kendall, 1995). There, the as it moved north toward the Asian plate (Golonka
Aruma Group consists of basal Laffan Formation (Co- et al., 1994; Morris et al., 2002). The Neotethys fore-
niacian) containing mainly slightly calcareous shales, deep narrowed during the Late Cretaceous, and the
which is overlain by the Halul Formation (Santonian), paleogeographic reconstruction by Golonka et al.
which consists of an argillaceous limestone. This is (1994) shows the development of a land mass in the
overlain by the Fiqa Formation (Campanian), which northeast in part of the present-day Arabian Pen-
consists of alternating calcareous shales, marls, and insula caused by the continued northward drift of
argillaceous limestone. These are capped by the Maas- the Arabian plate (Sharland et al., 2001; Morris et al.,
trichtian Simsima Formation containing fossilifer- 2002). The work of Sayed and Mersal (1998) in Jebel
ous and shaly limestone and dolomitic limestone rim- Rawdah in the northern Oman Mountains (a west-
ming the Omani foredeep (Alsharhan and Kendall, northwest- to east-southeast-trending, postobduction
1995). fold) shows that the Oman Mountains were emer-
The Simsima Formation, equivalent to the Aruma gent in the early Maastrichtian. Smith et al. (1995)
Formation in Oman, is interpreted to have been de- placed the ophiolite in the Oman Mountains as latest
posited on a shallow carbonate platform. In the Campanian to Maastrichtian, whereas later work by
Shah field, it consists of two large-scale deposition- Schreurs and Immenhauser (1999) showed that the
al cycles. The lower cycle consists of a coarsening- ophiolite was deposited in late Maastrichtian to Pa-
upward wackestone-mudstone and rudistid packstone- leocene. The proximity of a landmass to the north-
grainstone, with finely crystalline dolomite and east of the PNZ provides a likely source of the shale
minor anhydrite deposited in intertidal to supra- in the second Maastrichtian interval. Alsharhan and
tidal environments. The upper unit is composed of a Nasir (1996), in their research in the western Oman
vertical succession of fining-upward sequences con- Mountains, showed that the Late Cretaceous Qalah
taining packstone, wackestone, shale, and very finely formation, which uncomformably onlaps the Sim-
crystalline dolomite deposited in a subtidal to inter- sima Formation, indicates that the obducted Semail
tidal setting. It becomes more restricted at the top. ophiolite was extensively weathered in a tropical en-
The subsurface Simsima commonly contains inter- vironment. Skelton et al. (1990), in their study of the
particle, moldic, and vuggy porosity with locally good Qalah and Simsima Formations, suggest the Semail
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 253

ophiolite was emergent, and that a more open shelf logical interpretation as summarized in Table 1. Anhy-
environment exists. The absence of evaporites in the drite occurs throughout the reservoir interval as nod-
second Maastrichtian interval is consistent with de- ules (Figure 8), void and fracture filling (Figures 9, 10),
position on a shallow, gently dipping ramp under and, occasionally, as thick intervals of coalesced nod-
humid conditions. ules (Figure 11).
The obducted complex was strongly emergent at Oil production from the upper Maastrichtian res-
the start of the Maastrichtian, the Aruma Sea trans- ervoir production occurs in the dolomitized, mud-
gressed onto its western flank, and the ophiolite dominated peloidal packstones, skeletal-peloidal
subsided through that stage, forming a relatively grain-dominated packstones, and rudist rudstones.
open shelf in the north, becoming broader and more The oil production from the lower Maastrichtian in-
restricted farther south (Skelton et al., 1990). The con- terval occurs in dolomitized peloidal grainstones and
tinued narrowing of the Neotethys could also pos- packstones.
sibly explain the restriction in the marine circu- The field has produced less than 1% of the origi-
lation that resulted in the lagoonal sedimentation nal oil in place (OOIP) based on an estimate of nearly
during sea level lowstands in the upper Maastrichtian 1.5 billion bbl obtained from the reservoir model
interval. Overall, the abundant evidence of subaer- completed as part of this study. Production is highly
ial exposure, thin beds, dolomitization, and anhy- variable; initial rates range between several hundred
drite suggests that deposition during the first Maas- barrels of oil per day (BOPD) to more than 3000 BOPD.
trichtian occurred on a very gently dipping, shallow, Some wells are exceptional producers; one well has
arid, and restricted ramp setting that transitioned produced more than 5 million bbl of oil and some
between normal-marine conditions to restricted la- water out very quickly. Although representing a very
goonal environments. significant oil resource, the Maastrichtian reservoir
provides several challenges that this study addresses,
including (1) the presence of multiple barriers and
RESERVOIR DESCRIPTION baffles that compartmentalize the reservoir areally
and vertically; (2) the areal variation of oil gravity
The Maastrichtian reservoir is composed mainly of (13 –218 API); (3) the absence of a well-defined oil-
dolomite and limestone with thin beds of organic- water contact; (4) a water zone in the uppermost part
rich shale. Individual shale layers ranges from less of the reservoir interval; and (5) the distribution of
than 2 cm to 1 m (0.78 in. to 3.3 ft) or more. The possible fractures and faults that may provide an ex-
second Maastrichtian shale separates the Maastricht- planation for why some wells water out very quickly
ian into an upper and lower portion with signifi- whereas other wells produce at very low water cuts
cantly different depositional characteristics. The for many years.
type log given in Figure 7 shows the vertical extent Prior published reports on the Maastrichtian res-
of the first and second Maastrichtian intervals. The ervoir at Wafra field are very limited. Brief mention
upper part of the Maastrichtian is also known as the is made by Nelson (1968) and Alsharhan and Nairn
first Maastrichtian reservoir. The lower portion is com- (1997). Danielli (1988) provides some information
monly referred to as the second Maastrichtian res- about the shallow reservoirs at Wafra field, includ-
ervoir. This chapter will use upper and lower Maas- ing the Maastrichtian.
trichtian, instead of first and second Maastrichtian
because second Maastrichtian shale does not extend
across the entire modeled area. SEQUENCE-STRATIGRAPHIC FRAMEWORK
The lower Maastrichtian is composed largely of OF THE MAASTRICHTIAN
dolomite with local limestone and shale, whereas
the upper Maastrichtian is composed largely of lime- The Maastrichtian reservoir represents a shallow-
stone with locally abundant dolomite and shale. The ing-upward, lower order sequence culminating in a
shale in the upper Maastrichtian is organic rich, with regionally correlative subaerial exposure surface that
total organic content as much as nearly 50% of the marks the top of the Maastrichtian and is interpreted
rock volume and it is largely nonmarine. The shale in to be equivalent to the top of tectonostratigraphic
the lower Maastrichtian contains very little organic megasequence (TMS) AP9 of Sharland et al. (2001).
matter, less than 1% of the rock volume, and is largely Abdel-Kireem et al. (1994) described an unconformi-
marine in origin, which is supported by the palyno- ty at base of the Paleocene and the top of the late
254 / Dull et al.

FIGURE 7. Type log cored well 3 for the Maastrichtian showing the reservoir interval, high-frequency sequence-
stratigraphic picks, and lithologic core description. The well-log curves are track 1 (CGR = corrected gamma ray;
SGR = spectral gamma ray; VCLGR = clay volume from gamma ray); track 2 (core porosity; PHIE = effective porosity);
track 3 (PHIE; BVO = bulk volume oil; BVW = bulk volume water); and track 4 (facies). Depth in feet.

Maastrichtian in northeastern Iraq. The TMS AP9 face is overlain by a marine incursion that is the base
marks the end of the Cretaceous ophiolite obduction of the Paleocene-age Umm Er Radhuma Formation.
on the northern Arabian plate, resulting in widespread A stratigraphic cross section through the five cored
regression (Beydoun, 1991, 1993). This exposure sur- wells showing the identified evidence for subaerial
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 255

Table 1. Interpreted environments of deposi- dation space, making it uncertain if a facies change
tion from the shale palynology study of the was caused by eustatic sea level changes or topogra-
Maastrichtian.
phy. The sequence-stratigraphic modeling for the
Well Depth Interpretation HFS Maastrichtian was based on 413 m (1356 ft) of core
(ft) from five wells and well logs from 123 wells. The core
description and the correlation of the five Maastricht-
5 2567 Marine M00
ian cores have revealed what is interpreted as nine
5 2619 No palynomorphs M00
HFSs of an overall regressive sequence. Figure 7 is the
4 2743 Possible marine M10
type log for the Maastrichtian showing nine of the
3 2784 Marsh and embayment M20
HFSs and the lithofacies from the core description.
5 2780 No palynomorphs M20
With only five cored wells, it was essential to use the
4 2796 Marine M20
well logs from the other wells to construct the 3-D
3 2854 Restricted lagoon and fresh water M30
stratigraphic framework for the full Maastrichtian
4 2851 Possible marine M40
3 2932 Distal lagoon, marine influence M50
reservoir interval. As documented below, a synthetic
5 2851 Marine M50 lithofacies curve was generated from the petrophysi-
4 2885 Probable marine M50 cal data to assist in the fieldwide stratigraphic inter-
pretation and correlation.
The HFS for the Maastrichtian was identified using
exposure as hardgrounds and karst, interpreted envi- established bounding surface criteria that indicate a
ronments of deposition from the palynology, and in- base-level-fall to base-level-rise turnaround. The bound-
terpreted HFSs that were used to define the sequence- ing surfaces of the HFS were identified most com-
stratigraphic framework of the Maastrichtian is given monly by subaerial exposure surfaces or unconfor-
in Figure 12. The Maastrichtian is an overall regressive mities expressed as hardgrounds that show burrow
sequence composed of a transgressive systems tract and root casts filled with anhydrite commonly capped
(TST) and a highstand systems tract (HST). Figures 8–11 by organic terrestrial lagoonal shale (Figure 32). Less
and 13–31 show the core slab photos for the HFS M10 commonly, the bounding surfaces were identified by
to M80 interval for cored well 3. The location of the karst breccia with solution-enhanced open fractures
cored wells and cross section location are shown on and void-filling anhydrite occasionally associated with
the map of the field in Figure 3. subaerial hardgrounds (Figure 33). Uncommonly, the
The establishment of an appropriately detailed HFS were identified by (1) tidal flat caps of dolomud-
sequence-stratigraphic framework provides the foun- stone and dolowackestone, with pseudomorphs of
dation for the reservoir models as the stratal relation- anhydrite after gypsum (Figure 34), (2) lithofacies tract
ships, and the genetically related sequences largely offset from high-energy peloidal dolograinstones to
control the spatial distribution of the facies and pe- subtidal-marine dolowackestones and mud-rich dolo-
trophysical properties of the reservoir (Kerans and packstones (Figure 35), or (3) rudist and skeletal dolo-
Tinker, 1997; Deutsch, 2002). The basic framework rudstone with tidal-flat caps (Figure 36).
of the Maastrichtian is the HFS. The HFS, as defined
by Kerans and Tinker (1997), is bounded by a base- Fieldwide Correlation
level-fall to base-level-rise turnaround and is com- The fieldwide sequence-stratigraphic correlation
posed of genetically related cycles and cycle sets. A of the 10 HFSs of the Maastrichtian identified in the
cycle is defined as the smallest set of genetically cored wells was extended to all the wells based on a
related lithofacies representing a single base-level rise synthetic facies curve (FACIESG) and clay volume from
and fall. This is comparable to the parasequence of gamma-ray curve (VCLGR). Both curves were used to
Van Wagoner et al. (1988, 1990). The cycle set is a assist in the fieldwide correlation.
bundle of cycles that show a consistent trend of ag- The correlation between well-log response and res-
gradation or progradation (Kerans and Tinker, 1997). ervoir quality was essential to predicting facies on
The HFS was chosen as the basic building block for noncored wells with well logs. A review of the core
correlation in the Maastrichtian because it could be material, measured porosity, and measured perme-
correlated with a great deal of confidence. Individual ability indicated seven lithofacies or rock types. In
cycle correlation in the Maastrichtian is hampered by the course of core review, it became necessary to dis-
large distances between wells, abundance of expo- tinguish the mud-lean dolopackstones from the mud-
sure surfaces, and thin beds related to low accommo- rich dolopackstones and the dolorudstone from the
256 / Dull et al.

FIGURE 8. Core slab photos 2905 – 2920 ft (855 – 890 m)


FIGURE 10. Core slab photos 2938 – 2956 ft (895 – 900 m)
for cored well 3.
for cored well 3.

FIGURE 9. Core slab photos 2887 – 2905 ft (879 – 855 m) FIGURE 11. Core slab photos 2869 – 2887 ft (874 – 879 m)
for cored well 3. for cored well 3.
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 257

FIGURE 12. Stratigraphic cross section through the five cored wells showing the identified evidence for subaerial
exposure, interpreted environments of deposition from the palynology, and interpreted HFSs that were used to define
the sequence-stratigraphic framework of the Maastrichtian. The Maastrichtian is an overall regressive sequence
composed of a transgressive system tract and a highstand system tract. Depth in feet.

dolofloatstones. In addition, the cemented hardground (1997). The FACIESG curve was calculated using the
or dolostone and dolomudstone were combined be- core lithofacies data from cored wells 2– 5.
cause they were indistinguishable on the well logs Seven distinct carbonate lithofacies and shale were
and poor reservoir rocks. Figure 37 is a graph of po- identified from the core description that could be dis-
rosity and permeability of the dolopackstones show- tinguished using well-log curves. The pore system is
ing the difference between the mud-rich and mud-lean all secondary in nature and is primarily intercrystal-
dolopackstones. The dolorudstones and dolofloatstones line because of the fabric-destructive dolomitization.
shown in Figure 38 also show a significant difference The original depositional textures appear to have in-
in porosity and permeability. Figure 39 is a graph of fluenced the development of the intercrystalline
core porosity and permeability for the seven litho- pores. The mud-dominated lithofacies (lime mud-
facies indicating the grain-rich facies; dolograinstone, stone, dolomudstone, dolowackestone, and dolo-
mud-lean dolopackstone, dolorudstone, and mud-rich floatstone) (Figure 40a–c) are of poor reservoir quality
dolopackstone are the primary reservoir facies (litho- because of the predominance of original fine lime
facies 4–7). Lithofacies 1–3, dolomudstone, dolowacke- micrite matrix or dolomitized muddy matrix resulting
stone, and dolorudstone are of low reservoir quality in connected pore system-dominated very fine inter-
and potential barriers or baffles to flow. crystalline micropores. Figure 40d–g are thin-section
A method of predicting lithofacies in noncored wells photomicrographs of the more grain-dominated litho-
was developed based on the multiple cross-product, facies (mud-rich dolopacktstone, dolorudstone, mud-
nonlinear method as documented by Lawrence et al. lean dolopackstone, and dolograinstones). These rock
258 / Dull et al.

FIGURE 15. Core slab photos 2753 – 2771 ft (839 – 844 m)


FIGURE 13. Core slab photos 2722 – 2740 ft (829 – 835 m) for cored well 3.
for cored well 3.

FIGURE 14. Core slab photos 2740 – 2753 ft (835 – 839 m) FIGURE 16. Core slab photos 2771 – 2784 ft (844 – 848 m)
for cored well 3. for cored well 3.
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 259

FIGURE 17. Core slab photos 2784 – 2802 ft (848 – 854 m) FIGURE 19. Core slab photos 2815 – 2833 ft (858 – 863 m)
for cored well 3. for cored well 3.

FIGURE 18. Core slab photos 2802 – 2815 ft (854 – 858 m) FIGURE 20. Core slab photos 2833 – 2851 ft (863 – 868 m)
for cored well 3. for cored well 3.
260 / Dull et al.

FIGURE 21. Core slab photos 2851 – 2869 ft (868 – 874 m) FIGURE 23. Core slab photos 2956 – 2974 ft (900 – 906 m)
for cored well 3. for cored well 3.

FIGURE 22. Core slab photos 2920 – 2938 ft (890 – 895 m) FIGURE 24. Core slab photos 2974 – 2989 ft (906 – 911 m)
for cored well 3. for cored well 3.
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 261

FIGURE 25. Core slab photos 2989 – 3007 ft (911 – 916 m) FIGURE 27. Core slab photos 3025 – 3040 ft (922 – 926 m)
for cored well 3. for cored well 3.

FIGURE 26. Core slab photos 3007 – 3025 ft (916 – 922 m) FIGURE 28. Core slab photos 3040 – 3058 ft (926 – 932 m)
for cored well 3. for cored well 3.
262 / Dull et al.

FIGURE 29. Core slab photos 3058 – 3076 ft (932 – 937 m)


for cored well 3.

FIGURE 30. Core slab photos 3076 – 3094 ft (937 – 943 m) FIGURE 31. Core slab photos 3094 – 3202 ft (943 – 975 m)
for cored well 3. for cored well 3.
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 263

FIGURE 32. Core slab photo from cored well 3 (2857 ft;
870 m) showing uneven and eroded hardground with
burrows and root casts filled with anhydrite overlain by
organic-rich lagoonal shale.

fabrics are of better reservoir quality because of the


predominance of more coarsely crystalline dolomite,
producing the larger, better connected poor system and
lack of microporosity associated with a muddy matrix.
In addition, the permeability of the grain-dominated
lithofacies is higher because of common fine interpar-
ticle pores and associated late-stage etching. Table 2
shows the mean porosity and permeability of the car-
bonate lithofacies ranked by reservoir quality.
Lithofacies data from well 1 were not used because
of difficulties ensuring a reasonable depth match
given the very limited amount of core data. The well-
log curves PHIE (effective porosity), Sw (water satura-
tion), VCLGR (clay volume from corrected gamma- FIGURE 33. Subaerial hardground overlying karst breccia
ray curve), CGR (corrected gamma ray for uranium), with solution-enhanced open fractures and void-filling
and SGR (spectral gamma ray) were used to calculate anhydrite. Cored well 5 at depth of 2765 ft (842 m).
264 / Dull et al.

The FACIESG curve was critical to the fieldwide


correlation for the Maastrichtian because it provided
lithofacies curves to assist in identifying the cycle hi-
erarchy. The cycle hierarchy and the HFSs identified
in the core description, coupled with the FACIESG
curves on the noncored wells, were used for field-
wide correlation and to construct the 3-D sequence-
stratigraphic framework.
The VCLGR curve generated from the corrected
CGR was used in the sequence-stratigraphic correla-
tion, particularly in the upper Maastrichtian HFS M40
to M00. The core description revealed that many of the
hardground surfaces were capped with organic-rich
shale (Figure 41). The hardgrounds are consistently
identified by their low porosity and permeability and
appear as mudstones on the FACIESG curves. The shale
can easily be distinguished with the VCLGR curve as
shown on the type log given previously in Figure 7.

FIGURE 34. Dolomudstone with anhydrite after gypsum


crystals (cored well 4 at 2854 ft [869 m]).

the FACIESG curve. The correlation coefficient varied


between 0.41 and 0.79. A relatively low correlation
between actual facies and predicted facies (FACIESG)
FIGURE 35. Core slab photographs from cored well 3
was found for wells 2 and 4 and was attributed to
showing cycle top of HFS M70 of high-energy peloidal
having less core data as compared to the other two dolograinstones capped with subtidal-marine low-energy
wells and are thus likely to be less precisely depth mud-rich peloidal dolopackstones and dolowackestones
matched than the other cored wells. (3007 – 3025 ft; 916 – 922 m).
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 265

back in the middle Maastrichtian. The maximum


flooding surface (MFS) in HFS M50 at Wafra field
is interpreted to be equivalent to the MFS K180 of
Sharland et al. (2001). The maximum flooding event
resulted in the deposition of the second Maastrich-
tian shale across most but not all of the Wafra field
area.
The TST is composed of five HFS. The HFS M90 to
M60 represents overall shallowing-upward sequences
composed of cycles from subtidal, peloidal, and bio-
clastic dolowackestone to dolopackstone or dolo-
grainstone and, less commonly, subtidal mudstone
to peloidal dolograinstone cycles. The uppermost
HFS records the maximum transgression and is com-
posed of marine shale, argillaceous mudstone, and
peloidal wackestone. The existence of the peloidal
grainstones in the lower Maastrichtian in the M70
interval in cored well 3 seems to indicate a less-
FIGURE 36. Core slab photographs (2878 – 2905 ft; 877 – restricted and higher energy environment of deposi-
885 m) from cored well 5 showing dolofloatstone capped tion, with cycles that are generally thicker, indicat-
with intertidal dolomudstone, karst, and hardground ing a greater accommodation space. The deepening
defining the top of HFS M40.
HFS M50 includes an MFS that is marked by the
deposition of the second Maastrichtian shale. The
second Maastrichtian was probably deposited in more
Lower Maastrichtian HFS M90 to M50: open-marine conditions as evidenced by the lack of
Transgressive Systems Tract organic matter as compared to the upper Maastrich-
The deposition in the lower Maastrichtian reser- tian and as suggested from the palynological inter-
voir from the HFS M90 to HFS M50 interval repre- pretation from HFS M50 in wells 3, 4, and 5 as given
sents a marine transgression with a maximum flood in Table 1.

FIGURE 37. Graph of porosity


and permeability for lithofa-
cies showing the difference
on reservoir quality between
mud-rich (4) and mud-lean
dolopackstone (6).
266 / Dull et al.

FIGURE 38. Graph of porosity


and permeability for lithofacies
showing the difference on reser-
voir quality between dolofloat-
stone (3) and dolorudstone (5).

Upper Maastrichtian HFS M40 to M00: combination of nondeposition or low accumulation


Highstand Systems Tract rates, and condensation.’’ Hardgrounds are surfaces
The upper Maastrichtian HST is composed of five that are cemented before the next sedimentation
shallowing-upward HFS sequences commonly capped event. The core data show that the HST is typically
with hardgrounds and organic-rich shale. Flugel (2004, characterized by thin cycles, indicating less accom-
p. 206) describes hardgrounds as subtidal cementa- modation space, and also exhibits a greater variabil-
tion at the water-sediment interface as ‘‘related to a ity in facies with relative sea level changes. The lack

FIGURE 39. Graph of porosity


and permeability for the seven
lithofacies showing that the
more grain-dominated rocks
have higher permeability. The
best reservoir rock is the grain-
dominated lithofacies shown
in red (7 = dolograinstone; 6 =
mud-lean dolopackstone) and
green (5 = dolorudstone; 4 =
mud-rich dolopackstone). The
poor reservoir rock and poten-
tial baffles or barriers to flow
are shown in blue (3 = dolo-
floatstone; 2 = dolowacke-
stone; 1 = dolomudstone).
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 267

of accommodation space and resulting thin cycles in ity and on the well logs as a sharp change from low
the upper Maastrichtian is reflected in the number to high water saturation (Figure 44). The HFS M10
and types of hardgrounds. surface separates the oil-productive interval from
Two types of hardgrounds dominate the upper known water production from the HFS M00. A ma-
Maastrichtian: a subtidal firmground to incipient hard- rine transgression across the top of the Maastrichtian
ground and an intertidal to supratidal hardground formation overlies the unconformity surface M00
as described by Hillgartner (1998). Hillgartner (1998), that marks the top of the Maastrichtian. HFS M00 in
p. 1095, described the subtidal firmground as ‘‘low ac- cored well 5 shows a karsted interval capped with an
cumulation rate in a lagoonal environment, which intertidal hardground that is overlain with marine
favors consolidation and incipient cementation of shale (Figure 45).
the sediment at the water-sediment interface.’’ The The palynological and x-ray diffraction data from
upper Maastrichtian firmgrounds show characteris- the shale for the HFS M50 to HFS M00 suggest that
tics similar to those described by Hillgartner (1998): the shale deposition transitioned from a quiet la-
highly bioturbated with overlying sediment infilling goonal setting to marine settings that are interpreted
the burrows and an irregular surface probably caused as localized marine incursions that resulted from eu-
by weak erosion and/or differential cementation. The static sea level rise.
overlying sediment infilling the burrows in this case
is organic-rich lagoonal shale as shown in Figure 41.
The intertidal to supratidal hardgrounds tend to be GEOSTATISTICAL RESERVOIR MODEL
flat with a great deal more of associated anhydrite-
infilling burrows, fenestral pores, and evidence of A geostatistical model for the Maastrichtian res-
exposure such as tension gashes. The intertidal to sup- ervoir was generated to assess OOIP and reservoir
ratidal hardgrounds are occasionally found above development potential. Figure 3 shows the area mod-
microkarsts. eled and the distribution of wells with appropri-
The recognition of hardgrounds of low porosity ate well-log data (e.g., the synthetic facies curve
and permeability in the upper Maastrichtian is im- FACIESG described above along with porosity and Sw
portant not only for cycle correlation but for po- curves). The modeled area covers 144 km2 (55.6 mi2).
tential impact on reservoir compartmentalization. The modeled interval includes the entire Maastricht-
The incipient subtidal and intertidal hardgrounds as ian reservoir and the top of the Maastrichtian (HM00)
described by Hillgartner (1998) were found to have to the top of Hartha. The overall model grid dimen-
lateral extents from 0.1 to 1 km (0.06 to 0.6 mi). Some sions are 106  136  622 cells (8.97 million cells
of the hardground-capped cycles cannot be correlat- total). The areal cell size is 100  100 m (330  330 ft),
ed in the upper Maastrichtian, suggesting that the and the vertical layering is nominally 0.3 m (1 ft)
hardgrounds may not act as continuous barriers to between the HM00 and HM90 markers. The nominal
flow throughout the entire reservoir. vertical layering is 0.6 m (2 ft) from the HM90 marker
Floatstones and rudstones dominate the middle to the base of the model.
part of the HST from the M50 to the M20 marker. The geostatistical property model was generated
The floatstones-rudstones are composed of broken using an industry standard workflow (Deutsch, 2002;
to intact rudists (Figure 42), bivalves, and gastropods. Kelkar and Perez, 2002; Goovaerts, 1997). Porosity was
The floatstones commonly contain the large benthic distributed by sequential Gaussian simulation (sGs)
foraminifer, Loftusia (Figure 43). The floatstone- constrained by sequence-stratigraphic layers. Sw was
rudstone less commonly contains corals, stromato- distributed using a colocated cokriging with sGs al-
poroids, and bryozoans. The deposition of the rudist gorithm also constrained by sequence-stratigraphic
and skeletal debris is interpreted to be biostromal, layers. Tables 3 and 4 summarize the input data and
with no evidence of mounding. An abrupt facies semivariograms used to build the reservoir property
change is indicated from the HFS M20 to M10 from model. Although not used to constrain porosity or
restricted lagoon to more open-marine deposition of Sw distribution, several additional reservoir proper-
peloidal and bioclastic packstones. ties, such as FACIESG and Pe (photoelectric effect),
A subareal exposure surface occurs at the top of were also distributed using sGs constrained by a strat-
HFS M10 (Figure 12) that is porosity destructive and igraphic layer to assist in understanding the develop-
provides the seal on the oil-productive interval. This ment potential of the reservoir, particularly the se-
is shown on the cores as low porosity and permeabil- quences below M20.
268 / Dull et al.

FIGURE 40. Thin-section photomicrographs illustrating the seven dolomite lithofacies that compose the Maastrichtian
arranged by reservoir quality. (a) Argillaceous lime mudstone with a few bioclasts and composed primarily of silt-size
micrite and micrite that has been neomorphosed to microspar then etched, creating the microporosity. Organic matter
is present as fine filamentous algal matter. Porosity occurs as micropores and reduced molds. The permeability is poor
because the connected pore system is dominated by the very fine intercrystalline pores (Phi = 17.9%, k = 0.16 md).
(b) Bioclastic dolowackestone or grain-poor argillaceous and organic dolomite composed of fine-grained dolomite
matrix cut by coarse-grained dolomite laminae or burrows. The matrix is composed of organic matter and detrital clays
but the dolomite crystals form the framework. Organic matter is common as pellets and flattened tabular grains.
Porosity occurs in micropores, and permeability is poor because of the replacement of the fine carbonate matrix with
tightly interlocking dolomite crystals (Phi = 4.5%, k < 0.10 md). (c) Bioclastic dolofloatstone with the most common grains
are benthonic foraminifera (Rotalids) and Loftusia. Other bioclasts include ostracods, bivalve debris, gastropods, and
echinoderm plates. Organic matter is found as reddish brown pellets and possibly bitumen. Uncommon angular fragments
of bones are also present. The original matrix is not preserved and has been replaced or etched. Dolomite is the major
mineral. Porosity is a combination of molds and secondary interparticle pores. Permeability is low as result of the
variations in porosity and the frequency of organic laminae (Phi = 14.4%, k < 2.21 md). (d) Mud-rich dolopackstone in
which the fabric-destructive dolomitization has largely destroyed the original depositional fabric. No evidence of
carbonate mud matrix is preserved, and the organic matter is preserved as pellets. The finer dolomite crystal laminae
suggest that the original fabric contained some lime mud. Porosity occurs as secondary intercrystalline pores with
some local intracrystalline pores. Permeability is low because of the generally interlocking mosaic of the dolomite
crystals. Larger isolated pores exist, but are connected by finer pore throats (Phi = 17.6%, k < 5.00 md).
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 269

FIGURE 40. (cont.). (e) Dolorudstone is a mixture of fine-to medium-grained dolomite that has partly replaced grains
and relatively coarse dolomite, destroying much of the original rock fabric. The fabric suggests an original bioclastic
grainstone that was overpacked. No matrix appears to have been present. The framework grains are commonly rudist
debris, bivalve debris fragments of Orbitoides with possible stromatoproids and/or Loftusia. The porosity is all secondary
as intercrystalline pores. Calcite cement has reduced some of the finer and more isolated pores. Permeability is higher
because of the occurrence of fine interparticle pores (Phi = 19.8%, k < 80.0 md). (f ) Mud-lean peloidal dolopackstone
(Phi = 15.3%, k < 9.5 md) that has undergone fabric-destructive dolomitization and dissolution of the original rock. Large
tabular molds may have been molluskan debris. Only ghosts of what appear to be medium-grain-size peloids, ooids,
or other rounded grains are found. Organic matter is present as fibrous or brown amorphous pellets. Porosity is
composed exclusively of secondary pores, primarily intercrystalline pores, although reduced molds are also found.
Some intracrystalline pores exist but are uncommon and ineffective. Permeability is good probably because of
the uneven distribution of pores and appears to be associated with carbonaceous laminae, suggesting that the original
fabric may have been a poorly sorted grainstone. (g) Peloidal dolograinstone with a mixture of fine- to medium-
crystalline dolomite that has partly replaced grains and coarsely crystalline dolomite, which has largely destroyed the
rock fabric. Extensive pores created by leaching of undolomitized limestone have been extensively cemented
by calcite. No matrix appears to have been present. Partly mimetic-replaced bioclasts suggest small benthonic fora-
minifera, molluskan debris, algae, and fragments of large foraminifera. Very uncommonly are grain outlines com-
plete and almost always rounded or tabular. Porosity occurs as secondary intercrystalline pores. Permeability is
high because of late-stage etching and an open dolomite framework (Phi = 19.9%, k = 1108 md).
270 / Dull et al.

Table 2. Average core porosity and permeability locations as well as select exploration targets in the
by lithofacies.
Maastrichtian reservoir’s deeper portions.
Lithofacies* Mean Permeability Mean Figure 48 shows the variation in Pe and, hence,
(md) Porosity limestone and dolomite throughout the reservoir.
Note that the dolomite is much more abundant in the
1 0.022 0.086 more productive, upper parts of the Maastrichtian.
2 1.445 0.080
3 20.550 0.114
4 11.477 0.134
MAASTRICHTIAN DEVELOPMENT POTENTIAL
5 81.037 0.182
6 114.817 0.200
The Maastrichtian reservoir contains more than
7 294.868 0.202
1 billion bbl of oil in place. The degree of compart-
*1 = dolomudstone; 2 = dolowackestone; 3 = dolofloatstone; mentalization of the Maastrichtian can be shown by
4 = mud-rich dolopackstone; 5 = dolorudstone; 6 = mud-lean
dolopackstone; 7 = dolograinstone.

As shown in Table 3 and Figures 46 and 47, a gen-


eral increase in porosity with depth is present, along
with an increase in Sw. Porosity is very irregularly
distributed in the upper parts of the reservoir, but is
generally best developed in the M10 and M20 inter-
vals. The Sw distribution shown in Figure 47 provides
the best depiction of the distribution of productive
and potentially productive parts of the Maastrichtian
reservoir at Wafra. The middle parts of the Maas-
trichtian show isolated regions of high porosity and
low Sw that may be near-term exploration targets.
Developing a proper sequence-stratigraphic frame-
work and using it to constrain 3-D reservoir models is
key if the 3-D models are to be used to prioritize infill

FIGURE 41. Core slab photo from cored well 2 of black FIGURE 42. Core slab photo of rudist that appears to be
coaly shale in the upper Maastrichtian deposited on an intact, center filled with blue anhydrite (cored well 5,
uneven subaerial hardground. (12816 ft [858 m]). 2722 ft [829 m]).
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 271

area of the model also shows significant HCPV de-


velopment in the M40 that trends northeast to
southwest along the paleohigh. The northeast trend
of prospective reservoir development appears to
continue outside the model area beyond well con-
trol. No Maastrichtian well completions exist in this
area, and the model suggests significant exploration
potential to the north of the model area along the
Maastrichtian paleohigh and along paleodeposition-
al strike. Note that if the reservoir property models
used to generate the HCPV distributions were not
constrained by stratigraphy, but instead by lithology,
the trend shown in Figure 50 probably would not
have been captured, much less be used, to assist devel-
opment and exploration planning. Additional stud-
ies, incorporating seismic attributes, are in progress
to further refine the 3-D reservoir property maps and
to enhance the usefulness of the models to assist res-
ervoir development.

SUMMARY

The methodology for developing the sequence-


stratigraphic framework for the Maastrichtian at
Wafra field included a workflow to integrate core
and well-log data enabling the construction of the
3-D sequence-stratigraphic framework. The work-
flow for the Maastrichtian consisted of

 using the five cores to define the depositional


environment and cycle hierarchy
 correlating the five cores to define the ten HFSs
 developing synthetic facies curves from the well-
log data that enabled the use of the 123 well logs
with FACIESG and VCLGR curves to construct
the 3-D stratigraphic framework

The sequence-stratigraphic framework is essential


FIGURE 43. Core slab photo of Loftusia dolowackestone to the geostatistical modeling and to reservoir simu-
(cored well 4, 2826 ft [861 m]).
lation because it controls the spatial distribution of
petrophysical properties and identifies the chrono-
stratigraphic surfaces that act as barriers and baffles
maps of hydrocarbon pore volume (HCPV) for the to flow that compartmentalize the reservoir.
HFS M10 and M40 shown in Figures 49 and 50. The The modeling of the Maastrichtian reservoir has
trend of the highest HCPV for HFS M10 is coinci- revealed
dent with present-day structural trends and repre-
sents a peloidal dolopackstone shoal complex of the  a large oil resource of more than 1 billion bbl of
late HST. The M10 is the most extensive and pro- oil in place
ductive interval of the Maastrichtian. Whereas, the  that entrapment of hydrocarbons is primarily
HFS M40 shows a northeast to southwest linear trend stratigraphic
that likely parallels paleotopography. The northeast  a highly compartmentalized reservoir
272 / Dull et al.

FIGURE 44. A stratigraphic cross section of the cored wells shows the varying water saturation observed in the reservoir
interval and the grain-dominated lithofacies that typically have low water saturation. The M00 at the top of the
Maastrichtian from production tests is known to be water productive. Besides the M00, the M90 interval is the only
Maastrichtian sequence that shows nearly 100% water saturation. Depth in feet.

 multiple oil-water transition zones and vertical hydrocarbon distribution of the Maas-
 no single definitive oil-water contact trichtian that will have a substantial impact on the
 a water zone above and below the oil reservoir reservoir development.
 additional development potential associated
with paleohighs and facies trends ACKNOWLEDGMENTS

The sequence-stratigraphic framework and geo- The authors thank the Ministry of Petroleum and
statistical modeling have quantified a very signifi- Mineral Resources, Kingdom of Saudi Arabia; Saudi
cant asset and provided great insight into the areal Arabian Texaco; and Chevron Energy Technology
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 273

Table 3. Summary of the well-log porosity and


water-saturation data used to construct the
reservoir property models listed by HFS.

Stratigraphic Average Average


Interval Porosity Sw

M00 0.166 0.795


M10 0.150 0.531
M20 0.137 0.635
M30 0.131 0.611
M40 0.130 0.670
M50 0.125 0.761
M60 0.127 0.607
M70 0.135 0.676
M80 0.165 0.779
M90 0.202 0.896

FIGURE 45. Core slab photo of top of HFS M00 in cored


well 5 (2567 – 2568 ft; 782.4 – 782.7 m) showing hard-
ground capped with shaly dolomudstone and a dark
organic-rich marine shale marking a marine incursion in
the late Maastrichtian. Note the large lithoclast above
the irregular hardground surface.
274 / Dull et al.

Table 4. Summary of the variograms used in the property distribution of the geostatistical models.
Interval Range x (m) Range y (m) Azimuth Range z (m) Nugget Sill Form

M00 2460 1350 N1258E 6.0 0 1 Exponential


M10 1660 1220 N608E 2.7 0 1 Exponential
M20 820 660 N458E 4.6 0 1 Exponential
M30 1830 1260 N258E 4.4 0 1 Exponential
M40 1600 1110 N258E 3.4 0 1 Exponential
M50 1100 710 N408E 2.6 0 1 Exponential
M60 2060 1140 N108E 4.3 0 1 Exponential
M70 1720 1380 N358E 5.5 0 1 Exponential
M80 1760 1020 N408E 6.6 0 1 Exponential
M90 1860 1000 N358E 9.8 0 1 Exponential

FIGURE 46. Perspective view of Wafra Maastrichtian reservoir model showing porosity distribution. Note that porosity
distribution is very discontinuous, even in the generally higher porosity, lowermost stratigraphic zones. Within the
uppermost stratigraphic zones, only the M10 and M20 intervals show consistent high-porosity distribution across
large areas of the field. Vertical exaggeration = 5. Map shows the structure on top of the Hartha Formation (base
of Maastrichtian).
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 275

FIGURE 47. Perspective view of Wafra Maastrichtian reservoir model showing Sw distribution. Note the general increase
in Sw with depth. Also note that some of the middle stratigraphic zones (e.g., M40) have locally low Sw regions. Vertical
exaggeration = 5. Map shows structure on the top of the Hartha Formation (base of Maastrichtian).

FIGURE 48. Perspective view of Wafra Maastrichtian reservoir model showing Pe (photoelectric effect) distribution. Note
the general increase in Pe with depth showing the irregular distribution of limestone ( yellow) and dolomite (blue).
Note, however, that in general, there is much more dolomite in the upper parts of the Maastrichtian. Vertical
exaggeration = 5. Map shows structure on the top of the Hartha Formation (base of Maastrichtian).
276 / Dull et al.

FIGURE 49. Hydrocarbon pore volume


(HCPV) map for the M10 to the M20
interval. The area shown in orange is
greater than 10 ft (3 m) of HCPV. The
wells that have been completed in the
Maastrichtian are also shown. Contour
interval is 2 ft (0.6 m).
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 277

FIGURE 50. Map of hydrocarbon pore


volume (HCPV) for M40. The greatest
of amount of HCPV is coincident with
the maximum dolograinstone thickness.
This observation strongly supports the
premise that the hydrocarbon trapping
is primarily stratigraphic. The contour
interval for M40 HCPV map is 1 ft (0.3 m).
278 / Dull et al.

Company for their permission to share the results dian Society of Petroleum Geologists Memoir 17, p. 1 –
of this study. We also thank the reviewers for their 47.
Goovaerts, P., 1997, Geostatistics for natural resources
insight and guidance to this study: Steven L. Bachtel,
estimation: New York, Oxford University Press, 483 p.
ConocoPhillips Company; Sherry Becker, ExxonMobil
Harris, P. M., S. H. Frost, G. A. Seiglie, and N. Schneider-
Upstream Research Company; and P. Mitch Harris, mann, 1984, Regional unconformities and deposition-
Chevron Energy Technology Company. al cycles, Cretaceous of the Arabian Peninsula, in J. S.
Schlee, ed., Interregional unconformities and hydro-
REFERENCES CITED carbon accumulation: AAPG Memoir 36, p. 67 – 80.
Hillgartner, H., 1998, Discontinuity surfaces on a shallow-
Abdel-Kireem, M. R., A. M. Samir, and H. P. Luterbacher, marine carbonate platform (Berriasian, Valanginian,
1994, Planktonic foraminifera from the Kolosh For- France and Switzerland): Journal of Sedimentary Re-
mation (Paleogene) of the Sulaimaniah – Dokan re- search, Section B: Stratigraphy and Global Studies,
gion, northeastern Iraq; Tuebingen University Neues v. 68, no. 6, p. 1093 – 1108.
Jarhrbuch Geol.: Paleontology Monatsh, v. 9, p. 517 – Kelkar, M., and G. Perez, 2002, Applied geostatistics for
527. reservoir characterization: Society of Petroleum Engi-
Alsharhan, A. S., and S. J. Y. Nasir, 1996, Sedimentological neers, 264 p.
and geochemical interpretation of a transgressive se- Kerans, C., and W. S. Tinker, 1997, Sequence stratigraphy
quence: The Late Cretaceous Qahlah formation in the and characterization of carbonate reservoirs: SEPM
western Oman Mountains: United Arab Emirates Uni- Short Course Notes 40, 130 p.
versity Sedimentary Geology, v. 101, no. 3 – 4, p. 227 – Lawrence, T. D., A. Kohar, B. Sukamto, and H. Pramono,
242. 1997, Sonic density well log data editing with pseudo-
Alsharhan, A. S., and C. G. St. C. Kendall, 1995, Facies var- curve generation — Indonesian examples using a mul-
iation, depositional setting and hydrocarbon poten- tiple crossproduct non-linear method: World Petro-
tial of the Upper Cretaceous rocks in the United Arab leum Congress, p. 163 – 167.
Emirates: Cretaceous Research, v. 16, p. 435 – 449. Morris, A., M. W. Anderson, A. H. F. Robertson, and K. Al-
Alsharhan, A. S., and A. E. M. Nairn, 1990, A review of the Riyami, 2002, Extreme tectonic rotations within an
Cretaceous formations in the Arabian Peninsula and eastern Mediterranean ophiolite (Baer-Bassit, Syria):
Gulf: Part III. Upper Cretaceous (Aruma Group) stratig- Earth and Planetary Science Letters, v. 202, no. 2,
raphy and paleontology: Journal of Petroleum Ge- p. 247 – 261.
ology, v. 13, p. 247 – 266. Nelson, P. H., 1968, Wafra field— Kuwait-Saudi Arabia neu-
Alsharhan, A. S., and A. E. M. Nairn, 1997, Sedimentary tral Zone. Second Regional Technical Symposium, So-
basins and petroleum geology of the Middle East: New ciety of Petroleum Engineers, Saudi Arabia Section,
York, Elsevier, p. 843. Dhahran, 16 p.
Beydoun, Z. R., 1991, Arabian plate hydrocarbon, geology Philip, J. M., J. Roger, D. Vaslet, F. Cecca, S. Gardin, and
and potential — A plate tectonic approach: AAPG Stud- A. M. S. Memesh, 2002, Sequence stratigraphy, bio-
ies in Geology 33, 77 p. stratigraphy and paleontology of the Maastrichtian –
Beydoun, Z. R., 1993, Evolution of the northeastern Paleocene Aruma Formation in outcrop in Saudi Arabia:
Arabian plate margin and shelf: Hydrocarbon habitat GeoArabia, v. 7, no. 4, p. 699–718.
and conceptual future potential: Revue de l’Institut Sayed, M. G. S. A., and M. A. Mersal, 1998, Surface geology
Français du Pétrole, v. 48, p. 311 – 345. of Jebel Rawdah, Oman Mountains: GeoArabia, v. 3,
Danielli, H. M. C. D., 1988, The Eocene reservoirs of Wafra no. 3, p. 401 – 414.
field, in A. J. Lomando and P. M. Harris, eds., Giant Schreurs, G., and A. Immenhauser, 1999, West-northwest
oil and gas fields — A core workshop: SEPM Core directed obduction of the Batain group on the east-
Workshop 12, p. 119 – 154. ern Oman continental margin at the Cretaceous –
Deutsch, C. V., 2002, Geostatistical reservoir modeling: Tertiary boundary: Tectonics, v. 18, no. 1, p. 148 –
New York, Oxford University Press, 376 p. 160.
El-Nakhal, H. A., and Z. R. El-Naggar, 1994, Review of the Sharland, P. R., R. Archer, D. M. Casey, S. H. Hall, A. P.
biostratigraphy of the Aruma Group (Upper Cretaceous) Heward, A. D. Horbury, and M. D. Simmons, 2001,
in the Arabian Peninsula and surrounding regions: Arabian plate sequence stratigraphy: GeoArabia Spe-
Cretaceous Research, v. 15, p. 401 – 416. cial Publication 2, 371 p.
Flugel, E., 2004, Microfacies of carbonate rocks: Analysis, Skelton, P. W., S. C. Nolan, and R. W. Scott, 1990, The
interpretation and application: Berlin, Spring-Verlag, Maastrichtian transgression onto the northwestern
p. 976. flank of the Proto-Oman Mountains: Sequences of
Golonka, J., M. I. Ross, and C. R. Scotese, 1994, Phan- rudist-bearing beach to open shelf facies, in A. H. F.
erozoic paleogeographic and paleoclimatic modeling Robertson, M. P. Searle, and A. C. Ries, eds., The ge-
maps, in A. F. Embry, B. Beauchamp, and D. J. Glass, ology and tectonics of the Oman region: Geological
eds., Pangea global environments and resources: Cana- Society (London) Special Publication 49, p. 521 – 547.
The Sequence Stratigraphy of the Maastrichtian Reservoir at Wafra Field / 279

Smith, A. B., N. J. Morris, A. S. Gale, and W. J. Kennedy, Hastings, H. Posamentier, J. C. Van Wagoner, C. A. Ross,
1995, Late Cretaceous carbonate platform faunas of and C. G. St. C. Kendall, eds., Sea-level change: An in-
the United Arab Emirates – Oman border region: Ox- tegrated approach: SEPM Special Publication 42, p. 39 –
ford University Bulletin of the Natural History Museum 46.
(Geology Series), v. 51, no. 2, p. 91–119. Van Wagoner, J. C., R. M. Mitchum, K. M. Campion, and
Van Wagoner, J. C., H. W. Posamentier, R. M. Mitchum, V. D. Rahmanian, 1990, Siliclastic sequence stratigra-
P. R. Vail, J. F. Sarg, T. S. Loutit, and J. Hardenbol, phy in well logs, cores, and outcrops: Concepts for
1988, An overview of the fundamentals of sequence high-resolution correlation of time and facies: AAPG
stratigraphy and key definitions, in C. K. Wilgus, B. J. Methods in Exploration Series 7, 55 p.
Porter, M. L., A. R. G. Sprague, M. D. Sullivan, D. C. Jennette, R. T. Beaubouef,
T. R. Garfield, C. Rossen, D. K. Sickafoose, G. N. Jensen, S. J. Friedmann,

8
and D. C. Mohrig, 2006, Stratigraphic organization and predictability of
mixed coarse- and fine-grained lithofacies successions in a lower Miocene
deep-water slope-channel system, Angola Block 15, in P. M. Harris and L. J.
Weber, eds., Giant hydrocarbon reservoirs of the world: From Rocks to
reservoir characterization and modeling: AAPG Memoir 88/SEPM Special
Publication, p. 281 – 305.

Stratigraphic Organization and


Predictability of Mixed Coarse- and
Fine-grained Lithofacies Successions
in a Lower Miocene Deep-water
Slope-channel System,
Angola Block 15
M. L. Porter C. Rossen
ExxonMobil Production Company, Houston, ExxonMobil Development Company,
Texas, U.S.A. Houston, Texas, U.S.A.

A. R. G. Sprague D. K. Sickafoose
ExxonMobil Upstream Research Company, ExxonMobil Exploration Company,
Houston, Texas, U.S.A. Houston, Texas, U.S.A.

M. D. Sullivan G. N. Jensen
Chevron Energy Technology Company, ExxonMobil Production Company,
Houston, Texas, U.S.A. Houston, Texas, U.S.A.

D. C. Jennette S. J. Friedmann
Apache Oil, Houston, Texas, U.S.A. Lawrence Livermore Laboratory, Livermore,
California, U.S.A.
R. T. Beaubouef
ExxonMobil Exploration Company, D. C. Mohrig
Houston, Texas, U.S.A. Department of Earth, Atmospheric
and Planetary Sciences, Massachusetts
T. R. Garfield Institute of Technology, Boston,
ExxonMobil Production Company, Massachusetts, U.S.A.
Houston, Texas, U.S.A.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215880M883273

281
282 / Porter et al.

ABSTRACT

R
egional seismic mapping identified lower and middle Miocene slope
channels as significant exploration targets for Angola Block 15. Seventeen
exploration wells, followed by four appraisal wells, established these
slope-channel complexes as a world-class development opportunity. ExxonMo-
bil’s current development activity targets stacked turbidite-dominated reservoirs
in long-reach, high-angle wellbores tied back to tension leg platform (TLP) and
close-moored floating production, storage, and offshore loading facilities.
One of the major development targets in Block 15 is the Burdigalian-aged
(Bur1) slope-channel reservoirs. The Bur1 slope-channel system was one of many
lower Miocene sediment fairways that provided a mechanism for the delivery of
coarse-grained turbidites and mixed muddy and sandy debrites into the Lower
Congo basin. This slope-channel system traverses across the block in an east–west
direction and can be continuously mapped on adjacent seismic data sets across a
30–40-km (18–25-mi) reach. Three conventional cores and 28 well penetrations
calibrate updip to downdip changes in lithofacies type and channel architecture.
Map patterns of the Bur1 slope-channel system show distinctive changes in
sinuosity, channel confinement, and degree of amalgamation broadly related to
concurrent growth of salt-related structures. Channel-complex confinement is
more pronounced, and vertical amalgamation is better developed in segments
that cross structural highs. The Bur1 channel system shows weaker lateral
amalgamation, greater sinuosity, and less erosional confinement in structural
lows.
The episodic fill of the Bur1 slope-channel system can be better understood
by a hierarchical arrangement of unconformity-bounded stratal units. Within
these unconformity-bounded channel sets, nested channels form composite
channel complexes that show distinctive trends in lithofacies type and vertical
facies succession. Compared to other offshore Angola slope-channel systems,
this Bur1 system is noteworthy because of the relatively coarse granule to cobble
grain sizes encountered. Well logs and high-resolution seismic data calibrated to
conventional cores show that the lower parts of the channel complexes are
dominated by sandy-muddy debrites, slumps, and injected sandstones. These
facies are typically overlain by coarse-grained, gravelly, and well-amalgamated
sandy turbidites. The overlying facies succession is more variable, but commonly
consists of interbedded sandy and muddy turbidites, injected sandstones, and a
range of both muddy and sandy debrites.

BURDIGALIAN DEEP-WATER slope-channel systems (Bouchet et al., 2005), and the


CHANNEL SYSTEMS growth of ExxonMobil’s Kizomba development to
more than 550 thousand bbl/day of production in 3 yr
Introduction shows the prolific nature of deep-water depositional
The past 5 yr have seen a remarkable expansion of systems that exist in Block 15 (Figure 1).
offshore exploration and development activity in Seismic stratigraphy applied to conventional and
Angola. The highly productive hydrocarbon system high-resolution three-dimensional (3-D) data sets of-
in the Lower Congo basin, coupled with large struc- fered a compelling method to understanding the
tures, well-established sediment fairways, and excel- Tertiary history of the Angolan continental margin
lent seismic coverage has made this part of west (Beaubouef et al., 1998; Schollnberger and Vail, 1999;
Africa a world-class development opportunity. Total’s Gottschalk, 2002). Recent studies of the Zaire Fan
Girassol field development showed the huge poten- show a compelling view of successive fan development
tial that exists for deep-water production from Angola’s driven by erosion, sediment bypass, and punctuated
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 283

FIGURE 1. Location map of the Bur1 slope-channel system, offshore Angola. The Bur1 channel system forms a part of
the greater Kizomba development. Water depth ranges from 750 to more than 1300 m (2460 to more than 4265 ft)
in the study area. A 50-km (31-mi) transect of the Bur1 slope-channel system is possible because of the contiguous
3-D seismic coverage in the 400-km2 (154-mi2) study area. The distribution of oil and gas fields in the Lower Congo basin is
shown by green polygons. The inset photo shows one of the two close-moored tension leg platform (TLP) and floating
production, storage, and offloading vessel (FPSO) facilities producing hydrocarbons from the Bur1 system.

deposition in the updip submarine canyon and levee- system sets) commonly mapped with seismic and
channel system (Babonneau et al., 2002; Droz et al., sequence-stratigraphic methods (Sprague et al., 2002,
2003). The Pleistocene to Holocene history of the 2005).
Zaire Fan shows a linked shelf-slope-basin deposi- This chapter shows the seismic expression, map
tional system and provides an analog for interpreting trends, and reservoir elements of a major lower Mio-
the genetic packaging of complex deep-water depo- cene slope-channel system in offshore Angola. De-
sitional fan elements through sequence-stratigraphic velopment drilling reveals a complex, multistage
concepts (e.g., Garfield et al., 1998; Raposo and Sykes, lithofacies succession of coarse- and fine-grained
1998; Goulding et al., 2000; Temple and Broucke, turbidites and mixed debrites. These lithofacies form
2004). For reservoir-scale studies, placement of sub- predictable fining-upward channel fills in larger,
seismic-scale stratigraphy in sinuous deep-water chan- unconformity-bounded channel-complex sets that
nel fills is critical in developing a depositional model can be mapped across tens of kilometers. Regional map-
suitable for development drilling (e.g., Kolla et al., ping of the Bur1 system shows the position and extent
2001). A hierarchical approach to the physical stra- of reservoir unit changes along depositional dip, and
tigraphy of deep-water systems allows the smallest these variations can be attributed to sea-floor gradient
stratigraphic elements (beds) to be placed in in- changes associated with salt movement. A deposi-
creasing larger composite stratal packages (complex tional model is developed from the superposition of
284 / Porter et al.

FIGURE 2. General stratigraphy of the Lower Congo basin. The Bur1 slope-channel system is contained within the
Oligocene – Miocene Malembo formation. The Malembo and the underlying Landana Formation contain the bulk of
postrift, sag-phase sedimentation in the Lower Congo basin. The lower-right diagram shows a generalized depositional
model of a linked shelf-slope-basin depositional system (e.g., Garfield et al., 1998). The Bur1 reservoirs discussed in
this chapter represent only a small segment of a very large, deep-water depositional system. The study area is
approximately located at cross section AA0 on this profile and represents a deep-water confined channel complex
system in a midslope depositional setting.

unconformity-bounded packages that shows the rela- Brice et al., 1982; Emery and Uchupi, 1984; Burke,
tionships of turbidites and debrites in a nested stratal 1996) (Figure 2). A postrift, early sag phase filling of
hierarchy of deep-water channel complexes. the basin is represented by marine deposits of the
Cretaceous Pinda Formation. Source rocks are mostly
GEOLOGIC SETTING in the Upper Cretaceous Iabe Formation, which over-
lies the Pinda, and these shale-dominated rocks ac-
Deep-water slope-channel systems are one of the cumulated under low sedimentation rates. Fully ma-
important reservoir types in the hydrocarbon system rine conditions developed with continued sea-floor
of the Lower Congo basin. The Lower Congo basin is spreading of the south Atlantic, and this early Tertiary
a Cretaceous to Holocene sedimentary basin devel- drift phase is represented by marine shales, sandstones,
oped on the west African continental margin. The and carbonates of the Landana Formation. Middle to
early history of the basin is represented by a thick, late Tertiary basin fill is contained in the Oligocene
Aptian-aged salt succession that formed over rift to Miocene Malembo formation. The Malembo for-
basin deposits associated with the Early Cretaceous mation is a 6-km (3.7-mi)-thick, time-transgressive
breakup of Gondwana (Scrutton and Dingle, 1976; lithostratigraphic unit that contains strata deposited
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 285

FIGURE 3. Seismic traverse through the Bur1 slope-channel system. The upper panel shows the seismic expression
and structural configuration of the lower and middle Miocene along BB0. The Bur1 channel system is marked by dashed
horizons along a 55-km (34-mi) transect. Some of the well penetrations in the Bur1 are shown with corresponding
log responses. The lower panel is an interval amplitude extraction of the Bur1 slope-channel system. Note the confinement
of the system near or at structural highs and more weakly confined, sinuous map pattern developed in the intervening
synclines. Well penetrations, including cored wells, are shown by symbols on the seismic amplitude map.

in nearshore to deep-water slope and basinal envi- Regional tectonic studies and structural analysis
ronments (Temple and Broucke, 2004). of the Lower Congo basin have established a linked
At least three major episodes of clastic input across extensional-contractional structural style for the An-
the slope and into the basin are recognized in the golan continental margin (Larson and Ladd, 1973;
Oligocene and Miocene of the basin (Ardill et al., Rabinowitz and LaBreque, 1979; Duval et al., 1992;
2002). The Bur1 deep-water slope system discussed in Hartman et al., 1998; Gottschalk, 2002). Structures
this chapter forms a part of the second major influx of have developed as a result of salt-related deformation
coarse-grained clastic sediments in the lower Mio- associated with detachment on the Aptian salt. Updip
cene. Biostratigraphic analysis of cores and cuttings parts of the basin-margin slope are characterized by
in exploration wells shows that this slope-channel extensional faults and tilted fault blocks, and downdip
system is early Burdigalian in age. Nannofossil as- sections show a full suite of salt structures, compres-
semblages in bracketing hemipelagic slope mudstone sional anticlines, and thrust faults (e.g., Gottschalk
successions, tied to aged-dated European reference et al., 2004).
sections, show the Bur1 slope-channel system devel-
oped ca. 20 Ma. Internal age subdivision of the Bur1 THE BUR1 SLOPE-CHANNEL SYSTEM
channel system is problematic because of a lack of
age-diagnostic faunal assemblages and multiple epi- Sinuosity and Width
sodes of erosion, bypass, and deposition. Correlated A seismic amplitude extraction across the reservoir
ages of capping mudstone sequences that bracket the interval shows the large-scale map pattern of the Bur1
overall slope-channel system suggest about 2 m.y. of slope-channel system (Figure 3). The Bur1 system shows
deposition, ending at about 18 Ma. low to moderate sinuosity across a 40-km (25-mi)
286 / Porter et al.

reach and a well-developed east-to-west orientation. able sediment influx to the basin, local sea-floor gra-
This depositional trend is remarkably consistent, pre- dients associated with active salt movement, and a
sumably related to the regional slope gradient into regional depositional gradient that crosses linked ex-
the Lower Congo basin during deposition. Salt dia- tensional and contractional tectonic domains of a
pirs, extensional faults, and compressional folds seg- rapidly prograding shelf-slope system.
ment the slope profile into several large, north–
south-oriented anticlines and intervening synclines. Levee Confinement
These structural elements, early in their growth his- The Bur1 channel system is bounded by moderate-
tory, clearly influenced the width of this deep-water to low-amplitude seismic reflections that dip away at
channel system. The Bur1 channel system is more very low angles (<18) from the reservoir-prone axis
confined, narrow, and straight as it courses over struc- (Figure 4A). Downlap positions of these dipping strata
tural highs (Figure 3). Widths of the confined channel can be kilometers away from the trend of the channel
system are 700 –2200 m (2300 – 7200 ft) across posi- system (M. Grove, 2005, personal communication).
tive structural elements. Channel system sinuosities, Well penetrations show that these seismic facies are
measured from interpreted thalweg segments across low-net-to-gross silty mudstones interpreted here as
three anticlines in the study area, range between 1.06 flanking levees. Relief from the thalweg to levee crest
and 1.15 (n = 5). In each intervening syncline, the is on the order of 10–120 m (33–393 ft). Multiple epi-
channel system shows less erosional confinement sodes of levee growth and erosion are evident from the
and widths three to six times greater than over struc- stratal truncations observed on seismic traverses and
tural highs. Higher channel sinuosities (range 1.12 – comparison of levee thickness with the laterally ad-
1.73, n = 5) are developed in these laterally expanded jacent, more erosionally bounded channel complexes.
parts of the Bur1 channel system (Figure 3). Flanking levees are best developed at the base of the
channel system. Levee geometries are laterally more
System Depth and Relief of the Thalweg restricted in the younger parts of the system and
High-resolution 3-D seismic data show that the show classic geometries well documented in Pleisto-
composite fill thickness of the Bur1 slope system cene successions from the Amazon and Gulf of Mexico
varies along a 40-km (25-mi) depositional fairway in (Pirmez et al., 1997).
Block 15. This variation is a product of erosional and
depositional confinement of the channel complexes, Changes in Slope-channel System Confinement
proximity to growing salt structures, and accommo- across Structural Highs and Synclinal Lows
dation provided by extensional fault networks. Av- Erosional confinement of the Bur1 channel system
erage Bur1 thickness in the Kizomba development is is most pronounced near or across the current posi-
about 120 m (390 ft), with a range between 70 and tive structural elements. Channel segments that cross
more than 220 m (230 and more than 720 ft). Paleo- positive structural elements are relatively straight in
relief that developed on channel margins above the plan view, and this straight geometry sharply con-
channel thalweg can be estimated from the thickness trasts with the more highly sinuous channel patterns
of channel fill, stratal onlap positions, dimensions of in the intervening synclines. Significant positive re-
rotated slump blocks and their related displacement lief at the sea floor was not a factor during Bur1 de-
lengths, and cross sections from well-imaged, indi- position, as the Bur1 system crosses each of the salt-
vidual channel profiles. These aggregate data types cored anticlines or north – south extensional fault
suggest that the relief of Bur1 channels during depo- systems at right angles, without channel diversion
sition never exceeded 50 m (164 ft), and in some of or deflection. Net erosion of the sea floor by high-
the more weakly confined parts of the system, relief energy turbidity currents and levee growth was suf-
may have been as little as 5–10 m (16–33 ft). Thus, ficient to maintain the channel system’s position and
the Bur1 system is not a submarine canyon fill or the gradient across growing positive tectonic elements.
product of a single erosional event followed by depo-
sition. Instead, the composite Bur1 sediment fill is a Factors Controlling the External Form of the
product of multiple episodes of sediment gravity flow Confined Channel System
scour, headward erosion, and channel-margin slump- Extensive 3-D seismic coverage shows that lower
ing, alternating with episodic sedimentation by tur- Miocene slope-channel systems are large-scale de-
bidity and debris flows. Updip to downdip variations positional features that encompass more area than
in Bur1 channel system fill were controlled by vari- the well-imaged, high-amplitude channel packages
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 287

FIGURE 4. Seismic cross section through the Bur1 slope-channel system. (A) The seismic expression of the Bur1 is varied
because of multiple rock types and a stratigraphy built up through multiple episodes of erosion and deposition. The axial
parts of the system are seismically characterized as multicyclic, high-amplitude, semicontinuous seismic facies. Off-axis
and channel-margin successions show moderate- to low-amplitude response, and these strata onlap the erosional boundary
of the system. An Earth model of mixed-impedance channel fills and nonchannelized depositional elements (B) is
convolved with a seismic wavelet to produce a modeled seismic response (C) that is similar to the actual seismic traverse.

commonly illustrated in reservoir studies. Lateral di- term channel system trend across the slope and into
mensions are large, on the order of 2 – 20 km (1.2 – the basin. The paucity of coarse-grained lithofacies
12 mi). These low-amplitude seismic facies represent (i.e., coarse-grained sandstones, gravels, and conglom-
an integral component of the depositional system erates) in these flanking levee packages suggests large-
and are fundamentally related to the genesis of the scale sediment bypass down the channel system fair-
associated channel belts. Establishing the exact chro- way. Flow stripping of fine-gained suspension clouds,
nostratigraphic relationships between these deposits sea-floor gradients modified by mass transport com-
and those of the channel fills is difficult, but it is clear plex emplacement, and shelf-edge control of sediment
that they contain a stratigraphic record of protracted delivery access points determine the offset spacing
overbank processes. Thickness patterns and geom- of the channel system fairways.
etries suggest that these are net depositional units
formed via aggradation adjacent to evolving channel
complexes that were deposited from numerous tur- KIZOMBA FIELD DEVELOPMENT
bidity currents with strong lateral gradients in veloc-
ity, concentration, and depositional rates. Recogni- Field History
tion of these units is important in understanding the The early geoscience work and drilling history
evolution of the slope-channel system because these leading up to the Kizomba development is presented
widely dispersed, low-net facies establish the long- by Reeckmann et al. (2001). The phased development
288 / Porter et al.

FIGURE 5. Reservoir subdivision of the Bur1 slope-channel system. The Bur1 channel system can be subdivided into
four unconformity-bounded channel sets. This reservoir subdivision is based on the physical relationships of strata in
each channel set and contains lithofacies deposited in both channel and intrachannel areas. This approach avoids
overestimation of reservoir connectivity in assuming that all sandstones are time equivalent and recognizes lateral facies
changes into more shale-prone strata that baffle the reservoirs. Depth in meters.

consists of Kizomba A, a combination 36-slot TLP with Bur1 Well Penetrations in Block 15
a close-moored, 2.2-MMBO capacity FPSO (floating A total of 28 wells penetrate or are completed in
production, storage, and offloading vessel) (Figure 1). the Bur1 deep-water channel system (Figure 3). More
First oil was produced in August 2004, and the facility than half of the wells target multitiered, gravelly
in mid 2006 is producing more than 250,000 bbl/day. sandstone reservoirs in the Bur1 slope-channel sys-
Drilling activity in Kizomba A continues, with pres- tem. The high-angle, long-reach well design used in
sure support wells from five subsea drill centers. The Kizomba is optimized by volume-based interpre-
Kizomba B development, a virtual twin of Kizomba A, tation workflows in high-resolution 3-D seismic
began production in July 2005. In addition to res- cubes.
ervoirs accessible from the TLP, Kizomba B takes
production from the Dikanza subsea development
10 km (6 mi) to the northwest. Current production Distribution of Cored and Uncored Wells
from Kizomba B exceeds 250,000 bbl/day. Production- Four conventional cores (554 m [1817 ft] total)
focused activities in the Kizomba field include dril- establish the vertical lithofacies arrangement in the
ling pressure support wells, establishing production Bur1 channel-complex system. Three of the cored
in unpenetrated compartments, and shooting a four- wells establish a 27-km (16-mi) updip to downdip
dimensional, high-resolution seismic survey to man- transect of channel system sets in the Kizomba devel-
age field depletion. opment (Figure 5). The cores are described in detail,
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 289

FIGURE 6. Gravel-prone lithofacies in the Bur1 cores. Gravel-prone lithofacies in the Bur1 channel system are typically
thick bedded, moderate to poorly sorted bedsets bounded by sharp, erosive boundaries. The coarsest examples are R1
or R3 types that contain greater than 30% granule, pebble, or small cobble clasts. S1 and S3 bedsets contain 5 – 30%
granule to pebble clasts. Low-angle cross-stratification and clast imbrication are well developed in the R1 and S1 bedsets.
Color bar is 20 cm (8 in.) long.

and core results are used for seismic facies calibra- channel systems studied in subsurface and outcrop
tion, rock property studies, and inputs to geologic data sets (Raposo and Sykes, 1998; Sykes et al., 1998;
modeling. Forty kilometers (25 mi) to the north, wells Campion et al., 2000).
in Xikomba field produce from a time-equivalent
but geographically distinct Burdigalian deep-water
channel system, and these cores show a comparative Facies Association 1
channel-fill succession of turbidites and debrites. This is a gravel-prone facies association consisting of
clast-supported sandy gravels that display massive,
ungraded to normally graded clast fabrics (Figures 6, 7).
DEEP-WATER LITHOFACIES IN CORED WELLS
Description
Four facies associations are observed in cores from Bed thickness averages 1.3 m (4.2 ft), with a range
the Bur1 channel-complex system. These facies asso- between 0.3 and 8 m (1 and 26 ft) thick. Gravels and
ciations are distinct lithofacies successions of geneti- sandy conglomerates show a framework-supported
cally related bedsets, each characterized by distinctive texture and contain greater than 30% granule- to
sedimentary textures, bed thicknesses, and petrologic pebble-size clasts. Elongate pebble- to cobble-size
constituents. All the deep-water facies associations in clasts show well-developed planar to inclined, imbri-
the Bur1 slope system are common to other confined cated clasts between sharply erosional bed boundaries.
290 / Porter et al.

FIGURE 7. Gravel-prone, traction-dominated lithofacies associations. A vertical progression of lithofacies in gravel-prone,


traction-deposited bedsets consists of a lower part of very large shale rip-up clasts that are overlain by clast-supported
gravels and very coarse sandstones. These lithofacies associations are vertically well amalgamated and form the bases
of thick channel complexes as represented by a blocky well-log response. Core segment is 5 m (16 ft) long. Gamma-
ray (GR) and deep induction resistivity (ILD) log curves are depicted on the well log.

Very coarse sandstones are massive in texture and fabric, mapped distribution in high-resolution seismic
marked by thin, basal accumulations of granule con- data, and relation to channel-defined erosional cuts
glomerate with 5–30% granule and pebble clasts. Nor- suggest deposition as large gravelly bedforms (e.g.,
mal grading is common, and overall sorting is char- Winn and Dott, 1979). Gravel units developed during
acterized as poor to very poor. More than 80% of clasts maximum sediment throughput in active channel
described in this facies association are well-rounded, thalwegs by the accumulation of coarse bed load from
quartzofeldspathic, igneous, and schistose metamor- multiple flow events.
phic clasts, with minor amounts of well-indurated
sandstone and metavolcanic lithologies. Isolated
calcite-cemented concretions show convex upper and Facies Association 2
lower boundaries (Figure 8). Bed thickness, textures, This is a sand-prone facies association of thick-
and grain-size trends compare favorably with coarse- bedded, well- and poor-sorted structureless to rippled
grained deposits categorized as R1 and R3 bed types in sandstones (Figure 9).
Lowe (1982).
Description
Interpretation Structureless sandstones are medium to very fine
The gravel-rich beds in this facies association rep- grained, and bed thickness ranges between 0.06 and
resent bed-load –dominated deposition from high- 0.6 m (0.2 and 2 ft) thick, with an average bed thick-
concentration turbidity currents. The imbricated clast ness of 0.2 m (0.6 ft). Lower bed contacts are planar
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 291

FIGURE 8. Gravel-prone, traction-dominated lithofacies associations. Pebble- to granule-size lithofacies show massive
to weakly stratified textures in medium to thick (0.3 – 2-m; 1 – 6.6-ft) bedsets. This lithofacies association may contain
discrete, 0.5 – 1-m (1.6 – 3.3-ft) intervals of calcite cements. Convex, curved upper and lower boundaries defined by
the cementation front suggest an ellipsoidal, concretionary shape to these cements. Well-log response is blocky. Core
segment is 5 m (16 ft) long. Gamma-ray (GR) and deep induction resistivity (ILD) log curves are depicted on the well log.

and erosional; upper bed contacts are planar to gently al truncation of dewatering and fluid-escape struc-
curved. Sedimentary fabric is structureless and well tures (Figure 9). Grain fabric is typically massive and
sorted, although sparse swirled lamination is pres- homogeneous. Most of these thick-bedded sand-
ent in thicker bedded examples. These beds typically stones show a slight fining-upward grain-size trend.
overlie the gravel-prone lithofacies. Current-rippled Pebble-size floating shale clasts, swirled flow fabric,
sandstones include fine- to very fine-grained sand- and small granules are common components of these
stone and interbedded silty mudstone. Bed thickness beds. The macroscopic presence of clay-size matrix
ranges from 0.06 to 0.5 m (0.2 to 1.6 ft) and averages is inferred by the abundance of vertical dewatering
0.2 m (0.6 ft) thick. This facies association commonly pipes, swirled matrix flow structures, and floating,
caps the massive sandy turbidites, and contacts are matrix-supported clasts. Particle size analysis shows
conformable and generally planar in core. Sedimen- 1 – 8% clay content in the sandy matrix of these
tary structures are dominated by 1–5-cm (0.4–2-in.)- bedsets.
thick laminasets of current ripples. Minor amounts of
bioturbation and soft-sediment deformation are pres-
ent in the muddy sandstone beds. Interpretation
A distinctive suite of more poorly sorted sand- The vertical facies transition from underlying sandy
stones with a silty or muddy matrix is also developed conglomerates to these thick-bedded, well-sorted
in the sand-prone facies association. These medium sandstones suggest rapid suspension deposition by
to fine sandstones are 0.3 –1.5 m (1– 5 ft) thick, and sand-rich turbidity currents in a channel-axis setting.
bed-planar bed contacts are well defined by erosion- Rapid deposition is suggested by the thick-bedding,
292 / Porter et al.

FIGURE 9. Sand-prone, suspension-dominated lithofacies associations. Well-amalgamated sandstones show massive


to normal grading and sharp, conformable bed contacts. Minor amounts of granule or shale clast-rich lags define the
bases of many bedsets in this lithofacies association. Muddy sandstones of debrite origin are developed locally in
the core suite. Core segment is 5 m (16 ft) long. Gamma-ray (GR) and deep induction resistivity (ILD) log curves are
depicted on the well log.

conformable bed contacts, and uniform grain-size dis- Facies Association 3


tribution (Arnott and Hand, 1989). Massive sandstones This is an interbedded sand and mud lithofacies
are comparable to the Ta/Tb turbidite subdivision of association. Massive, current-rippled, and laminat-
Bouma (1962). ed bed types are identified on the basis of inter-
In contrast to the well-sorted sandstones, mixed nal sedimentary structures and presence of grading
grain sizes and sedimentary structures in more poorly (Figures 10–12).
sorted, clay-bearing sandstone beds suggest a sandy
debris-flow origin. The dispersed clast fabric, meter- Description
scale bedding, and abundant fluid-escape structures Interbedded sandy and muddy facies consist of
suggest that matrix strength was a profound control medium- to very fine-grained sandstone, siltstone,
on the movement and deposition of these sandy de- and mudstone in fining-upward, 0.1 – 0.3-m (0.3 –
brites (e.g., Middleton and Hampton, 1976; Iverson, 1-ft) beds. These beds are comparable to classic Bouma
1997). The matrix strength imparted by the small (1962) bed types and include current-rippled sand-
amount of clay dampened fluid turbulence and hin- stone, wavy to parallel-laminated sandy siltstone,
dered flow-related textural size segregation. Numeric and parallel-laminated mudstone (Figure 10). Sand-
and physical process modeling suggests that some stones are typically very well sorted.
thick-bedded sandstones previously described as the Structureless mudstones occur in 0.1 – 4-m (0.3 –
product of amalgamated Ta beds may, in fact, be the 13-ft)-thick beds with planar to highly contorted bed
product of sandy debrite deposition (Angevine et al., contacts. The matrix has a swirled or deformed ap-
1990; Shanmugam, 1996). pearance and commonly shows shear banding and
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 293

FIGURE 10. Spectrum of sandy and interbedded sandy-muddy lithofacies in Bur1 cores. This lithofacies suite spans a
complete range of sand-to-mud ratios and is best described as classic turbidites. Bed thickness is strongly related to sand
content. Well-amalgamated Ta and Tb beds are medium to thick bedded, whereas Tc , Tcd, and Tde are mud dominated
and thin bedded. Color bar is 20 cm (8 in.) long.

internal fabric variations on a centimeter scale. Float- channels. Accumulation and preservation of these
ing sand-size quartzose grains and larger organic detri- fine sandstones and mudstones were enhanced by
tus are common in these mudstones (Figures 13, 14). their off-axis position, removed from the erosive ef-
fects of subsequent high-concentration flows in the
channel thalwegs. It is likely that sediment gravity
Interpretation flows varied substantially in strength, sediment vol-
The current-rippled, heterolithic nature and di- ume, and texture during the active phase of channel-
versity of lamina types suggest that these mud-prone complex development. This shale-prone succession
beds were deposited from suspension from turbidity accumulated in the waning stages of channel fill and
currents (Arnott and Hand, 1989). Bed fabric in muddy abandonment, when channel avulsions likely changed
turbidites succession is comparable to Bouma Tde beds the locus of channel deposition to another part of the
(e.g., Bouma, 1962). Rippled sandy turbidites were slope complex.
deposited under waning-flow conditions in channel-
margin and proximal overbank settings. Dilute sedi-
ment plumes associated with each turbidity current Facies Association 4
spread laterally from the axial parts of deep-water This is an injected and slumped facies association.
294 / Porter et al.

FIGURE 11. Interbedded sandy-muddy lithofacies associations. Mud-prone lithofacies show parallel to wavy lamina sets
and a shaly log response in the intervening intervals between channelized reservoir packages. The core segment is
5 m (16 ft) long. Gamma-ray (GR) and deep induction resistivity (ILD) log curves are depicted on the well log.

Description clasts commonly show internal bedding or lamination


Concordant (sills) and discordant (dikes) sand- to be truncated at block margins. The blocks are em-
stones occur in a wide variety of shapes, thicknesses, bedded in sandy or muddy matrix.
and geometries in the Bur1 cores (Figure 15). Sand-
stones are well sorted, and grain size is very restricted, Interpretation
in the very fine- to fine-sand range. The internal struc- The injected sandstones are locally remobilized
ture is commonly massive, and some of the thicker sands derived from well-sorted, sandy bedforms. Sand-
discordant beds have small shale rip-up clasts at the stone dikes and sills in interchannel shales may be
upper and lower contacts (Figure 16). Injected sand- related to sediment loading under and adjacent to
stone thickness ranges from millimeter scale to 0.75 m rapidly deposited gravelly and sandy bedforms. Later
(2.5 ft) in dikes, and 0.4–1 m (1.3–3.3 ft) thick in sills. sand remobilization into overbank and capping aban-
In some sections, more or less isolated sills of sand- donment shale successions likely occurred during shal-
stone pass upsection into a complex network of ver- low burial of the sandy channel complexes. Lithology-
tical and horizontal injected sandstones throughout related permeability contrast between the sandy
a 10–15-m (33–49-ft) interval. channel fills and mud-prone overbank facies with
Stratigraphically below coarse-grained reservoir sec- compaction led to overpressure in porous sands. Hydro-
tions, many cored intervals contain 0.5–5-m (1.6– static pressure release forced these well-sorted sands
16-ft) blocks of folded and contorted sandstone and into confining mudstones as a series of dikes and sills.
mudstone blocks. Internal fabrics are always deformed The origin of the slumped strata is more problem-
and never parallel to master bedding surfaces. Large atic and may represent more than one depositional
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 295

FIGURE 12. Interbedded sandy and muddy lithofacies associations. Muddy debrites form the upper part of a well-
developed fining-upward well-log motif (arrow). The muddy debrites contain 2 – 20% sand in the clay-rich matrix. These
muddy debrites are overlain by thin-bedded Tc , Tcd, and Tde turbidites. Thickness of this lithofacies association is
commonly 2 – 8 m (6.6 – 26 ft). The core segment is 5 m (16 ft) long. Gamma-ray (GR) and deep induction resistivity (ILD)
log curves are depicted on the well log.

process and episode of deposition. Proximity of slump fill consists of thick-bedded gravels and very coarse
deposits to major erosional surfaces implies channel sandstones (Figure 17). Upper parts of the channel
instability and undercutting by energetic turbidity fills consist of interbedded sand-prone turbidites and
currents. Crosscutting relations of deformed strata mixed muddy and sandy debrites, thin-bedded muddy
with injected sandstones suggest that slump blocks turbidites, and injected sandstones. Multiple, stacked
failed from oversteepened cut banks and were buried channel-fill deposits can be grouped with laterally
prior to compaction by deposition of bed-load – associated channel-margin and overbank deposits to
dominated bar deposits. form genetically distinct channel complexes. Seismic
mapping of erosion surfaces, together with corre-
FACIES ARCHITECTURE OF THE BUR1 lation of low-net facies related to channel-complex
DEEP-WATER CHANNEL SYSTEM abandonment and lateral migration, provides a four-
fold, unconformity-bounded subdivision of the Bur1
Vertical and Lateral Facies Stacking Arrangements into channel-complex sets (e.g., Sprague et al., 2002)
High-resolution seismic data sets and numerous (Figures 5, 18).
well penetrations establish some general trends of
vertical and lateral facies relationships in the Bur1 Coarse-grained Gravels in Confined
slope-channel system. In most areas where current Channel Complexes
development activities are focused, gravel-prone litho- The gravel-prone facies succession shows a high-
facies are common in mapped channel-axis posi- impedance seismic character that allows detailed geo-
tions. The lower one third to one half of the channel body extractions in the Bur1 channel system (Figure 19).
296 / Porter et al.

FIGURE 13. Spectrum of debrite types in the Bur1 cores. A complete spectrum of lithofacies with floating, outsized
lithoclasts, swirled or deformed primary bedding geometries, dewatering pipes, and interstitial clay is present in the
Bur1 system. Completely sand-prone or completely mud-prone examples occupy the ends of a true spectrum of bed
types and are interpreted here as sandy debrites and muddy debrites. Most of the debrites cored in the Bur1 fit into the
middle part of the bed-type spectrum and are pragmatically separated into sandy-muddy debrites (poorer reservoir
quality) and muddy-sandy debrites (better reservoir quality) on the basis of sand in the matrix, water saturation,
permeability values, and bed thickness. Color bar is 20 cm (8 in.) long.

Seismic facies mapping in high-resolution 3-D cubes sition as determined from isochron maps. Instead,
indicates these geobodies to be an excellent proxy gravel bars splay 30 –808 off the downcurrent flow
for the size, distribution, and shape of both traction- direction. Interpretation of seismic amplitudes in the
deposited gravel bedforms and intervening intra- channel system suggests that gravel-prone units are
channel units, such as thick- to thin-bedded turbidites, not uniformly deposited across the width of the Bur1
and mixed sandy-muddy debrites (e.g., Goulding et al., system. Strong stratigraphic control of these units
2000; Beaubouef, 2004). by channel migration and incision is inferred by
Horizon-keyed amplitude extractions along the sinuous map patterns and cross-sectional geometries
erosional base of the Bur1 channel system show in- in channel complexes. Although late-stage incision
terpreted gravel bars and coarse-grained thalweg de- or unrelated subsequent erosion by turbidity currents
posits to be elongate and paddle shaped (Figure 19). cannot be discounted, the oblique orientation of the
Downflow-directed lengths are 5– 10 times longer preserved gravelly bars to the overall trend of the
than corresponding widths. The long-axis orientation sinuous channel suggests a primary, preserved depo-
of the gravel bars is never parallel to the thalweg po- sitional extent to these deposits.
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 297

FIGURE 14. Mud-prone lithofacies associations. Mixed muddy debrites form highly varied successions of thick- and
thin-bedded, contorted facies successions. Log response is serrate to bell shaped. Thin, interbedded Tc and Tcd turbidites
commonly punctuate the thicker debrite-rich intervals. Permeability contrast between turbidite-deposited sandstones
and muddy debrites is indicated by the degree of hydrocarbon saturation. The core segment is 5 m (16 ft) long. Gamma-
ray (GR) and deep induction resistivity (ILD) log curves are depicted on the well log.

Implications of the Gravel-prone Facies Successions The lower Miocene shelf margin, identified on re-
The abundance of gravel-prone lithofacies in chan- gional seismic lines, was at least 40–50 km (25–31 mi)
nel complexes of the Bur1 system illustrates some im- to the east of the study area. The competency of trans-
portant geologic features and processes that operate porting sediment gravity flows was considerable.
within large slope-channel systems. These gravelly Clast sizes of as much as 45 cm (17 in.) in diameter
sediments are clearly associated with channel erosion are present in the cored suite, and pebbles to small
and fill, organized in bedforms, and do not represent cobbles comprise most of the gravel-prone bedsets.
eroded coarse-grained remnants of older strata. Frame- Larger framework clast sizes likely exist in these de-
work clast petrology is essentially that of Proterozoic posits but are not represented in the cored lithofacies
igneous and metamorphic rock suites of the Angolan because of the mechanical difficulty of coring in largely
shield, and this implies an updip, direct shelfal input unconsolidated reservoirs by extended-reach wells.
of gravels or cannibalization of older, coarse-grained Low-angle cross-stratification, imbricate clast fab-
shelf-margin fluvial-mouth bar deposits by shelf- rics, and grain-size trends suggest that the gravelly
edge failure and headward erosion up the channel units were deposited as traction carpets by high-
systems. A relatively narrow shelf and Tertiary-aged concentration turbidites (e.g., Winn and Dott, 1977;
regional tectonic uplift were important factors in Lowe, 1982). Effective transport of such coarse-grained
keeping the lower Miocene slope-channel systems of lithofacies more than 50 km (31 mi) from the time-
Angola charged with coarse-grained sediments (Brice equivalent shelf margin implies tremendously energetic
et al., 1982). sediment gravity flows into the Lower Congo basin.
298 / Porter et al.

FIGURE 15. Mud-prone lithofacies associations. Thick successions of muddy debrites are uniformly gray in color and
may show discrete, well-sorted disconformable intervals of fine-grained sandstone. Crosscutting relations of the
sandstone with the bedding geometry of muddy debrites suggest a postdepositional injection origin of the sandstones.
The shaly log response is dependent on the amount of injected sandstone; those intervals with abundant sandstone
injections show a highly serrate log pattern. Core segment is 5 m (16 ft) long. Gamma-ray (GR) and deep induction
resistivity (ILD) log curves are depicted on the well log.

Deposition of the gravel units was focused in the Depositional History of the Bur1
deeper thalweg part of channel complexes. Erosional Slope-channel System
bedset boundaries suggest episodic development of The integration of high-resolution 3-D data, cores,
the barforms, and numerous high-concentration and well logs demonstrates a composite sedimentary
flows likely contributed to the volume and arrange- record for the Bur1 slope-channel system. Cored litho-
ment of the gravels preserved in the channel fills. facies show both turbidity current and debris-flow
Because the gravelly units may represent only a deposits in the composite fill of the channel system
small fraction of the total sediment load in large, (Figure 20). Multiple episodes of erosion and depo-
turbulent sediment gravity flows, the preservation of sition produced at least four unconformity-bounded
thick, amalgamated gravelly packages implies that much channel-complex sets over the span of 2 m.y. Each of
greater amounts of sand and mud have completely these channel sets shows both lateral and vertical
bypassed the study area. It is likely that these missing connectivity and can be subdivided into an axis
sediments built downdip depositional elements of the dominated by gravel-rich traction deposits, highly
large, linked deep-water slope and basin depositional amalgamated sandstones, and laterally flanking
system. It is without a doubt that these depositional strata, which contain thin-bedded sandstone and
elements of the lower slope and basin deep-water systems mudstones, injected sandstones, and slumps depos-
will be better understood as exploration pushes into the ited in channel-margin, levee, and overbank areas of
ultradeep parts of the Lower Congo basin. the Bur1 slope-channel system.
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 299

FIGURE 16. Injected sandstone and debrite lithofacies associations. This segment shows some of the facies organization
at the base of the Bur1 system. The slope-channel system boundary (white dashed line) can be tied to seismic data
as a regional mapping horizon. The contact is sharp and uncorformably separates the Bur1 channel system from
older slope mudstones (gray lithologies). The lithofacies succession that sits directly on this unconformity is highly
variable. In this core, sandy debrites, slump blocks, and injected sandstones overlie the contact. Core segment is
5 m (16 ft) long. Gamma-ray (GR) and deep induction resistivity (ILD) log curves are depicted on the well log.

The general similarity of channel stacking pattern lower Miocene deep-water system. We also express
and constituent channel-fill elements in a 40-km our gratitude to Sociedade Nacional de Combustiveis
(25-mi) reach of the Bur1 channel system suggests de Angola (Sonangol) for the opportunity to study
that sediment gravity flow erosion and deposition and develop these world-class reservoirs. Our work
extend over long distances (tens to hundreds of kilo- includes substantial contributions from numerous
meters). The Bur1 succession in the Kizomba devel- prospect and development geoscientists in the Ex-
opment gives a glimpse into the variations of facies ploration and Development companies of ExxonMo-
architecture constructed by highly energetic and com- bil. We thank John Ardill, Dana Butters, Tim Fahrer,
petent sediment gravity flows traversing the mid- E-Chien Foo, Matt Grove, Randy Kissling, Dave Mason,
slope of the Lower Congo basin. Lisa McBee, Jose Sequeira, Mark Rosin, and Bill Tate
for their insight to the geology of Bur1 slope-channel
system. Reviews by Laurence Droz (University of Brest)
ACKNOWLEDGMENTS and Carlos Pirmez (Shell Oil) greatly improved the
manuscript. The views expressed in this study are
The authors thank ExxonMobil, together with An- those of the authors and ExxonMobil and do not
gola Block 15 coventurers British Petroleum, Statoil, necessarily reflect those of the concessionaire or the
and ENI for permission to publish our work on this Block 15 contractor group.
300 / Porter et al.

FIGURE 17. Depositional environments interpreted from the vertical succession of lithofacies association in the Bur1
channel system. Each of the channel complexes in the Bur1 shows a progression of mixed debrites, turbidites, and
slumps that are overlain by turbidite-dominated facies associations of well-amalgamated gravel- and sand-prone
traction deposits. Because of the admixture of muddy debrites in the channel complexes, log response would suggest a
nonamalgamated, disconnected prediction of channel-fill connectivity. Erosion associated with the gravelly lithofacies
produces lateral and vertical amalgamation of the channel complexes, and dynamic well performance demonstrates
better connectivity than would be inferred by log response. Gamma-ray and Vshale log response of stacked facies
associations is on the leftmost track with core photo annotations; resistivity, porosity, density, and water-saturation
curves are shown on the right log tracks. Depth in meters.
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 301

FIGURE 18. Vertical lithofacies succession and stratal


hierarchy in the Bur1 slope-channel system. Detailed
facies description of Bur1 cores shows the relationship
of bed and bedset elements to larger stratal elements
in a deep-water hierarchy. In this example, yellow
bedsets are gravel- or sand-prone lithofacies deposited
by high- and low-concentration turbidity currents.
Brown-colored units are sandy and muddy sand de-
brites. Gray-colored lithologies are an admixture
of muddy turbidites, debrites, and hemipelagic shales.
The vertical facies log shows the axial expression of
three stacked channel complex sets in this part of the
Bur1 slope-channel system. Leftmost well-log curves
are gamma ray and Vshale ; shallow and deep resistivity
curves are displayed on the right track. The length
of the core description is 110 m (360 ft).
302 / Porter et al.

FIGURE 19. Seismic expression of gravel-prone units in the Bur1 slope-channel system. Four seismic horizon-keyed
extractions of the high-impedance, lower parts of Bur1 channel complexes show the geometry and dimensions of
gravel-rich units. (A) Scrollwork pattern of gravel-rich units deposited within sinuous channels of the Bur1 system. These
units represent the lower half of the composite channel complex fills and are overlain by sand- and mixed sand- and
mud-prone lithofacies. (B) Stratigraphically younger channel complex shows smaller proportion of gravel units in an
overall more sand-prone channel complex set. Scrollwork pattern is developed from laterally migrating channels, but
gravelly units are more isolated than in the channel complex set illustrated in (A). (C) Strongly confined part of the Bur1
system shows the abundance of gravel thalweg units in green. Well penetrations that establish intervening areas are
sandstones deposited by high-concentration turbidity currents. (D) Volume seismic sculpt shows relationship of gravels
with overlying sandstones in a single channel complex set of the Bur1. Well penetrations show blue units as high-
impedance gravels and the red units as lower impedance sandstones.
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 303

FIGURE 20. Depositional model for the Bur1 slope-channel system. The vertical and lateral relationships of turbidite
and debrite deposits are represented in this cross-sectional depositional model of the Bur1 confined slope-channel
system. Core-calibrated logs define the axial facies succession. Debrite and gravelly channel fills define the deepest parts
of erosionally based channel sets, marked by light-blue lines. The overall vertical succession shows most of the gravel-
prone lithofacies in the lower half of the channel system, and the upper fill succession contains more mixed sandstone,
mudstones, and muddy debrites. Seismic facies mapping is used to place depositional elements in the off-axis and
margin areas of the channel system. These lateral facies are thin-bedded mixed sandstones and mudstones deposited
in channel-related, stacked levee margins.
304 / Porter et al.

REFERENCES CITED Gottschalk, R. R., 2002, The Lower Congo basin, deep-
water Congo and Angola: A kinematically linked
Angevine, C. L., P. L. Heller, and C. Paola, 1990, Quan- extensional/contractional system (abs.): AAPG Bulle-
titative sedimentary basin modeling: AAPG Continu- tin, v. 86, p. 66.
ing Education Course Note Series 32, 133 p. Gottschalk, R. R., A. V. Anderson, J. D. Walker, and J. C.
Ardill, J. A., T. C. Huang, and O. McLaughlin, 2002, The Da Silva, 2004, Modes of contractional salt tectonics
stratigraphy of the Oligocene to Miocene Malembo in Angola Block 33, Lower Congo basin, west Africa,
formation of the Lower Congo basin, offshore Angola in 24th Annual Gulf Coast Section SEPM Research
(abs.): AAPG Annual Meeting Expanded Abstracts, Conference, Salt-sediment Interactions and Hydro-
p. 9. carbon Prospectivity: Concepts, Applications, and
Arnott, R. W. C., and B. M. Hand, 1989, Bedforms, pri- Case Studies for the 21st Century, Houston, Texas,
mary structures and grain fabric in the presence of December 5 – 8, 2004, 30 p.
suspended sediment rain: Journal of Sedimentary Goulding, F. J., T. R. Garfield, K. W. Rudolph, G. N. Jensen,
Petrology, v. 59, p. 1062 – 1069. and R. T. Beaubouef, 2000, Seismic/sequence stratig-
Babonneau, N., B. Savoye, M. Cremer, and B. Klein, 2002, raphy of deep-water reservoirs: 1. Seismic facies recog-
Morphology and architecture of the present canyon nition criteria: Past experience and new observations
and channel system of the Zaire deep-sea fan: Marine (abs.): AAPG Annual Meeting Program, v. 9, p. A56.
and Petroleum Geology, v. 19, p. 445 – 467. Hartman, D. A., W. A. Swanson, P. R. Smith, F. J.
Beaubouef, R. T., 2004, Deep-water leveed-channel com- Goulding, and C. A. Kelly, 1998, Structural develop-
plexes of the Cerro Toro Formation, Upper Cretaceous, ment of the continental margin of Congo and north-
southern Chile: AAPG Bulletin, v. 88, p. 1471 – 1500. ern Angola (abs.): AAPG Bulletin, v. 82, p. 1923.
Beaubouef, R. T., T. R. Garfield, and F. J. Goulding, 1998, Iverson, R. M., 1997, The physics of debris flows: Reviews
Seismic stratigraphy of depositional sequences: High in Geophysics, v. 35, p. 245 – 296.
resolution images from a passive margin slope setting, Kolla, V., P. Bourges, J. M. Urruty, and P. Safa, 2001, Evo-
offshore west Africa (abs.): AAPG Bulletin, v. 82, p. 1980. lution of deep-water Tertiary sinuous channels off-
Bouchet, R., B. Levallois, G. Mfonfu, and J.-F. Authier, shore Angola (west Africa) and implications for reser-
2005, Optimizing development of Angola’s Girassol voir architecture: AAPG Bulletin, v. 85, p. 1371 – 1405.
field: World Oil, v. 226, no. 3, p. 45 – 48. Larson, R. L., and J. W. Ladd, 1973, Evidence from magnetic
Bouma, A. H., 1962, Sedimentology of some flysch de- lineations for the opening of the South Atlantic in the
posits: Amsterdam, Elsevier, 162 p. Early Cretaceous: Nature, v. 246, p. 209 – 212.
Brice, S. E., M. D. Cochran, G. Pardo, and A. D. Edwards, Lowe, D. R., 1982, Sediment gravity flows: II. Depositional
1982, Tectonics and sedimentation of the south Atlan- models with special reference to the deposits of high-
tic rift sequence, Cabinda, Angola, in J. S. Watkins and concentration turbidity currents: Journal of Sedimen-
C. L. Drake, eds., Studies in continental margin geology: tary Petrology, v. 52, p. 279 – 297.
AAPG Memoir 34, p. 518. Middleton, G. V., and M. A. Hampton, 1976, Subaqueous
Burke, K., 1996, The African plate: South African Journal sediment transport and deposition by sediment
of Geology, v. 99, p. 341 – 409. gravity flows, in D. J. Stanley and D. J. P. Swift, eds.,
Campion, K. M., A. R. Sprague, D. Mohrig, R. W. Lovell, Marine sediment transport and environmental man-
P. A. Drzewiecki, M. D. Sullivan, J. A. Ardill, G. N. agement: New York, John Wiley and Sons, p. 197 –
Jensen, and D. K. Sickafoose, 2000, Outcrop expression 218.
of confined channel complexes: Gulf Coast Section Pirmez, C., R. N. Hiscott, and J. D. Kronen Jr., 1997, Sandy
SEPM Foundation 20th Annual Research Conference, turbidite successions at the base of channel-levee
Houston, Texas, p. 127 – 150. systems of the Amazon Fan revealed by FMS logs and
Droz, L., T. Marsset, H. Ondreas, M. Lopez, B. Savoye, and cores: Unraveling the facies architecture of large
F.-L. Spy-Anderson, 2003, Architecture of an active submarine fans, in R. D. Flood, et al., eds., Proceedings
mud-rich turbidite system: The Zaire Fan (Congo – of the Ocean Drilling Program, Scientific Results,
Angola margin southeast Atlantic): Results from Zai- v. 155, p. 7 – 33.
Ango 1 and 2 cruises, AAPG Bulletin, v. 87, p. 1145 – Rabinowitz, P. D., and J. LaBreque, 1979, The Mesozoic
1168. South Atlantic Ocean and evolution of its continental
Duval, B., C. Cramez, and M. P. A. Jackson, 1992, Raft margins: Journal of Geophysical Research, v. 84, p. 5973–
tectonics in the Kwanza Basin, Angola: Marine and 6002.
Petroleum Geology, v. 9, p. 389 – 404. Raposo, A. J. M., and M. A. Sykes, 1998, Exploration for
Emery, K. O., and E. Uchupi, 1984, The geology of the deep-water reservoirs, offshore Angola (abs.): AAPG
Atlantic Ocean: New York, Springer-Verlag, 1050 p. Bulletin, v. 82, p. 1956.
Garfield, T. R., D. C. Jennette, F. J. Goulding, and D. K. Reeckmann, S. A., D. K. S. Wilkin, and J. Flannery, 2001,
Sickafoose, 1998, An integrated approach to deep- Kizomba, a deep-water giant field, Block 15, Angola,
water reservoir prediction (abs.): AAPG Bulletin, v. 83, in M. T. Halbouty, ed., Giant oil and gas fields of the
p. 1314. decade 1990 – 1999: AAPG Memoir 78, p. 227 – 236.
Lower Miocene Slope-channel System in Block 15, Offshore Angola / 305

Schollnberger, E., and P. R. Vail, 1999, Seismic stratigraphy reservoir presence and quality in offshore west Africa:
of the Lower Congo, Kwanza, and Benguela basins, off- E-Exitep Proceedings 2005, Veracruz, Mexico, 13 p.
shore Angola, Africa (abs.): AAPG Bulletin, v. 83, Sykes, M. A., D. Mohrig, C. Rossen, and T. R. Garfield, 1998,
p. 1338. Lithofacies associations within complex slope channel
Scrutton, R. A., and R. V. Dingle, 1976, Observations on reservoirs: Debrites and turbidites (abs.): AAPG Bulletin,
the processes of sedimentary basin formation at the v. 82, p. 1973.
margin of southern Africa: Tectonophysics, v. 36, p. 143– Temple, F., and O. Broucke, 2004, Sedimentological
156. models of the Oligocene and Miocene Malembo for-
Shanmugam, G., 1996, High density turbidity currents: mation in offshore Angola (Lower Congo basin) (abs.):
Are they sandy debris flows?: Journal of Sedimentary 1st Nigerian Association of Petroleum Explorationists–
Research, v. 66, p. 2 – 10. American Association of Petroleum Geologists West
Sprague, A. R. et al., 2002, The physical stratigraphy of Africa Deepwater Conference, Abuja, Nigeria, p. A46.
deep-water strata: A hierarchical approach to the anal- Winn, R. D., and R. H. Dott Jr., 1977, Large-scale traction
ysis and genetically related stratigraphic elements for produced structures in deep-water fan-channel con-
improved reservoir prediction (abs.): AAPG Annual glomerates in southern Chile, Geology, v. 5, p. 41 – 44.
Meeting Expanded Abstracts, p. 167. Winn, R. D., and R. H. Dott Jr., 1979, Deep-water fan-
Sprague, A. R. et al., 2005, Integrated slope channel de- channel conglomerates of Late Cretaceous age, south-
positional models: The key to successful prediction of ern Chile: Sedimentology, v. 26, p. 203 – 228.
9
Dubois, M. K., A. P. Byrnes, G. C. Bohling, and J. H. Doveton, 2006,
Multiscale geologic and petrophysical modeling of the giant Hugoton
gas field (Permian), Kansas and Oklahoma, U.S.A., in P. M. Harris and
L. J. Weber, eds., Giant hydrocarbon reservoirs of the world: From rocks
to reservoir characterization and modeling: AAPG Memoir 88/SEPM
Special Publication, p. 307 – 353.

Multiscale Geologic and


Petrophysical Modeling of the
Giant Hugoton Gas Field
(Permian), Kansas and
Oklahoma, U.S.A.
Martin K. Dubois, Alan P. Byrnes, Geoffrey C. Bohling, and John H. Doveton
Kansas Geological Survey, University of Kansas, Lawrence, Kansas, U.S.A.

ABSTRACT

R
eservoir characterization and modeling from pore to field scale of the
Hugoton field (central United States) provide a comprehensive view of a
mature giant Permian gas system and aid in defining original gas in place
and the nature and distribution of gas saturation and reservoir properties. Both
the knowledge gained and the techniques and workflow employed have impli-
cations for understanding and modeling reservoir systems worldwide that have
similar geologic age and reservoir architecture (e.g., Gwahar and North fields, Persian
Gulf). The Kansas–Oklahoma part of the field has yielded 34 tcf (963 billion m3)
gas throughout a 70-yr period from more than 12,000 wells. Most remaining gas
is in lower permeability pay zones of the 557-ft (170-m)-thick, differentially de-
pleted, layered reservoir system.
The main pay zones represent 13 shoaling-upward, fourth-order marine-
continental cycles comprising thin-bedded (6.6–33-ft; 2–10-m), marine carbon-
ate mudstone to grainstone and siltstones to very fine sandstones and have re-
markable lateral continuity. The pay zones are separated by eolian and/or sabkha
red beds having low reservoir quality. Petrophysical properties vary among 11 major
lithofacies classes. Neural network procedures, stochastic modeling, and automation
facilitated building a detailed full-field three-dimensional (3-D) 108-million-cell
cellular reservoir model of the 10,000-mi2 (26,000-km2) area using a four-step
workflow: (1) define lithofacies in core and correlate to electric log curves (training
set); (2) train a neural network and predict lithofacies at noncored wells; (3) popu-
late a 3-D cellular model with lithofacies using stochastic methods; and (4) popu-
late model with lithofacies-specific petrophysical properties and fluid saturations.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215881M883274

307
308 / Dubois et al.

INTRODUCTION North America. Covering southwest Kansas and parts


of the Oklahoma and Texas panhandles, these fields
The focus of this chapter is the definition and are situated in the Hugoton embayment of the Ana-
integration of core-defined lithofacies, core-derived darko basin (Figure 1). Since its discovery in 1922 and
petrophysical properties, wire-line-log response, and development in the 1950s, 34 tcf gas (963 billion m3)
estimation of reservoir properties to characterize a have been produced from greater than 12,000 wells
giant reservoir system at the core, well, and field scale. across 6200 mi2 (16,000 km3) in the Kansas and Okla-
Central to the effort is the use of core-defined litho- homa part of the Hugoton field (Figure 2). Unless
facies to train a neural network to predict lithofacies otherwise noted, the term ‘‘Hugoton’’ in this chapter
at wells without core. We discuss each step of the combines the Hugoton (Kansas), Panoma (Kansas and
workflow, many aspects of which could be applied Oklahoma), and the Guymon-Hugoton (Oklahoma)
in other settings. The results and summary are an early fields. Production is from the Lower Permian Chase
view of work in progress that is part of an ongoing Group and Council Grove Group (Figure 3). In most
collaborative project. areas inside the Panoma boundary, the gas column is
The importance of the Hugoton field study ex- continuous between the two stratigraphic intervals
tends beyond the borders of Kansas and Oklahoma. (Pippin, 1970; Parhman and Campbell, 1993) and
Both the knowledge gained and the techniques em- reaches a maximum thickness of 500 ft (150 m) in the
ployed have implications for understanding and mod- west-central part of the study area. One exception
eling reservoir systems worldwide that have similar may be in a relatively small part of the field near the
geologic age, reservoir architecture, production char- western margin in Morton County, Kansas, that is
acteristics, problems in determining water saturation, described by Olson et al. (1997) as being compart-
large data sets, split ownership, or maturity. The full- mentalized by faults. In Oklahoma, production out-
field model of the 10,000-mi2 (26,000-km2) reservoir lined as ‘‘other Council Grove’’ (Figure 1) is from in-
area provides a detailed three-dimensional (3-D) view tervals in the Council Grove that are as much as 300 ft
of 13 shoaling-upward cycles vertically stacked in a (100 m) below the lowest perforations in the Chase.
low-relief shelf setting. The nature of the model and The reservoir is shallow, with depth to the top of the
its construction provides a good analog for similar Chase ranging from 2100 to 2800 ft (640 to 850 m)
thin, stacked-cycles reservoir systems, including the and lower and upper productive limits, referenced to
Aneth field in the Paradox basin (Weber et al., 1994; sea level, of approximately +100 ft (+30 m) on the east
Grammer et al., 1996), fields in the prolific Permian and +1250 ft (+380 m) on the western updip mar-
basin of west Texas (Dutton et al., 2005), and the gin, respectively. Original wellhead shut-in pressure
Khuff Formation in Gwahar and North fields in the in Kansas was 437 psi (3013 kPa) (Hemsell, 1939), sig-
Persian Gulf (McGillivray and Husseini, 1992; Kon- nificantly less than half of a seawater gradient, and
nert et al., 2001). Fine-scale cellular models are par- similar, anomalous initial pressures were recorded in
ticularly important for modeling thin-layered, differ- Oklahoma (Sorenson, 2005). Average 72-h wellhead
entially depleted reservoir systems, and methods used shut-in pressure in Kansas in 2003 was 32 psi (221 kPa).
in building the model demonstrate the construction Annual production in 2004 was 265 bcf (7.5 billion m3).
of a cellular petrophysical model for a giant field. Early completions in the Chase were commonly open
The study also demonstrates the benefits of pooling hole with a slotted liner followed by a large acid treat-
proprietary geologic and engineering data in set- ment. After 1960, typical completions commonly in-
tings having split ownership (Sorenson, 2005). As the volve casing, perforation, and acidizing as many as
world’s fields mature, high-resolution modeling at six zones separately, followed by a large hydraulic sand
the full-field scale in data-rich environments will be- fracture treatment, sometimes exceeding 200,000 lb
come increasingly important, and we present a large- (91,000 kg) of sand, to the entire perforated interval
scale example for developing such models. (Hecker et al., 1995).
Although much has been published on the Hugo-
ton throughout the 70-yr life of the field, most of the
Background studies were broad in scope (Hemsell, 1939; Mason,
The combined Kansas Hugoton and Panoma, Texas 1968; Pippin, 1970). Sorenson’s (2005) recent article
Hugoton, and West Panhandle fields, with an esti- that presents a paleostructural and pressure history for
mated ultimate recovery of 75 tcf (2.1 trillion m3) gas the reservoir system stretching from the Texas Pan-
(Sorenson, 2005) represent the largest gas field in handle to west-central Kansas provides a good recent
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 309

FIGURE 1. Regulatory bound-


aries for Permian (Wolfcam-
pian) gas and oil fields, Kansas,
Oklahoma, and Texas. Wolf-
campian structure in feet, after
Pippin (1970) and Sorenson
(2005).

homa Panhandle. Following


the Kansas Corporation Com-
mission proration order per-
mitting a second well in each
unit, several studies on res-
ervoir characterization (Sei-
mers and Ahr, 1990; Caldwell,
1991; Olson et al., 1997) and
reservoir simulation (Fetko-
vitch et al., 1994; Oberst et al.,
1994) were published. Past
studies by industry have been
generally confined to areas
where they have assets and
data. This chapter provides
details of the most compre-
hensive reservoir characteri-
zation effort to date, both
geographically and strati-
graphically (the entire reser-
voir system, Chase and Coun-
cil Grove groups).
Work related to this study
is part of a collaborative, mul-
tidisciplinary study of the
Hugoton field, supported by
10 industry partners having
gas resources in southwest
Kansas and the Oklahoma
Panhandle (Dubois et al.,
2005). The primary purpose
of the 2-yr study is to devel-
op a comprehensive, field-
wide geologic model that can
be used to quantify and locate
remaining gas for improved
reservoir management. No
overview of the field history and prior work. Detailed field-scale model exists primarily (1) because of the
studies involving reservoir characterization have been immense size of the field and (2) because the most
limited geographically and stratigraphically. For ex- critical petrophysical data were proprietary. The cur-
ample, Seimers and Ahr (1990) investigated the Chase rent project was facilitated by the creation of a co-
in the Oklahoma Panhandle; Olson et al. (1997) operative geologic and engineering data set amassed
studied the Kansas Chase; and Heyer (1999) focused from participating companies, which allowed a com-
on the Council Grove in a small area of the Okla- prehensive global view of the entire reservoir system.
310 / Dubois et al.

is critical. Because only a relatively small number of


wells were cored and the reservoir volume is im-
mense, a methodology was developed for predicting
lithofacies from wire-line-log response at wells not
cored and then between wells.

RESERVOIR GEOLOGY
Regional Geology
The Hugoton field lies on the west side of the Hugo-
ton embayment of the Anadarko basin and is bounded
to the northwest by the Las Animas arch and to the
northeast by the central Kansas uplift. The Anadarko
FIGURE 2. Gas production from the Wolfcampian (Hugo- basin is an asymmetric foreland basin associated with
ton and Panoma fields) in the Kansas part of the study the early Pennsylvanian Ouchita–Marathon orogeny,
area through 2004. The Oklahoma part of the study area and the Hugoton embayment and the rest of the Kansas
produced 7 tcf (198 billion m3) Wolfcampian gas in the shelf form the flatter side of the asymmetry (Figure 4).
same period that 27 tcf (765 billion m3) was produced
in Kansas. The spike in production beginning in the early
The Anadarko basin was initiated and had its greatest
1980s was caused by infill drilling the Hugoton field in subsidence in Pennsylvanian-Morrowan time, with
Kansas. subsidence rates decreasing through the Permian. The
basin was nearly filled by the end of the Wolfcampian
when the Anadarko basin was covered by shelf carbon-
Although the field is very mature, individual com- ates (Kluth and Coney, 1981; Rascoe and Adler, 1983;
panies possess excellent modern wire-line-log, core, Kluth, 1986; Perry, 1989).
and engineering data in their asset areas because of Marine carbonate reservoirs thin toward the up-
drilling for deeper production and the infill drilling dip margin and many pinch out at, or just west of,
program that took place in the late 1980s and early the field margin, particularly in the Council Grove.
1990s.

Problem and Approach


A principal goal of the project is to quantify the
nature and distribution of gas saturation and reservoir
properties and original and remaining gas in place.
Determining the gas in place and its distribution is
hampered by three significant obstacles: (1) accurate
determination of water saturations using conven-
tional wire-line logs is not possible because of deep
invasion by filtrate for typical drilling programs (Olson
et al., 1997; George et al., 2004); (2) wellhead shut-in
pressures are strongly influenced by high-permeability
interval properties and do not accurately represent all
interval pressures; and (3) the reservoir is layered,
differentially depleted, and pressure data for individ-
ual layers are minimal. To define the remaining gas
in place, we used an integrated approach using core, FIGURE 3. Stratigraphic column, Hugoton field area, with
core-derived lithofacies-specific petrophysical relation- the names of gas fields in Kansas and Oklahoma adjacent
to the intervals from which they produce (compiled
ships, and engineering data. Because petrophysical
from Zeller, 1968; Pippin, 1970; Baars, 1994; D. P. Merriam,
property relationships (e.g., permeability-porosity, 2006, personal communication). The combined Hugoton
capillary pressure) differ among lithofacies, the con- and Panoma fields in Kansas and the Guymon-Hugoton
struction of a geomodel with appropriate lithofacies field in Oklahoma are lumped as Hugoton in this chapter.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 311

FIGURE 4. Distribution of major litho-


facies in the mid-continent during
the late Wolfcampian (modified from
Rascoe, 1968; Rascoe and Adler, 1983;
Sorenson, 2005; used with permission
from the Rocky Mountain Association
of Geologists and AAPG). Approximate
paleolatitude was 38N (Scotese, 2004).

basinward at a rate of approximately


1.2 ft/mi (0.24 m/km) to location on
the shelf where the rate of thicken-
ing increases by a factor of 10. The
axis of thickening is coincident with
an area of present-day steep dip and
may mark a shelf margin or the
axis of a steepened slope. It is also
nearly coincident with the edge of
a Virgillian-age starved basin and
the transition from marine carbon-
ate to marine shale (Rascoe, 1968;
Rascoe and Adler, 1983). Dubois and
Goldstein (2005) estimated the max-
imum relief across the Kansas part of
the shelf during Council Grove de-
position to have been 100 ft (30 m),
with a slope of approximately 1 ft/mi
(0.2 m/km). Notable is the absence
Red continental rocks, primarily very fine to coarse on the Hugoton shelf of dark, fissile shale, a common
siltstones, are thickest at the margin and thin basin- deep-water lithofacies in the Wolfcampian in out-
ward across the shelf (Figure 5). The red beds have crop in eastern Kansas and northeast Oklahoma (Maz-
been thought by many to be the lateral seal that, zullo et al., 1995; Boardman and Nestell, 2000), sug-
when accompanied by a Leonardian-age evaporite gesting that water depths on the Hugoton shelf were
top seal, created a giant stratigraphic trap (Garlough less than those at the present-day outcrop 300 mi
and Taylor, 1941; Mason, 1968; Pippin, 1970; Parh- (480 km) to the east. The closest equivalent to the typ-
man and Campbell, 1993). However, laterally con- ical deep-water lithofacies in Hugoton core are dark,
tinuous, high-porosity and high-permeability, ma- marine siltstones found near the base of the marine
rine, and continental sandstones are common at the carbonate intervals in four cycles, the Grenola (C_LM),
updip margin in the northwest part of the field. Funston (A1_LM), Wreford, and Ft. Riley. For this chap-
These rocks are gas productive inside the field bound- ter, we will refer to most of the extremely gently slop-
aries and water saturated outside the field despite ing area as shelf and the area of steeper dip and strati-
being in a higher structural position and without graphic thickening as the shelf margin.
evidence of a physical barrier. These conditions argue
against these rocks being a lateral seal and suggest Reservoir Lithofacies
that mechanisms other than lithofacies change alone The Hugoton in Kansas and Oklahoma produces gas
are responsible for trapping (Dubois and Goldstein, from 13 fourth-order marine-continental (carbonate-
2005). siliciclastic) sedimentary cycles (Figure 8), six in the
Present-day structure of Wolfcampian-age rocks Chase and seven in the Council Grove, reflecting rapid
was strongly influenced by a Laramide-age eastward glacioeustatic sea level fluctuations (Olson et al.,
tilt (Figure 6), whereas the Wolfcampian isopach 1997; Heyer, 1999; Boardman et al., 2000; Olszewski
(Figure 7) better reflects the shelf geometry. From the and Patzkowsky, 2003). The marine and continen-
west field margin, the Wolfcampian strata thicken tal lithologic units are laterally continuous and can
312 / Dubois et al.

FIGURE 5. Stratigraphic cross section of the Chase and Council Grove groups with the top of the Council Grove as the datum. At the wells, lumped lithofacies
TM
are those predicted by neural network models (small well symbols) or from core (large symbols) and are interpolated in Geoplus Petra between wells. The
lumped lithofacies include continental sandstone (L0), continental siltstone (L1-2), mud-supported carbonate and marine siltstone (L3-5), grain-supported
carbonate and dolomite (L6-9), and marine sandstone (L10). The Council Grove is thinnest at a midfield (midshelf) position. Log curves are gamma ray (left)
and corrected porosity (right).
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 313

FIGURE 6. (A) The present-day structure of the top of the Wolfcampian reservoir (top of Chase) is mostly a function
of eastward tilt during the Laramide orogeny. Note the shelf margin or area of steepened slope at the southeast margin
of the Hugoton field outline. (B) Three-dimensional view of the same area. The present-day structure on the top of
the Chase and base of Council Grove.

be traced across the shallow shelf to the outcrop in extent, siliciclastic sandstone (Seimers and Ahr, 1990;
eastern Kansas. The main pay zones are 13 thin Caldwell, 1991; Olson et al., 1997; Heyer, 1999; Dubois
(mean thickness 6– 70 ft [2 –21 m]) marine, mainly et al., 2003).
carbonate intervals, deposited during sea level high-
stands. Pay zones are separated by continental, mainly
siltstone (red-bed) intervals (mean thickness 6–25 ft Depositional Model
[2 – 8 m]) deposited during sea level lowstands, when Climate, shelf geometry, and glacially forced sea
most of the shelf was exposed. The siltstones gen- level changes all influenced sediment supply, depo-
erally have poor reservoir quality and vertically iso- sitional patterns, accommodation, and stabilization
lated, or restricted communication among, the 13 pay of both the marine and continental sediments. The
intervals (Seimers and Ahr, 1990; Oberst et al., 1994; Hugoton shelf was near the paleoequator (Figure 4),
Ryan et al., 1994; Olson et al., 1997). The principal and monsoonal climate conditions are likely to have
factor in determining the reservoir storage and flow prevailed at the annual scale (Parrish and Peterson,
capacity (hydrocarbon pore volume and permeabil- 1988). Generally, arid conditions accompanied gla-
ity) of Hugoton reservoir rock is primary deposi- cially induced sea level lowstands with more humid
tional texture. Diagenesis, both early and after burial, conditions and high sea level present during inter-
including leaching of grains and cements and early glacial periods (Rankey, 1997; Soreghan, 2002). Pre-
and late dolomitization, are important factors in en- vailing winds are thought to have been from the
hancing or reducing porosity (Seimers and Ahr, 1990; present-day west during winter and east during sum-
Olson et al., 1997; Luczaj and Goldstein, 2000). How- mer (Parrish and Peterson, 1988). Extremely low re-
ever, the dominant reservoir rocks are marine car- lief enabled rapid migration of the shoreline posi-
bonate with grain-supported textures and, to a lesser tion and changes in shelf hydrodynamics, with only
314 / Dubois et al.

Winfield near the updip margin of the field, although


it is more common to find the situation as they de-
picted elsewhere (sandstone at the base of the Towan-
da and top of the Winfield). Although similar in many
respects, the Council Grove cycles are typically more
asymmetric than the Chase cycles and tend to have
better developed, thin, packstone-grainstone litho-
facies at the base of the marine half-cycle. Figure 10
presents a composite of the vertical distribution of
lithofacies from model node wells that also help il-
lustrate the difference in symmetry between a Coun-
cil Grove and Chase cycle.
Intervals were defined within a simple cyclic frame-
work instead of a sequence-stratigraphic framework.
Existing formation or member tops that are half-
cycle boundaries between marine and continental
intervals represent a sequence boundary and flooding
surface. Because the transgressive systems tract (flood-
ing surface to maximum flooding) is relatively thin
and consistent in most of the cycles, little is gained by
correlation of an additional surface for sequence-
stratigraphic classification.
Idealized depositional models (Figure 11) for the
Council Grove Chase and Chase are generally sim-
ilar, but differences exist because of gradual changes
in climate, ambient sea level position, and sea level
fluctuation rate. Differences may be related to a shift
FIGURE 7. Isopach of the Wolfcampian reservoir (top of from more icehouse to more greenhouse conditions in
Chase Group to base of Grenola Limestone, Council Grove the Permian (Parrish, 1995; Olszewski and Patzkowsky,
Group). Wolfcampian rate of thickening increases by a 2003). The entire Hugoton shelf was above sea level
factor of 10 at the shelf margin. during maximum lowstand for all studied Council
Grove cycles, and continental red-bed siliciclastics
accumulated and were stabilized by vegetation and
minimal absolute changes in sea level (as little as built relief preferentially near the field’s west updip
100 ft [30 m]). These conditions set the stage for margin (Dubois and Goldstein, 2005). Accommoda-
the vertical succession of lithofacies repeated from tion for the carbonate sediments of the overlying
one sedimentary cycle to the next, as well as the re- marine half-cycle was reduced, leading to nondepo-
markable lateral continuity of thin lithofacies units sition, or pinch-outs, of several marine intervals in
in each cycle. the Council Grove at that position.
The cyclical nature of the Council Grove and At the end of each lowstand, a relatively rapid sea
Chase is widely recognized (Seimers and Ahr, 1990; level rise resulted in deposition of a thin (1 –4-ft; 0.3–
Caldwell, 1991; Puckette et al., 1995; Olson et al., 1.2-m) transgressive carbonate-siliciclastic interval
1997). Vertical succession of lithofacies in a shoaling- at the base of each marine half-cycle. Only in the
upward pattern in both the Council Grove and Funston (A1-LM) and Neva (C-LM) cycles are well-
Chase (Figure 9) is a result of depositional environ- developed marine siliciclastics (shaly siltstone) de-
ments changing across the shelf in response to posited during maximum flooding. After maximum
rapid sea level fluctuation. Differences in the style flooding, shallowing, accompanied by conditions that
(symmetry) and pattern (lithofacies) among the Chase fostered increased carbonate production, resulted in
cycles recognized by Olson et al. (1997) are confirmed, a shoaling-upward lithofacies stacking pattern. A fall
for the most part, in our study. Exceptions are that we in absolute sea level caused progradation of broad
do see fine-grained sandstone of marginal-marine facies belts (e.g., carbonate sand shoals), resulting in
origin at both the top and base of the Towanda and laterally extensive lithofacies bodies. With continued
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 315

sea level fall, continental sabkha, coastal-plain, and accompanied a sea level fall and withdrawal: sub-
savannah environments followed the retreating shore- tidal carbonate, tidal-flat carbonate, red siltstone
line and covered the carbonate surface. Evidence for and muddy siltstone with anhydrite (sabkha), and
prolonged direct subaerial exposure and erosion of finally, red siltstone with paleosols (coastal plain or
the carbonate surface is absent in all seven Council savannah).
Grove cycles in the nine cores examined. Instead of Although Chase deposition was similarly influ-
calcretes, microkarst, erosion, or other indicators of enced by absolute sea level, it differs from the Council
prolonged exposure in the upper part of the marine Grove in significant ways. Specifically, during Chase
carbonate, there is a vertical succession of lithofa- lowstand, the lateral extent of subaerial exposure on
cies that suggests continuous sedimentation that the shelf was generally more limited, and in some
continental intervals, tidal-flat siltstone and very
fine-grained sandstone are prevalent, particularly
in positions lower on the shelf. Fine-grained eolian
sandstone present in nearly all continental half-cycles
in the Council Grove is nearly absent in the Chase.
Marine transgressions in the Chase generally extend-
ed farther landward than they did during Council
Grove deposition, with marine sediments extending
beyond the updip margin of the field in all six cycles,
whereas they pinch out in four of the seven Council
Grove cycles studied. During the maximum high-
stand and the subsequent fall in sea level, carbonate
sand shoals of the Chase tend to be coarser grained,
and constituents include bioclasts and, occasionally,
ooids, instead of oncoids and peloids (in the Council
Grove), indicating more open-marine conditions.
Another significant difference between the Chase
and Council Grove is the presence of fine-grained
sandstone deposited in tidal-flat and marginal-marine

FIGURE 8. Formation- and member-level stratigraphy


correlated to the wire-line well log in the A-1 Flower well,
Stevens County, Kansas. Commonly used formation and
member letter-number combinations are shown for the
Council Grove. Twelve of the thirteen marine-continental
(carbonate-siliciclastic) sedimentary cycles that are gas
productive are shown (Grenola Limestone, C_LM is not
logged). Stratigraphic names that include ‘‘limestone’’
are marine half-cycles when combined with an adjacent
continental half-cycle, intervals with stratigraphic names
that include ‘‘shale’’ form a complete cycle. Color-coded
lithofacies are derived from core. Three were deposited
in a continental setting (L0 = sandstone; L1 = coarse
siltstone; and L2 = shaly siltstone) and eight in a marine
environment (L3 = siltstone; L4 = carbonate mudstone;
L5 = wackestone; L6 = very fine-crystalline dolomite; L7 =
packstone; L8 = grainstone and phylloid algal bafflestone;
L9 = fine-medium-crystalline moldic dolomite; and L10 =
sandstone). L0 is absent in this well. Wire-line-log abbre-
viations are caliper (CALI), gamma ray (GR), corrected
porosity (PHI_GM3), photoelectric effect (PEF), density po-
rosity (DPHI), neutron porosity (NPHI), core permeability
(K_MAX), and core porosity (CORE_POR). Logged interval
is 520 ft (160 m).
316 / Dubois et al.

FIGURE 9. Idealized Chase and Council


Grove Groups cycles. Chase cycles are
from Olson et al. (1997), used with
permission from the AAPG, and our
Council Grove cycles are similarly for-
matted. One exception is that we extend
the cycle and approximate sea level
curve through the continental half-cycle
based on earlier work (Dubois and
Goldstein, 2005). Five cycle types dis-
tinguished the basis of lithofacies
stacking pattern and inferred relative
sea level curve.

with normal-marine assemblages


are absent in most of the Council
Grove cycles at or near the west
margin of the field. Both the Chase
and Council Grove cycles exhibit
gradual changes through time that
may be related to third-order cy-
clicity (Boardman et al., 2000) and
the overall shift from icehouse to
greenhouse conditions that began
in late Pennsylvanian and contin-
ued until the end of glacial condi-
tions in the Permian (Parrish, 1995).
Most likely, as a consequence of the
climate change trend, Chase ma-
rine carbonate intervals tend to be
three to five times thicker than their
Council Grove counterparts, at least
in the Crouse through Cottonwood
interval (B1-LM–B5-LM) on the Hu-
goton shelf.

settings at either the top, base, or top and base of all Layered Reservoir and Differential Depletion
cycles above the Fort Riley in the northwest part of The Hugoton and Panoma fields appear to be
the field (Winters et al., 2005). one large reservoir system that may have filled and
In all Chase or Council Grove intervals, the nature changed pressure in stages (Sorenson, 2005). How-
of the lithofacies present in a succession are a func- ever, the six pay zones in the Chase are being de-
tion of the position on the shelf: the farther west pleted by production at different rates, as indicated
and updip, the greater the volume of siliciclastics, by different interval pressure tests and reservoir simu-
whether in the marine or continental setting. Ma- lation (Fetkovich et al., 1994; Oberst et al., 1994; Ryan
rine carbonate tends to be muddier to the west and et al., 1994). Table 1 lists zonal pressures showing
northwest, and grain-supported carbonate tends to differential depletion in both the Chase and Council
be finer toward the west, with the dominant grains Grove. Composite tests covering multiple zones ex-
being hardened pellets (round, very fine grained, hibit pressures that primarily represent the produced
micritic) and peloids (subrounded, fine grained, mi- zone(s) having the lowest pressure and the highest
critic) instead of oncoids, bioclasts, or ooids (found permeability as measured from core and drillstem test
in upper Chase only). Marine environments become (DST) analysis. The more recent of the wells tested
more restricted in a westerly direction, and rocks (2005) showed pressures as low as 19 psi (131 kPa) in
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 317

FIGURE 10. Vertical histograms


showing the average relative
distribution of lithofacies in two
Wolfcampian marine half-cycles
from wells having predicted
lithofacies data (node wells).
Data for the Crouse (B1_LM),
Council Grove Group, are from
1146 wells, and for the Krider,
Chase Group, data are from
1069 wells. Histograms and
probabilities demonstrate the
difference in symmetry in ver-
tical lithofacies distribution be-
tween the Chase and Council
Grove. Probability distributions
were used to condition litho-
facies modeling by sequential
indicator simulation between
node wells. Layer annotations
refer to layering in the half-
cycle respective models (dis-
cussed later). Abbreviations are
fine-grained (Fg), fine crystal-
line (Fxln), and fine to medium
crystalline (F-mxln).

using stochastic methods


and (4) populate model with
lithofacies-specific petro-
physical properties and fluid
saturations. Figure 12 pro-
vides a more comprehensive
overview. The following is a
discussion of each of the steps
for integrating core-defined
lithofacies, core-derived pe-
trophysical properties, wire-
the main pay zones, the Herington and Krider half- line-log response, and prediction of reservoir prop-
cycles. This explains why comingled wellhead shut- erties to characterize the reservoir system at the core
in pressures (32 psi [220 kPa] field average in Kansas) and well scale and at the field scale.
do not provide the necessary data for determination
of overall depletion.
Lithofacies Classification
Because petrophysical properties vary among litho-
STATIC MODEL WORKFLOW facies, fundamental to reservoir characterization and
the construction of a cellular reservoir model is the
Although many details exist, the workflow for population of cells with lithofacies. Determining the
developing the Hugoton geomodel can be described number of lithofacies classes and the criteria for de-
simply in four steps: (1) define interval tops from fining classes involved four standards: (1) maximum
logs, lithofacies in core, petrophysical properties for number of lithofacies recognizable by neural net-
lithofacies, and accurate log correction algorithms works using petrophysical wire-line-log curves and
to obtain a log database; (2) train a neural network other variables; (2) minimum number of lithofacies
and predict lithofacies at node wells; (3) populate a needed to accurately represent lithologic and petro-
3-D cellular model with lithofacies and porosity; physical heterogeneity; (3) maximum distinction
318 / Dubois et al.

FIGURE 11. Idealized depositional models for the Council Grove (A) and Chase (B) showing the distribution of dominant
lithofacies on the Hugoton shelf. Depicted are approximate depositional environments and associated lithofacies for
typical Chase and Council Grove cycles at maximum sea level lowstand and during the falling sea level stage of the
marine highstand.

Table 1. Pressures by zone for two relatively closely spaced wells.*

Group Zone 1994 Science Well 2005 Replacement Well


TM
DST-Sip psi (kPa) Composite psi (kPa) XPT -SIP* psi (kPa)

Chase Group Herington 120 (830) 104 (720) 19 (130)


Krider 88 (610) 104 (720) 21 (145)
104 (720) 30 (210)
Winfield SS 105 (720) 104 (720) 141 (970)
Winfield LS 121 (830) 104 (720) 217 (1500)
Towanda 230 (1590) 104 (720) 165 (1140)
U. Ft. Riley >400 (2750) 104 (720) 192 (1320)
Florence 398 (2740) 104 (720) 265 (1830)
Wreford 372 (2570) 104 (720) 219 (1510)
Council Grove Group A1_LM 400 (2760) 156 (1080) nt
B1_LM 350 (2410) 156 (1080) nt
B2_LM 131 (900) 156 (1080) nt
B3_LM 368 (2540) 156 (1080) 386 (2660)
B4_LM 215 (1480) 156 (1080) nt
B5_LM 160 (1100) 156 (1080) 348 (2400)

*The well drilled in 1994 was a research well (Flower A1, Figure 8), drilled with a foam fluid to limit filtrate invasion and formation damage.
Pressures are 24-h shut-in pressures from drillstem tests. The well drilled in 2005 is located 6 mi (10 km) north of the earlier well, and pressures
were recorded in the open hole by Schlumberger’s XPT TM tool, a repeat formation tester.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 319

FIGURE 12. Workflow for field-scale Hugoton model. Workflow can be divided into three broad tasks: (1) gather and
qualify data; (2) process data to provide basic geomodel input files (Develop/Define/Properties/Algorithms); and
(3) build the geomodel. The figure suggests the process is linear, whereas in reality, there are more feedback loops,
multiple iterations at subtask level, and testing and validation at smaller scales.
320 / Dubois et al.

of core petrophysical properties among classes; and analysis and wire-line-log data (Figure 13). In most
4) the relative contribution of a lithofacies class to cases, selected cores included either the entire Chase
storage and flow. An optimal solution using these (five) or Council Grove (six) interval or covered both
criteria resulted in 11 lithofacies distinguished on intervals (three). Examples of the 11 major lithofacies
the basis of rock type (siliciclastic or carbonate), classes are shown in Figures 14 and 15. Common sub-
texture (Folk, 1954, grain size for siliciclastics; Dun- classes are represented, but the examples illustrated
ham, 1962, classification for carbonates), and princi- do not show the range exhibited by a lithofacies. For
pal pore size (visual estimate). In classifying dolomite example, the marine carbonate packstone, packstone-
rocks, we did not consider depositional texture, but grainstone class (L7), includes rocks having a variety
instead, the present texture and pore size that is pri- of principal grain types and grain size but were de-
marily a function of crystal size and the presence or posited in a variety of environments (e.g., fine-grained
lack of molds of leached carbonate grains. Classes pellets = tidal flat and lagoon; peloid and oncoids =
based on differences in core petrophysical properties restricted shelf and shoals; bioclasts = open shelf and
coincided well with major lithofacies classes of rocks shoals).
and have fairly distinctive wire-line-log response to Relative proportions of the 11 lithofacies in 4250 ft
petrophysical properties, the principal variables used (1295 m) of core described in this study are shown in
for neural network prediction of lithofacies. Although Figure 16. Continental lithofacies comprising fine-
defining more classes might have improved petro- grained sandstones (L0), coarse siltstones (L1), and
physical prediction accuracy, the inability of neural fine or shaly siltstones (L2) represent 42% of the rock
networks to effectively recognize and distinguish finer volume, whereas marine carbonates and marine sili-
lithofacies classes discouraged finer class distinctions ciclastics represents 45 and 13%, respectively. Litho-
(e.g., discriminating between fine-grained packstone facies with the greatest storage and flow capacity
and coarse-grained packstone). in marine rocks include L6 to L10, consistent with
A quantitative, digital lithofacies description sys- the principal reservoir lithofacies defined in previ-
tem (Table 2) was used in describing core at 0.5-ft ous studies (Siemers and Ahr, 1990; Olson et al.,
(0.15-m) intervals. Three of the five factors illustrated 1996; Heyer, 1999). These lithofacies represent 31%
in the tables were sufficient to segregate lithofacies of the rock volume (Figure 16) and include very fine-
classes, although other digits were considered initial- crystalline dolomite (L6), fine-medium-crystalline
ly in the process of determining class boundaries. For dolomite (L9) with grain-moldic porosity, and litho-
each interval, as much as 12 variables were recorded. facies with grain-supported texture, including pack-
Factors recorded in addition to those in Table 2 stone (L7), grainstone (L8), and marine sandstone
included the degree of consolidation and fracturing, (L10). Continental, fine-grained sandstones represent
subsidiary pore size, cement mineralogy, bedding, only 6% of the rock volume, but are important to flow
water depth, faunal assemblage, and color. Classify- and storage in the Council Grove. All 11 lithofacies
ing lithofacies in a digital form facilitated changes in are present in the Chase Group, but continental sand-
classification criteria and correlation of lithofacies stones and very fine-crystalline, sucrosic dolomites
with core and log petrophysical properties involved are less common than in the Council Grove. Coarser
in the iterative process of determining optimal litho- crystalline dolomite with grain-moldic porosity, typi-
facies class boundaries. This digital system is designed cally dolomitized bioclastic or ooid grainstone or pack-
to provide a continuous numerical classification that stone, is absent, and marine sandstone is uncommon
corresponds to the continuum in lithological and petro- in the Council Grove. The very fine-crystalline do-
physical properties. In using this system, instead of a lomite is interpreted to have originally been mud-rich
mnemonic system, error in classifying a given sam- carbonate for the most part.
ple is generally only one class up or down, and there-
fore, the predicted property values are within a class
step up or down from the true value. Once an object Lithofacies Prediction
is numerically classified, mapping to alternate clas- To predict lithofacies using neural network anal-
sification schemes can be performed automatically. ysis, we used a standard, single-hidden-layer neural
Fourteen of approximately one hundred contin- network (Hastie et al., 2001) based on wire-line well
uous cores were selected for lithofacies analysis on logs in 1364 node wells throughout the Hugoton and
the basis of length (longest selected), geographic po- Panoma fields. As illustrated in Figure 17, the input
sition (sampling distribution), and availability of core vector to the neural network included two computed
Table 2. Digital lithofacies description system.*

Variables

1 2 3 4 5

Code Rock Type Dunham/Folk Classification Grain Size Principal Pore Size Argillaceous Content

9 Evaporite Cobble conglomerate vcrs rudite and cobble congl (>64 mm) cavern vmf (>64 mm) Frac-fill 10 – 50%
8 Dolomite Sucrosic/pebble conglomerate m-crs rudite/pebble congl (4 – 64 mm) med – lrg vmf (4 – 64 mm) Frac-fill 5 – 10%
7 Dolomite-limestone Baffle-boundstone/vcrs sandstone fn rudite/vcrs sand (1 – 4 mm) sm vmf (1 – 4 mm) Shale > 90%
6 Dolomite-siliciclastic Grainstone/crs sandstone Arenite/crs sand (500 – 1000 Mm) crs (500 – 1000 Mm) Shale 75 – 90%
5 Limestone Packstone-grainstone/med Ss Arenite/med sand (250 – 500 Mm) med (250 – 500 Mm) Shale 50 – 75%
4 Carbonate-siliciclastic Packstone/fn sandstone Arenite/fn sand (125 – 250 Mm) fn (125 – 250 Mm) Shale 25 – 50%
3 Siliciclastic-carbonate Wackestone-packstone/vfn Ss Arenite/vfn sand (62 – 125 Mm) pin-vf (62 – 125 Mm) Shale 10 – 25%
2 Marine siliciclastic Wackestone/crs siltstone crs lutite/crs silt (31 – 62 Mm) Pinpoint (31 – 62 Mm) Wispy 5 – 10%
1 Continental siliciclastic Mudstone-wackestone/vf-m silt fn-med lutite/vf-m silt (4 – 31 Mm) Microporous (<31 Mm) Trace 1 – 5%
0 Shale Mudstone/shale/clay Clay (<4 Mm) Nonporous Clean < 1%

Digital Description Lithofacies Lithofacies Class

1/>2 0 NM sandstone
1/2 1 NM siltstone
1/0-1 2 NM shaly siltstone
0.2/<3 3 Mar shale and siltstone
3-8/01 4 Mdst/Mdst-Wkst
3-8/2-3 5 Wkst/Wkst-Pkst
6-8/8/<3 6 Vfxln sucrosic (Dol)
3-8/4-5 7 Pkst/Pkst-Grnst
3-8/6-7 8 Grnst/PA Baff
7-8/8/>2 9 F-Mxln sucrosic moldic Dol
2/>2 10 Marine sandstone

*After Dubois et al. (2003). (A) Five-digit classification system used for core descriptions at 0.5-ft (0.15-m) intervals, gathered by visual observation with the aid of binocular microscope. A total of
seven other variables were recorded but were not used in determining lithofacies. (B) Digital code for 11 lithofacies. An example, 13323, is a continental siliciclastic, very fine-grained sandstone (203 –
410 ft; 62 – 125 m), with pinpoint porosity and wispy clay laminations (5 – 10% clay). Abbreviated are nonmarine (NM), marine (Mar), carbonate mudstone (Mdst), wackestone (Wkst), packstone,
(Pkst), grainstone (Grnst), phylloid algal bafflestone (PA Baff), dolomite (Dol), very fine crystalline (Vxln), and fine to medium crystalline (F-Mxln).
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 321
322 / Dubois et al.

geologic variables, a depositional environment indica-


tor (MnM) and a stratigraphic cycle relative position
(RelPos), in addition to the following wire-line-log pa-
rameters: gamma ray, logarithm of deep induction resis-
tivity, average of neutron and density porosity, neutron-
density porosity difference, and photoelectric factor
(PE, where available). No adjustment was made for thin-
bed or boundary effects. For each input vector, the net-
work computed a vector of output values representing
the corresponding lithofacies membership probabil-
ities, and the most probable lithofacies was assigned
for the logged interval. The network was trained based
on associations between core-defined lithofacies and
the log and geologic constraining variables. A com-
parison of core-defined lithofacies, lithofacies mem-
bership probabilities, and predicted discrete lithofa-
cies is shown in Figure 18.
The two geologic variables were derived from a
25-formation (or member) set of tops (Figure 8), which
are also the tops of marine or nonmarine (continen-
tal) half-cycles. Values for the depositional environ-
ment indicator were (1) nonmarine (continental) and
(2) marine and (3) intertidal, for the Herington and
Holmesville, where intertidal environments are domi-
nant. The MnM variable helps to distinguish between
lithofacies with similar petrophysical properties but
developed in different broad depositional environ-
FIGURE 13. Distribution of Hugoton cores (continuous) ments. Values for the stratigraphic cycle relative po-
for which lithofacies were defined at 0.5-ft (0.15-m) sition parameter (RelPos) range linearly with depth
intervals. from 0 at the bottom of each half-cycle interval to 1 at
the top, indicating position in each interval. Includ-
ing this curve allowed the network to encode infor-
mation regarding the fairly regular succession of litho-
facies commonly exhibited within each interval and,

FIGURE 14. Major lithofacies in Chase and Council Grove, lithofacies code 0-5. (A) Continental sandstone (L0): Example:
Blue Rapids (Council Grove, B1_SH), Cross H Cattle 1-6, 2652 ft (808 m). Coarse silt to very fine-grained sandstone,
mostly quartz, massive bedded, adhesive meniscate burrows (S. Hasiotis, 2005, personal communication). Low-relief
migrating eolian system. Digital classification: 13322. (B) Continental coarse siltstone (L1): Example: Stearns (Council
Grove, B4_SH), Newby 2-28R, 2963 ft (903 m). Coarse quartz silt, rhizolith (Rz) and root traces with reduction haloes
(Ho). Savannah, slow accumulation of silt by airfall, stabilized by vegetation and soil processes. Digital classification:
12213. (C) Continental shaly siltstone (L2): Example: Hooser (Council Grove, B3_SH), Newby 2-28R, 2944 ft (897 m).
Fine- to medium-grained quartz silt and clay, caliche (Ca), rhizolith (Rz), and root traces with reduction haloes (Ho).
Coastal plain, slow accumulation of silt by airfall, stabilized by vegetation and soil processes. Digital classification: 11114.
(D) Marine siltstone and shale (L3): Example: Funston (Council Grove, A1_LM), Newby 2-28R, 2872 ft (875 m). Very fine-
grained shaly siltstone. Siliciclastic-dominated shelf during maximum flooding. Plug $ = 4.6%; k = 0.0001 md. Digital
classification: 21104. (E) Mudstone and mudstone-wackestone (L4): Example: Crouse (Council Grove, B1_LM), Alexander
D-2, 2962 ft (903 m). Silty mudstone-wackestone, wispy laminations, and ministylolites (Ms), burrowed in part (Bh),
sparse normal-marine fauna, including fusulinids (Fs). Low-energy shelf at a time close to maximum flooding. Plug $ =
3.1%; k = 0.00239 md. Digital classification: 41113. (F) Wackestone and wackestone-packstone (L5): Example: Fort Riley
(Chase), Flower A-1, 2700 ft (823 m). Slightly dolomitized wackestone, normal-marine faunal assemblage includes
echinoids, brachiopods, bryozoan, and fusulinids. Intercrystalline micropores (blue in thin section) in dolomitized mud
matrix is dominant porosity (core slab and thin section stained with alizarin red). Low-energy normal-marine shelf.
Full-diameter $ = 15.2%; k = 0.413 md. Digital classification: 52111.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 323

thus, transfer some of that character to the sequence of The neural network code has been added to
predicted lithofacies in each well. The two curves were Kipling.xla, an ExcelR add-in for nonparametric
computed for the node wells using Visual Basic code regression and classification (Bohling and Doveton,
in an ExcelR spreadsheet run in a batch-processing 2000). For this study, the network was trained to
routine and exported as log curves in a Log Ascii Stan- match observed associations between logs and litho-
dard (LAS) file format. They were then combined with facies identified in core from a set of key wells shown
the wire-line-log curves to complete the feature vector. in Figure 13 (8 Chase wells with 3952 0.5-ft [0.15-m]
324 / Dubois et al.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 325

tuned to the training data and unable to generalize.


Increasing the damping parameter counteracts pre-
cise tuning but can result in the network developing
smoother representations of the boundaries between
lithofacies.
We used cross-validation to estimate the optimum
values for network size and damping parameter. The
cross-validation was done in two ways: (1) splitting
the entire training data set, regardless of well, into
random subsets, with two thirds of the cases used for
training and the other one third for testing (compari-
son of predicted and actual lithofacies); and (2) tak-
ing out each well in turn, training on the remaining
FIGURE 16. Relative proportions of 11 lithofacies in 4250-ft wells, and testing on the removed well. Training and
(1300-m) core from Chase and Council Grove groups, testing were repeated several times for each param-
Hugoton field. Eleven lithofacies are identified by a code eter combination to account for the random varia-
(L0–L10): L0 = sandstone; L1 = coarse siltstone; L2 = shaly tion between different realizations of the network.
siltstone; L3 = siltstone; L4 = carbonate mudstone; L5 = This computationally intensive cross-validation was
wackestone; L6 = very fine-crystalline dolomite; L7 = pack-
stone; L8 = grainstone and phylloid algal bafflestone;
performed using scripts in the R statistical computing
L9 = fine-medium-crystalline moldic dolomite; and L10 = language (R Development Core Team, 2005) using
sandstone. the nnet function developed by Venables and Ripley
(2002). The scripts computed several measures of cor-
respondence between actual and predicted lithofa-
intervals in the training set, and 10 Council Grove cies, including the average lithofacies misallocation
wells with 4593 0.5-ft [0.15-m] intervals). Fundamen- cost for all intervals in the test well. This value is
tal parameters controlling the network behavior are computed from a misallocation cost matrix that as-
the number of nodes in the hidden layer (network signs a cost for misallocation that is proportional to
size) and a damping or decay parameter. Increasing the distance in lithofacies code units in the lithofacies
the network size allows the network to match the spectrum from actual. For example, calling a pack-
training set more closely, but using too many hidden- stone (L7) a marine siltstone (L3) carries a higher cost
layer nodes leads to the network becoming precisely than confusing it with a grainstone (L8) with similar

FIGURE 15. Major lithofacies in Chase and Council Grove, lithofacies code 6-11. (A) Very fine-crystalline sucrosic
dolomite (L6): Example: Cottonwood (Council Grove, B5_LM), Beatty E-2, 2800 ft (853 m). Finely crystalline, sucrosic
dolomitized mudstone, locally with anhydrite cement and replacement in nodules and along fracture (An). Porosity
(blue in thin section) is microporous (intercrystalline) and pinpoint (molds-Mo). Restricted, protected lagoon. Plug $ =
13.9%; k = 1.37 md. Digital classification: 88120. (B) Packstone and packstone-grainstone (L7): Example: Winfield (Chase),
Flower A-1, 2579 ft (768 m). Medium- to coarse-grained bioclastic-oncoid packstone, patchy anhydrite cement (An).
Most porosity (blue in thin section) is intergranular. Carbonate sand shoal on open shelf. Full-diameter $ = 16.4%;
k = 5.98 md. Digital classification: 54520. (C) Grainstone or phylloid algal bafflestone (L8): Lithofacies have similar core
and wire-line-log properties and were lumped because of their small populations. Example 1: Cottonwood (Council
Grove, B5_LM), Alexander D-2, 3024 ft (922 m). Medium-coarse grained oncoid-peloid grainstone. Well-connected
intergranular porosity is blue in thin section. Carbonate sand shoal on restricted shelf. Full-diameter $ = 18.8%; k =
39.0 md. Digital classification: 56540. Example 2: Cottonwood (Council Grove, B5_LM), Newby 2-28R, 2992 ft (912 m).
Phylloid algal bafflestone. Phylloid algal blade molds (Pm) partially filled with anhydrite cement (An). Matrix is largely
peloid-pellet packstone (Pp). Phylloid algal mound on slightly restricted shelf. Full-diameter $ = 20.6, k = 1141 md. Digital
classification: 57770. (D) Fine to medium crystalline moldic dolomite (L9)-Example: Krider (Chase), Flower A-1, 2516 ft
(767 m). Fine-medium crystalline moldic dolomite. Large molds (Mo), possibly ooids and bioclasts, dominate the well-
connected pore system in a dolomitized medium-coarse grained grainstone. Patchy anhydrite cement (An) occludes
some porosity. Carbonate sand shoal on an open shelf. Full-diameter $ = 22.3%; k = 275 md. Digital classification: 88550.
(E) Marine sandstone (L10): Example: Herington (Chase), Flower A-1, 2485 ft (757 m). Planar (Px) and ripple (Rx) cross-
bedding and vertical burrows (Bv). Tidal flat. Very coarse silt to very fine-grained sandstone, well-sorted, subarkose (83%
of detrital fraction is quartz, by x-ray diffraction), well-connected intergranular porosity (blue), patchy anhydrite cement
(An). Full-diameter $ = 20.8%; k = 48.2 md. Digital classification: 23321.
326 / Dubois et al.

FIGURE 17. Schematic repre-


sentation of single hidden-
layer neural network used to
predict lithofacies from wire-line
logs and geologic constraining
variables. Inputs two geologic
constraining variables (MnM =
depositional environment indi-
cator; RelPos = relative position
in stratigraphic interval) and
includes an array of nuclear and
electrical wire-line-log curves:
gamma ray, (GR); logarithm
of deep induction log (LogILD);
average of neutron and den-
sity porosity (%N + %D)/2);
difference between neutron
and density porosity (%N  %D);
and photoelectric effect (PE).
Outputs are lithofacies occur-
rence probabilities.

petrophysical properties. Absolute accuracy in litho- added two additional models for the Chase below the
facies prediction, although desirable, is unrealistic, Towanda (with and without PE) to better represent
and our goal was to limit error to the nearest litho- the distribution of marine sandstone in that interval
facies class. Plots of median average misallocation cost and region. These six neural network models were then
versus damping parameter and network size (panel applied as appropriate to produce predicted lithofa-
variable) for the well-by-well cross-validation, includ- cies versus depth logs at 0.5-ft (0.15-m) intervals in
ing the photoelectric wire-line log (PE) in the Council the 1364 node wells distributed throughout the field
Grove (Figure 19), illustrate the median computed (Figure 20A). The batch prediction capability of the
over 40 average misallocation costs for each param- Kipling program was used in this case, with logs being
eter combination: five trials per well for each of the six read from LAS files and the predicted lithofacies curves
Council Grove wells with PE logs. Cross-validation being written out to LAS files.
plots for the Council Grove case without PE, and for Following the construction of predicted lithofacies-
the Chase, with and without PE, were similar (not depth curves, the predicted lithofacies curves were read
TM
illustrated). Although the network performed reason- into Petrel (Schlumberger 3-D modeling software)
ably well throughout a range of parameter values, we and upscaled to the resolution of the model layers
chose to use a network size of 20 hidden-layer nodes (roughly 2 ft [0.6 m] in marine intervals and 4 ft [1.3 m]
and a damping parameter of 1.0. The damping pa- in nonmarine intervals) by majority vote: the litho-
rameter chosen (1.0) exhibited consistently low mis- facies for each model cell intersecting a well is taken
allocation values, and the number of hidden-layer to be the most frequently occurring lithofacies in that
nodes (20) was chosen over configurations with fewer layer in the well. Voxel-based methods were chosen
nodes that tended to overgeneralize. over object-based methods for facies and property mod-
Initially, we trained four neural networks: Chase eling because of the relatively dense well control and
with and without PE logs and Council Grove with geometry of the lithofacies bodies being modeled
and without PE logs. Predictions using the model (thin and laterally extensive). Sequential indicator sim-
including PE were used wherever possible. The Chase ulation (Deutsch and Journel, 1998) as implemented
models included all 11 lithofacies, but the Council in Petrel was used to generate lithofacies values in
Grove models included only 9 because the fine- to all model cells conditioned on the upscaled litho-
medium-crystalline dolomite and marine sandstone facies values in the node well cells. The development
do not occur in the Council Grove in sufficient vol- of variograms for the lithofacies by data analysis on a
ume to be considered as separate classes. Later, we zone-by-zone basis (24 zones and 11 lithofacies) is
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 327

FIGURE 18. Comparison of predicted lithofacies versus core-defined lithofacies. Illustrated vertical plots of lithofacies
membership probabilities, predicted discrete lithofacies (most probable), and lithofacies in core at the 0.5-ft (0.15-m)
scale for the Chase and Council Grove from two separate wells in the training set (Youngren and Stuart, respectively).
Neural networks were those used in estimating lithofacies in node wells (trained on all wells). Probabilities were not
used as an input for modeling, but they do illuminate some of the misallocations (actual lithofacies is commonly in
second place).

difficult and not justified considering the well control mate depositional strike, whereas in the continental
and lithofacies geometry. Regions were therefore clas- and tidal-flat half-cycles, a 30,000  30,000-ft (9144 
sified using a limited number of variograms for each 9144-m) range was used. Vertical range was set at ap-
lithofacies and large horizontal ranges. Using large proximately two times the mean layer height in the
horizontal ranges encouraged the stochastic simula- zone, and nuggets ranged from 0.1 to 0.22 on the
tion to produce facies bodies that are as laterally ex- basis of limited data analysis. The very large number
tensive as possible, but still subject to conditioning of conditioning data helped reduce the sensitivity of
data at the node wells. This provided facies bodies the simulated results to the variogram model param-
consistent with geologic depositional models but eters, making the predictions more deterministic.
constrained by node well data. Lithofacies bodies Table 3 shows the proportional distribution of litho-
are laterally extensive because of both the width of facies defined in core in the training set (keystone
the lithofacies spectrum used in determining class wells) and predicted lithofacies at the node wells and
(lumped lithofacies) and the geologic conditions in the lithofacies model in the study area. Consid-
under which they were deposited. Node well spac- ering bias by the geographic distribution of sam-
ing is much closer than the width of the lithofacies pling and the scale change from 0.5 ft (0.15 m) to
bodies. For the marine half-cycles, the major axis was much thicker layering, consistency in the distribution
set at 30,000 ft (9144 m) and minor axis at 25,000 ft at each scale can be characterized as good. Most dis-
(7620 m) and the azimuth equal to 118, the approxi- crepancies are between the core-defined lithofacies
328 / Dubois et al.

FIGURE 19. Example results


for cross-validation analysis to
determine optimal values of
neural network size (number
of hidden-layer nodes) and
damping parameter. Results
shown are the median over
eight Council Grove wells
with core-defined lithofacies
and wire-line photoelectric
curve. The procedure was to
perform five trials per well;
leave out each well in turn,
train on the other seven wells,
and predict on the subject
well. Median average mis-
allocation cost versus damping
parameter and network size
for all wells was then plotted.
A network size of 20 hidden-
layer nodes and a damping
parameter of 1.0 were chosen.

(training wells) and the lithofacies predicted by fluid saturation. Petrophysical properties vary among
neural networks (node wells) and are caused by the the 11 major lithofacies. Principal lithofacies-specific
position of the training wells on the shelf. Training petrophysical properties analyzed and discussed here
wells are skewed to the west (because of core avail- include routine helium and in-situ porosity, routine
ability), whereas the node wells are more evenly air and in-situ Klinkenberg gas permeability, grain
distributed. An exception is grainstone (L8), where density, capillary pressure, and gas-water relative
the neural networks as trained are not effectively permeability.
predicting this facies. Fortunately, this facies repre- Data for routine porosity, permeability, and grain den-
sents only a small part of the volume (2.3% in the sity were compiled for more than 6000 full-diameter
training wells) and is commonly mistaken as pack- core and plug samples from measurements performed
stone, a close neighbor, in the neural network train- by commercial laboratories and the Kansas Geolog-
ing exercises. ical Survey. Lithofacies were determined for more
than 3500 samples. Core-plug sampling was designed
to represent the range in porosity, permeability, and
Core Petrophysics lithofacies in the study area. Routine air permeability
Previous petrophysical studies of the Hugoton and was generally measured under a confining pressure of
Panoma fields have generally used average proper- approximately 400 psi (2.8 MPa); in-situ Klinkenberg
ties assigned to formations (e.g., Siemers and Ahr, gas permeability (high-pressure gas or liquid equiva-
1990; Olson et al., 1996). Porosity and permeability lent) and in-situ porosity were measured under a hy-
characteristics of reservoir-quality wackestone, pack- drostatic confining pressure of 800 psi (5.5 MPa) and
stone, and grainstone lithofacies in the Council Grove, a hydrostatic in-situ stress equal to the net effective
Texas County, Oklahoma, were reported by Heyer stress. A correlation between routine helium and in-
(1999). Byrnes et al. (2001) and Dubois et al. (2003) situ porosity was determined for 245 core samples
presented lithofacies-specific petrophysical proper- representing the range in porosity and lithofacies. In
ties for the Council Grove in Panoma field and il- addition, capillary-pressure curves were obtained for
lustrated the similarities between low-permeability 252 samples using air-mercury capillary-pressure intru-
carbonates and sandstone. Fundamental to construc- sion analysis. Data for 32 gas-water drainage-relative
tion of the reservoir geomodel is the population of permeabilities were compiled, and effective gas per-
cells with the basic lithofacies and their associated meability at critical water saturation was measured
petrophysical properties—porosity, permeability, and on more than 200 cores.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 329

FIGURE 20. (A) Map illustrating the location of 1364 node wells used for the static model construction. Fourteen of the
wells have core-defined lithofacies (circled) and the balance has lithofacies predicted by neural networks. The 1364 is a
mix of wells with Council Grove (1146) and Chase (1069) wells with lithofacies defined by neural net (854 of the 1364 have
both Council Grove and Chase lithofacies). Only wells with lithofacies defined at least to the top of the Fort Riley (Chase)
or the Florena Shale, B5_SH (Council Grove), were considered. (B) Map showing 8765 wells with formation-member level
tops was used for building the structural and stratigraphic framework for the Hugoton geologic model.

Porosity and Grain Density Permeability


Routine (unconfined) helium porosity ($He) values Core-measured in-situ Klinkenberg permeabilities
range from 1 to 26% (Figure 21), with in-situ porosity (kik) range from 0.00002 to 400 md (2  108 –0.4 mm2).
($i) values averaging 0.54 porosity units less than Permeability is a function of several variables, includ-
routine porosity values (i.e., $i = $He  0.54) for all ing primarily pore-throat size, porosity, grain size and
lithofacies and porosity rocks. Comparing porosity packing (which controls pore body size and distribu-
among lithofacies, average porosity decreases with tion), and bedding architecture. Lithofacies-specific
decreasing grain size (i.e., very fine sandstone to very equations were developed to predict permeability using
fine-medium siltstone) in the continental siliciclas- porosity as the independent variable because porosity
tics and from grainstone to mudstone in the carbon- data are the most economic and abundant and be-
ates. Continental rocks exhibit an average grain cause porosity is well correlated with the other vari-
density of 2.70 ± 0.08 g/cm3 (error expresses 2 stan- ables for most lithofacies.
dard deviations), limestones exhibit a grain density More than approximately 75% of all rocks in the
of 2.72 ± 0.06 g/cm3, and dolomites exhibit a grain Hugoton exhibit an in-situ Klinkenberg permeability
density of 2.82 ± 0.08 g/cm3. of less than 1 md (0.001 mm2). Obtaining accurate
330 / Dubois et al.

Table 3. Relative distribution of 11 lithofacies in core, node wells, and cellular model.*
Height Source 0.5 ft (0.15 m) 0.5 ft (0.15 m) Variable** Variable**
Actual (%) NNet Predicted (%) Upscaled (%) Modeled (SIS) (%)

Lithofacies Code Training Node Wells Node Wells All Cells

0 5.6 2.2 1.0 1.1


1 23.3 19.7 17.0 16.7
2 12.9 9.6 7.1 6.7
3 7.5 9.6 9.0 9.1
4 5.4 4.3 3.6 3.9
5 14.5 20.1 22.2 22.5
6 2.8 4.9 3.9 3.8
7 14.7 24.7 25.9 25.2
8 2.3 0.2 0.2 0.2
9 5.6 1.4 3.8 3.8
10 5.4 3.4 6.3 7.1
Count (N) 8545 993,146 183,949 107,765,147

*Core-defined lithofacies for 14 wells were used in neural network Training for lithofacies prediction in 1364 Node Wells. Lithofacies 0.5 ft (0.15 m)
in node wells were upscaled to model layer thickness (Variable Upscaled). Sequential indicator simulation (SIS) was used to populate the cellular
model (All Cells) between the node wells.
**Model layer h. Average of mean h = 3.3 ft (1 m). Range of mean h = 1.9 – 5.2 ft (0.57 – 1.58 m). Lithofacies 0 – 2 tend to be in thicker layers.

permeability values for rocks with low permeability between matrix and well test permeability is discussed
requires correction for the influence of confining in a following section.
stress and care to avoid core data influenced by stress- As with many sedimentary rocks, the relationship
release microfractures. Figure 22 shows that both between permeability and porosity can be approxi-
continental and carbonate rocks in the study area mated by power-law functions, although the rela-
exhibit a significant decrease in permeability below tionship changes slightly in some lithofacies at po-
approximately 1 md (0.001 mm2) because of the in- rosities below approximately 6%. Each lithofacies
fluence of Klinkenberg gas slippage and confining exhibits a relatively unique kik  $i correlation that
stress combined. This decrease is consistent with other can be represented using a power-law equation of the
tight gas rocks (Byrnes et al., 2001). form (Figure 24)
Cores with identified macrofracturing exhibit in-
creasingly greater permeability with decreasing ma- kik ¼ AfBi
trix permeability (Figure 23). The permeability of
the fractured cores with matrix permeability below where kik is in millidarcys (md); porosity is in percent
0.5 md (0.0005 mm2) can be attributed to the core (%); and values for A and B are shown in Table 4.
permeability measurement reflecting the permeabil- Subparallel trends are apparent for the continental
ity of fractures in the sample with the matrix contri- siliciclastics, the sucrosic dolomites, and the mudstone-
bution being small or negligible. Full-diameter anal- to-grainstone limestones. For these trends, the standard
ysis, generally performed at confining pressures less error of prediction ranges from a factor of 3.3–9.1. At
than 400 psi (2.8 MPa), commonly exhibit significant $i > 6%, permeability in a grainstone-bafflestone can
difference from plug values for permeabilities below be 35 times greater than mudstone and 150 times
0.5 md (0.0005 mm2), even for samples where frac- greater than marine siltstone of similar porosity. These
tures were not identified but microfractures may have differences illustrate the importance of identifying
been present. To obtain accurate matrix permeability- lithofacies to correctly predict permeability from wire-
porosity correlations, full-diameter permeability data line-log porosity. Differences in permeabilities be-
for permeabilities less than 0.5 md (0.0005 mm2) that tween continental very fine sandstone and coarse
were more than 2 standard deviations outside the siltstone and between coarse siltstone and very fine-
plug-defined matrix permeability-porosity trend were medium siltstone are approximately 2.5 times,
not used in final matrix correlations. The correlation whereas the difference between marine sandstone
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 331

determination of formation water saturation from


electric wire-line-log response is problematic because
of deep mud filtrate invasion with conventional mud
programs because of the low reservoir pressure (Ol-
son et al., 1997; George et al., 2004). Because water
saturations could not be reliably determined for most
wells using logs, saturations were estimated based on
matrix capillary-pressure properties and determina-
tion of the free-water level (level at which gas-brine
capillary pressure is zero). Air-mercury capillary-pressure
data were compiled and measured for 252 samples
ranging in porosity, permeability, and lithofacies, and
relationships developed that allowed the prediction
of a capillary-pressure curve for any given lithofacies
and porosity.
To examine the lithofacies dependence of thresh-
old-entry pressure, gas-column height, and pore-
throat size (Figure 26), laboratory capillary-pressure
data were converted to reservoir gas-brine capillary-
pressure data using the standard equation (Purcell,
1949; Berg, 1975): Pcres = Pclab (scosures/scosulab); where
Pcres is the gas-brine capillary pressure (psia) at reser-
voir conditions, Pclab is the laboratory-measured cap-
illary pressure (psia), and scosures and scosulab is the
interfacial tension (s, dynes/cm) times the cosine of
the contact angle (u, degrees) at reservoir and labo-
ratory conditions, respectively. For the Hugoton and
Panoma fields, the interfacial tension is approximate-
ly 63–65 dynes/cm for the initial reservoir gas pres-
FIGURE 21. Histogram of routine helium porosity for sures of about 400–450 psi (2.8–3.1 MPa) and tem-
Chase and Council Grove nonmarine continental (NM)
peratures of 90–1008F (32–388C) (Hough et al., 1951;
sandstones and siltstones (A) and limestones (B). Litho-
facies codes and descriptions are provided in text. Porosity Jennings and Newman, 1971). Conversion of capillary
generally increases with increasing grain size in siliciclastics pressure to height above free-water level to determine
and with decreasing mud content from mudstone through the water saturation in any given rock type as a func-
grainstone (Baff = bafflestone; Grst = grainstone; Pkst = tion of height above the free-water level requires the
packstone; Wkst = wackestone; Mdst = mudstone). conversion of capillary-pressure data to height above
free-water level. This conversion was performed using
the standard relation (Hubbert, 1953; Berg, 1975): H =
and marine siltstone is approximately 10 times. Pcres/(C(rbrine  rgas)), where H is the height (ft) above
Although models for permeability prediction using free-water level, Pcres is the capillary pressure (psia) at
porosity were developed for utility, we recognized reservoir conditions, rbrine and rgas are the density of
that the dominant control on permeability is the brine and gas at reservoir conditions (rbrine = 1.16–
pore-throat size (Figure 25). 1.19 g/cm3 and rgas = 0.025–0.035 g/cm3, which are
reasonable intermediate values for these fields, and C
is a constant (0.433 [psia/ft]/[g/cm3]) for converting
Water Saturation and Capillary Pressure density to pressure gradient in psia/ft. From the air-
It is important to consider the presence of water mercury capillary-pressure data, pore-throat diameter
in the pore space of low-permeability reservoirs both was calculated using the modified Washburn (1921)
for accurate volumetric calculations and because relation: d = 4Cscosu/Pc, where Pc = capillary pres-
water occupies critical pore-throat space and can sure (psia); C = 0.145 psiacmmm/dyne; u = contact
greatly diminish gas permeability, even in rocks at angle (1408); s = interfacial tension (484 dynes/cm);
irreducible water saturation (Swi). In the Hugoton, and d = pore-throat diameter (mm). This relation
332 / Dubois et al.

FIGURE 22. Crossplot of routine air


permeability (kair) versus in-situ Klin-
kenberg gas permeability (kik) for Coun-
cil Grove rocks (gray circles) and Chase
rocks (black triangle). Influence of both
confining stress and Klinkenberg cor-
rection increase with decreasing perme-
ability. Values of kik can be predicted
approximately from kair using log10kik =
0.059 (log10kair)3  0.187 (log10kair)2 +
1.154 log10kair  0.159, where perme-
abilities are in millidarcys (md). Vari-
ance is caused by both differing routine
conditions and rock response.

Synthetic capillary-pressure curves


were constructed based on capillary
pressure-porosity-permeability-litho-
facies relationships exhibited by the
252 cores analyzed, representing the
range in lithofacies and permeability.
Capillary pressures in each lithofa-
cies can be represented to be a func-
tion of porosity. Modeled capillary-
pressure curves for two important
lithofacies (Figure 27) illustrate that
with decreasing porosity (and asso-
ciated permeability), threshold entry
heights and transition zone heights
increase. Example capillary-pressure
curves for different lithofacies at a
assumes that the nonwetting phase (i.e., gas) enters given 10% porosity (Figure 28) illustrate the signifi-
the pores through circular pore throats. cant differences in Sw that can exist among lithofa-
Figure 26 illustrates selected capillary-pressure cies at any given height above free-water level. Because
curves for samples of different permeability. Differ- differences decrease with increasing height, satura-
ences among capillary-pressure curves for the various tions for all lithofacies approximately approach a
lithofacies correspond to variations in threshold-entry similar irreducible saturation at gas-column heights
pressure, pore-throat diameter, and water saturation above approximately 300 ft (90 m), except for low-
for various gas-column heights above the free-water porosity rocks where saturation differences are still
level, including the thickness of the transition zone evident. Using the capillary-pressure model, it was
from Sw = 100% to approximately Swi. possible to predict water saturation for any given litho-
Capillary pressures and corresponding water satura- facies and porosity at any given height above free-
tions (Sw) vary among lithofacies and with porosity- water saturation and, thus, populate every grid cell
permeability and gas-column height. Threshold entry in the 3-D geomodel with water saturation values.
pressures and corresponding heights above free-water
level are well correlated with permeability (Figure 26).
This is consistent with the relationship between Relative Permeability
pore-throat size and permeability. The figure shows Gas and water drainage relative permeability curves
that for rocks with in-situ Klinkenberg gas permeabil- reveal several characteristics similar to other low-
ity below approximately 0.003 md (3  106 mm2), permeability rocks. Water permeability, even at 100% Sw,
threshold entry heights are greater than the gas- is less than Klinkenberg gas permeability and decreases
column heights available in the Hugoton, and there- with decreasing permeability. Gas relative perme-
fore, the samples have Sw = 100%. ability is less than the absolute gas permeability at
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 333

FIGURE 23. Crossplot of in-situ Klinkenberg permeability versus in-situ porosity for whole core identified as fractured
(asterisk), whole core that were not identified as fractured but may contain microfractures (gray circle), and unfractured
core plugs (black triangle). Permeabilities shown are either measured in-situ values or routine values corrected to in-situ
conditions using the equation presented in Figure 22. Whole core (full-diameter) values diverge from matrix (plug)
values at porosities less than about 10% and matrix permeability of approximately 0.5 md, reflecting the influence of
microfracture(s) on permeability in whole core samples with porosity less than 10%. Above 10% porosity, influence of
microfracture(s) is small.

all water saturations greater than zero and gas relative showed that for rocks below approximately 8% po-
permeability decreases significantly as Sw increases rosity, or approximately 0.5 md (0.0005 mm2), micro-
above 50%. Relative permeabilities can be reasonably fractures in core significantly increased permeability.
modeled using Corey-type equations (Figure 29), sim- A fundamental question for these data is as follows:
ilar to other low-permeability rocks (Byrnes, 2003). are the microfractures present in the subsurface, or
are they a stress release or coring-induced phenom-
enon? This question can only be answered by com-
Permeability at Different Scales paring upscaled matrix permeabilities with unfrac-
Fundamental to modeling the permeability distri- tured full-diameter permeabilities and with DST- or
bution in the Hugoton is the need to understand the well test-calculated permeabilities. Comparing care-
relative function of matrix and fracture flow and the fully examined unfractured full-diameter permeabil-
possible scale dependence of permeability. Figure 23 ity values with core-plug values measured on plugs
334 / Dubois et al.

Table 4. Power-law coefficients (A) and exponents


(B) by lithofacies for in-situ Klinkenberg (md)
versus in-situ porosity (%) trendlines illustrated
in Figure 24.
Lithofacies In-situ In-situ
Code Klinkenberg Klinkenberg
Permeability Permeability
Coefficient Exponent
kik = A$iB kik = A$iB
A B

0 1.000E – 09 7.90
1 3.715E – 10 7.90
2 1.585E – 10 7.90
3 1.995E – 11 8.31
4 2.088E – 10 7.98
5 2.967E – 08 6.26
6 1.967E – 09 7.10
7 1.527E – 07 6.17
8 3.631E – 09 8.24
9 2.553E – 07 6.30
10 1.995E – 10 8.31

taken from the full-diameter cores (Figure 30) indi-


cates that for homogenous samples, matrix proper-
ties apply to the full-diameter core scale.
The ability to compare well-scale permeability with
matrix permeability is limited because so few wells
have DST or well test data for thin intervals for which
core data are available and which were tested prior
to hydraulic fracturing, which complicates artificial
fracture-enhanced permeability with reservoir perme-
ability. In four key research wells, permeability was
measured using DST for multiple intervals for which
core analysis was also performed. To compare with
core permeabilities, full-diameter and plug perme-
abilities were arithmetically averaged (representing
parallel-flow contribution from each depth interval)
to determine average interval permeabilities. Correla-
tion between DST and upscaled full-diameter and plug
core permeabilities (Figure 31) shows good correlation
FIGURE 24. Crossplot of in-situ Klinkenberg permeability for intervals with permeability greater than about
(kik) versus in-situ porosity ($i) for the continental silici- 0.5 md (0.0005 mm2). For interval permeabilities below
clastics (A), limestones (B), and dolomites (C). Lithofacies
and codes are discussed in the text. Each lithofacies ex-
0.5 md (0.0005 mm2), full-diameter permeabilities
hibits a relatively unique kik  $i correlation that can exhibit nearly constant permeability between 0.5
be represented using a power-law equation of the form and 3 md (0.0005 and 0.000033 mm2), characteris-
kik = A$ Bi ; values for A and B are shown in Table 4. For tic of microfracture-influenced permeability. Matrix-
some samples, routine permeability values were con- scale plug permeabilities can be either higher or lower
verted to in-situ values using the equation in Figure 22,
than DST permeabilities.
and routine porosity was converted to in-situ porosity
using the equation in the text. Trend lines exhibit a stan- Variance in the DST-matrix permeability correla-
dard error of prediction ranging from 3.3 to 9 times tion is partially or predominantly related to the lim-
depending on the lithofacies. ited vertical sampling of the core plugs and difficulty
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 335

FIGURE 25. Crossplot of principal pore-throat diameter (PPTD, micrometers) versus in-situ Klinkenberg permeability
(kik) for lithofacies in Chase and Council Grove groups and sandstones from outside the Hugoton embayment for com-
parison. The good correlation through eight orders of magnitude shows the predominant influence that pore-throat size
exerts on permeability and explains permeability changes with grain size and Dunham classification at a given porosity.
Second Y-axis shows corresponding threshold entry heights necessary for gas column to enter sample for gas pressure
and temperature conditions in the Hugoton area and discussed in text. The correlation between kik and PPTD can be
expressed: PPTD = 2.2 kik0.42. Standard error of prediction for this correlation is a factor of 1.7.

FIGURE 26. Selected capillary-pressure


curves for rocks of different perme-
ability illustrating general curve char-
acteristics. Capillary pressure has been
converted to height above free-water
level (at which Pc = 0) using equations
in text. These curves illustrate how
threshold entry height and transition
zone height increase with decreasing
permeability.
336 / Dubois et al.

ability for rocks with permeability greater than 0.5 md


(0.0005 mm2), both full-diameter and plug data reflect
matrix properties, and the good correlation with DST
permeabilities indicates that the reservoir is not frac-
tured at the scale of investigation of the DST. The
better correlation of plug and DST permeabilities for
intervals with permeability below 0.5 md, and the
fact that upscaled permeabilities from plug data are
greater than or equal to DST permeabilities for three
of four intervals, can be interpreted to indicate that
these intervals are also unfractured. These data, and
less precise data from other wells, indicate that the
production characteristics of many wells in the Hu-
goton are consistent with matrix properties control of
flow, without significant natural fracture contribu-
tion. Data and statistics on the fraction of wells that
exhibit production greater than what would be pre-
dicted from matrix properties have not yet been com-
piled and calculated.

Log Petrophysics
The application and validation of statistical and
petrophysical concepts to porosity estimation is an
important factor in reserve calculations. As always,
accuracy is a major concern, but special attention
must be paid to potential bias introduced by gas ef-
fects and lithology variation. Cumulative effects of
the factors can result in significant under- and over-
estimation of total hydrocarbons in place. As dis-
cussed earlier, the Chase and Council Grove are in-
vaded extensively by mud filtrate during drilling with
conventional mud programs, so that water satura-
FIGURE 27. Model capillary-pressure curves converted tions estimated from resistivity logs are adversely
to height above free-water level for nonmarine coarse overestimated. Pore volumes remain unchanged re-
siltstone (A) and packstone and packstone-grainstone gardless of invasion, so that density and neutron logs
(B) for different porosities. Threshold entry heights and
can be used with confidence for porosity estimation
transition zone heights increase with decreasing porosity
for all lithofacies. The siltstones exhibit both greater entry in the Hugoton. However, these logs must be eval-
heights and higher transition zones than corresponding uated carefully to eliminate, first, borehole environ-
porosity packstones. Threshold entry heights for coarse mental effects, and then the factors associated with
siltstones less than 6% porosity exceed existing closure variable mineralogy and variable gas effects. Because
in the field, indicating these rocks are water saturated.
we used only relatively modern log suites (post-1980)
Curves for packstones less than 6% porosity are not
modeled. and only neutron and density porosity for estimating
porosity, there was no need to calibrate, normalize, or
otherwise preprocess the logs other than checking
in representing some pore properties that are larger quality.
in scale than core plugs. The single, phylloid algal Because the density and neutron measurements
bafflestone interval exhibits significantly lower ma- are from devices that make contact with the borehole
trix permeability because core plugs did not sample wall, the highly anomalous porosities caused by wash-
the larger scale vuggy nature of this lithofacies, which outs must be removed prior to detailed analysis. Po-
exhibits high permeability. Because microfractures rosity cutoffs of 20–22.5% were applied by lithofacies
do not contribute significantly to measured perme- to eliminate washout effects. This was followed by a
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 337

FIGURE 28. Model capillary-pressure curves


converted to height above free-water level
for lithofacies L0-L10 at a constant poros-
ity of 10%. Figure 27 illustrates how curves
change as a function of porosity for two
lithofacies. In this figure, for a given po-
rosity, comparison between curves at any
given height above free water shows that
water saturations are generally greater
for the siliciclastics than the carbonates.
The differences among curves illustrate
the importance of knowing lithofacies to
accurately predict water saturation.

log alone, because the neutron log


appears to be adversely and erratically
affected by clay minerals in the fine
fraction. Siliciclastic lithofacies were cal-
ibrated separately, and the relation be-
tween each one’s density porosity ($d)
to core porosity is shown in Figure 32. A
striking distinction is present between
the continental lithofacies (L1 and L2)
and the marine siliciclastic lithofacies
(L3), which exhibits relative homoge-
remedial elimination of shoulders caused by the trun- neity. It should be noted here that the carbonate mud-
cation. Porosities substituted into these washout in- stone lithofacies (L4) typically has high silt content,
tervals corresponded to average porosities of the litho- and its population can be characterized as exhibiting
facies assigned to the interval. This procedure was properties more like a fine-grained siliciclastic than a
found to be effective in removing washout effects and pure carbonate mudstone. By contrast, in the carbon-
was performed using an automated process with lim- ate lithofacies L5 to L9, the average of the neutron ($n)
ited manual intervention to preserve intervals with and density ($d) porosities, calibrated to a limestone
high, but real, porosities such as those that occur in matrix, was found to be the best estimator of porosity,
more coarsely crystalline dolomites (L9). Following en- compensating for changes in lithology between lime-
vironmental correction, the calculation of porosities stone and dolomite, as well as minor saturations of
from logs could then be developed as a reliable pro- residual gas.
cedure, particularly when validated through the use
of porosities from the extensive core database as the Compensation for Gas Effects
calibration standard. Although gas effects are minor in the Council Grove
and can be compensated for by the averaging of neu-
Porosity Log Calibration among Lithofacies tron and density porosities, they can be significant in
The subdivision of the Chase and Council Grove into the Chase (Figure 33) and must be accommodated in
lithofacies allowed a statistical strategy of porosity-log an expanded equation set. The equation commonly
calibration that was sensitive to mineralogy effects applied to estimate the porosity with compensation
while simultaneously accommodating the effects of for gas takes the form $ = (($n2 + $d2)/2)0.5. The equa-
gas saturation in residual or greater volumes. Initial tion closely approximates the formula derived by
regression analysis that related core porosities to log Gaymard and Poupon (1968) from a petrophysical
measurements in the Council Grove showed a distinc- model of a gas-filled reservoir. An alternative and em-
tive difference in calibration model between the sili- pirical equation that is also widely used is a simple
ciclastic lithofacies and the carbonate mudstone litho- weighted average of the neutron and density porosi-
facies (L0–L4) and the carbonate lithofacies (L5–L9). ties with a weighting of one third applied to the neu-
The best correlation of log and core porosities in the tron porosity and a two thirds weighting of the den-
siliciclastic lithofacies was exhibited for the density sity porosity (Asquith and Krygowski, 2004). This
338 / Dubois et al.

rosities related to neutron and density fractional log


porosities resulted in an equation of the form $ =
0.37$n + 0.62$d for limestone (lithofacies L5, L7, and
L8) and $ = 0.04 + 0.31$n + 0.53$d for dolomite (L6
and L9). The extra term in the dolomite equation
accommodates the lithology correction required for
logs calibrated to a limestone matrix. Porosities in the

FIGURE 29. Relative gas permeability (A) and water rela-


tive permeability (B) curves for 32 samples of various
lithofacies. Black curves represent predicted values for
the Corey-type equation model: krg = (1  (Sw  Swc,g)/(1 
Sgc  Swc,g))p (1  ((Sw  Swc,g)/(1  Swc,g))2), krw = ((Sw  Swc)/
(1  Swc))q (kw/kik), where Swc is the critical water satu-
ration, Swc,g is the critical water saturation for gas flow,
and all saturation terms are in fractions. Black curves rep-
resent mean values of exponents p = 1.3 for gas curve and
q = 8.3 for water curve, whereas gray-bounding curves
represent outer limits of curves using exponents p = 1.3 ±
0.4 and q = 8.3 ± 3, which represent the range exhibited FIGURE 30. Crossplot of full-diameter core porosity versus
by the sample set, which had 0.1 < kik < 50 md. plug porosity (A) and permeability (B) for samples in which
the full-diameter cores did not exhibit any apparent micro-
fracturing. Good correlation indicates that matrix-scale
properties apply to full-diameter scale. Variance can be
empirical equation closely emulates the gas correc- attributed to full-diameter core sampling multiple litho-
tion shown on service company neutron-density cross- facies or a range in porosity not sampled by the corre-
plots (Figure 34). A regression analysis of core po- sponding core plug.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 339

FIGURE 31. Crossplot of calculated


interval drillstem test (DST) formation
permeability versus average interval per-
meability calculated from full-diameter
core for four wells and from core plugs
in well 1. Routine core data were cor-
rected for confining stress, Klinkenberg,
and relative permeability effects so as
to correspond to reservoir-condition
values. Good correlation down to about
0.5 md shows matrix-scale control of
flow in the region of DST investigation.
Below 0.5 md, microfractures in full-
diameter core result in permeabilities
higher than in the unfractured reservoir.
Higher DST than core-plug permeabil-
ities can be interpreted to indicate that
formation is not fractured in the range
of investigation, and that plug sam-
pling density was probably not ade-
quate to properly sample lower range
of permeability.

FIGURE 32. Porosity calibra-


tion of siliciclastic zones based
on regression of core mea-
surements on density-log po-
rosity in limestone-equivalent
units for continental facies
(L1 and L2) and marine facies
(L3 and L4).
340 / Dubois et al.

FIGURE 33. Example of strong


(and atypical) gas effect on the
neutron and density-porosity
logs in the Towanda Limestone
for a well in Stevens County,
Kansas.

marine sandstone (F10) were


calibrated to log porosities
with the equation $ = 0.05 +
0.15$n + 0.47$d. Although
there are comparatively few-
er core measurements in the
continental sandstone (F0),
the use of the averaged neu-
tron and density porosities
provided an excellent match
with core porosities.

Free-water Level and


Trapping Mechanism
Estimating the free-water
level (FWL) position is crit-
ical for calculating water
saturations using capillary
pressures and the height
above FWL. It has been rec-
ognized that the Hugoton
field has a sloped gas-water
contact, and we interpret a
sloped FWL that is several
hundreds of feet higher at
the west updip margin than
on the east downdip limits
(Garlough and Taylor, 1941;
Hubbert, 1953, 1967; Pippin,
1970; Sorenson, 2005). In
this study, we have defined
the gas-water contact as the
lowest position in the res-
ervoir that a well can pro-
duce gas economically, with-
out substantial water, and
the free-water level as the

FIGURE 34. Neutron-density


porosity log crossplot of gas
zones in the Towanda Lime-
stone from the example Stevens
County well (see Figure 33).
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 341

datum where gas-brine capillary pressure is zero. As saturations using capillary pressure. Although others
shown in the section on petrophysics, initial reser- have presented general descriptions of the gas-water
voir desaturation may not occur for some lithofacies contact datum (e.g., Garlough and Taylor, 1941; Parh-
until several tens or hundreds of feet above the free- man and Campbell, 1993), it has not been rigorously
water level (threshold entry height). For typical res- defined by earlier workers. Our estimation of the FWL
ervoir rocks in the study area, packstone-grainstone (Figure 35) was derived using a combination of four
8–10% porosity, the FWL ranges from 50 to 70 ft (9 to indicators: (1) base of lowest perforations; (2) forma-
21 m) below the gas-water contact, a point at which the tion fluid resistivity estimated from wire-line logs;
water saturation is approximately 70% (see Figure 27). (3) calculation of the FWL for an estimated original
The Hugoton gas reservoir is a dry-gas, pressure- gas in place (OGIP); and (4) pressure measurements
depletion reservoir with very little or no support from of deep-water productive intervals. Within the limits
the underlying aquifer. Vertical water flow is con- of the Panoma field, we based the depth of FWL on
strained by low vertical water permeabilities through the average lowest reported productive perforations
low-porosity siltstone layers (k < 106 md [109 mm2] in the Council Grove (FWL = base of perforations +
for $ < 4%) and by low water relative permeability in 70 ft [20 m]), assuming that operators have been
carbonates with low water saturation. However, below efficient at identifying pay and avoiding water pro-
the transition zone, water can be produced freely, duction. Along the eastern and western margin of
and reservoir pressures (600 – 700 psi; 4.1 – 4.8 MPa) the Hugoton in Kansas, where there is no underlying
approach regional hydrodynamic pressures for the Council Grove production, and the western margin
depth (Sorenson, 2005). The low reservoir gas pres- in Oklahoma, we used a formation fluid resistivity
sures (450 psi; 3.1 MPa) and subhydrostatic water method for estimating the FWL at the field bound-
pressures below the transition zone were proposed by ary. Here, FWL was estimated to be 30 ft (9 m) below
Sorenson (2005) to be the result of water pressure the structural datum of the point in the Chase pay
equilibrating with reservoir rocks exposed at outcrop zones where the average apparent formation fluid
in eastern Kansas and gas cap expansion, and conse- resistivity (mix of all fluids) estimated from logs was
quent pressure decrease. at or near the resistivity of formation water. Limited
The Hugoton has long been considered a classic data in the Oklahoma Panhandle required that FWL
example of a giant stratigraphic trap (Garlough and be estimated by backcalculating the FWL necessary
Taylor, 1941; Parhman and Campbell, 1993) because to calculate an OGIP that would allow total gas pro-
of updip changes in lithofacies and petrophysical duction to equal 70% of OGIP. This method assumed
properties associated with these changes. However, that the Panhandle reservoir exhibited similar pressure
dips on the apparent gas-water contact and FWL that depletion and gas production as reservoirs in Kansas.
cross stratigraphic boundaries cannot be fully ex- The FWL subsea depth is approximately 30 ft
plained by lateral heterogeneities. Hubbert (1953, (9 m) at the east margin of the Hugoton and, mov-
1967) proposed a conceptual model for the Hugoton ing west, remains relatively flat to the east margin
being a hydrodynamic trap with trapping resulting of the Panoma, where it begins to rise at a rate of
from a hydraulic gradient coupled with permeabil- 15 ft/mi (2.85 m/km) to a datum of +250 ft (+80 m),
ity changes at the updip margin of the field. Pippin where it then rises at 50 ft/mi (9.4 m/km) to a height
(1970) cited Hubbert’s hydrodynamics and updip of +1000 ft (+300 m) at the western margin of the
pinch-outs of reservoir rock as the trapping mecha- Hugoton. The configuration closely parallels the gas-
nism. Olson et al. (1997) suggested that sealing faults, water contact described by Parhman and Campbell
at least in the western part of the field in Stanton (1993), although our estimate places it 100 ft (30 m)
and Morton counties, Kansas, compartmentalize the lower at the east margin. Our estimated gas-water
lower Chase reservoirs with the compartments hav- contact is +40 ft (12 m) at this position in the field
ing dramatically different gas-water contacts that rise (FWL + 70 ft [21 m]).
to the west. Sorenson (2005) suggested that the down-
dip flow of gas during expansion of the Hugoton gas Static Model Construction
bubble might be responsible for the gas-water contact The extremely large area (10,042 mi2; 26,008 km2),
geometry. small XY cells (660  660 ft; 201  201 m) and rela-
Determining the mechanism for an uneven FWL tively thin layers (169 layers, 3.3 ft [1 m] thick average;
was not an objective of our investigations, but FWL Figure 36) resulted in a 108-million-cell model that
had to be established for the calculation of water required subdividing the model into parts because of
342 / Dubois et al.

FIGURE 35. Free-water level (FWL) estimated for the Hugoton model area (A, B) is applicable for all production from
the Chase but only for Council Grove production that is inside the Panoma. The entire Wolfcamp (Chase and Council
Grove) is thought to have a common FWL across much of the area, but the Council Grove production outside the Panoma
(in Oklahoma) may be related to a different FWL. The FWL is at near sea level at the east Hugoton margin, climbs gradually
in a westerly direction to midfield, and then ascends more rapidly to a height of more than 1000 ft (300 m) above sea
level at the west field margin. The zero datum of the height above FWL surface for the Chase (C) and Council Grove (D)
correspond well to the edges of the Hugoton and Panoma field boundaries, respectively.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 343

FIGURE 36. Intersecting structural cross


sections (near midfield) in the Hugoton
model illustrate the 169 layers in the
24 zones from the combined six 3-D
grids. Marine zones are more finely
layered than the continental intervals
(vertical exaggeration is 200).

permeability were calculated at the


cell level using the petrophysical equa-
tions presented above. Water satu-
ration was calculated at the cell level
using capillary pressure curves devel-
oped for each cell based on the cell
lithofacies, porosity, and height above
free-water level. Finally, OGIP was
calculated for a bottom-hole pressure
equal to 450 psi (3.1 MPa) and com-
pressibility index (Z) equal to 0.92.

RESULTS AND DISCUSSION

The accuracy and utility of the


Hugoton geomodel can be measured
by several metrics, including pre-
computational limitations. Because the main pay diction accuracy of parameters like lithofacies, petro-
intervals are separated stratigraphically by intervals physical properties, and OGIP at the well to field
with relatively poor reservoir properties and lithofa- scale. The only direct measure for lithofacies is the
cies and property modeling is restricted by zones in comparison of predicted and core-defined lithofa-
the modeling process, we chose to subdivide the cies (Table 3). We can also qualitatively measure the
Chase and Council Grove model stratigraphically validity of the lithofacies model by (1) comparing
instead of geographically, with three multizone, multi- it with earlier work at smaller scales and (2) com-
formation models in the Chase and three in the paring the 3-D lithofacies patterns with deposition-
Council Grove. Each of the six models were built al models that have been proposed for the area and
with the same starting architecture and layering to for upper Paleozoic cyclic depositional systems in gen-
facilitate cutting out selected parts of the six models eral. Measures of accuracy for any parameter (e.g.,
and assembly into smaller models. These would have permeability) at the lease and field-scale are con-
a complete vertical section of the reservoir system, strained by lack of data directly measured at these
but limited aerial extent, and be more easily man- scales and the consequent need to compare other pa-
aged for further analysis and reservoir simulation. rameters, such as OGIP. OGIP requires integration of
The structural framework for the six models was many variables, which may be inaccurate due to er-
based on a structural tops database for 8765 wells ror in a single input variable or improper upscaling
(Figure 20B). Our data set of tops in Oklahoma is less and integration of accurate, but different scale, vari-
complete than it is in Kansas. Layering within the ables. Ultimately, measures of accuracy and utility of
models used the following hierarchy: (1) division a geomodel at the lease and field scales are com-
between formation and members, (2) further subdi- monly defined by comparison of predicted and mea-
vision between continental and marine intervals, and sured production and pressure history, where the
(3) further subdivision into layers based on min- predicted pressure and production data are obtained
imum vertical thickness of the key lithofacies in the from input of the static geomodel (the focus of this
node wells (Figure 20A). Layers averaged 2 ft (0.6 m) chapter) into a reservoir-flow simulator to obtain a
thick in the marine intervals and 4 ft (1.2 m) thick in dynamic model. The workflow involved in calibration
the continental intervals. Horizontal and vertical of the geomodel with dynamic data is not discussed in
344 / Dubois et al.

this chapter or shown in Figure 12, but is an important In cross section at the intercycle scale, backstepping
part of testing and refining a geomodel. Dynamic of major lithofacies associated with changes in water
modeling has been performed of 9- and 12-mi2 (23- depth and energy are evident in the Chase marine
and 31-km2) areas, including histories for 28 and 37 carbonate, particularly in the two dolomite litho-
wells, respectively, to test and refine the geomodel facies. The position on the shelf where continental
discussed here. These simulations are ongoing and are siliciclastics thicken at a higher rate also backsteps in
both generally validating the present geomodel and a similar manner. Marine carbonates in the middle
showing how uncertainty in some properties (e.g., three of the seven Council Grove cycles (Figure 37C)
free-water level) must be reduced to provide a model pinch out at the western field margin. The gradual
that is sufficiently constrained for use in accurately shift in the paleoshoreline position, first to lower on
predicting field performance. the shelf and then back to a higher position, is be-
lieved to be related to a change in sea level amplitude.
The symmetry demonstrated by sedimentation pat-
Model Lithofacies and Properties terns around the middle cycle of the seven fourth-
Lithofacies and property distribution in the 3-D order Council Grove cycles suggests the possibility of
Hugoton cellular geomodel presented here on a larger higher order cyclicity in the latter part of the middle
scale and with finer resolution are consistent with Paleozoic icehouse.
earlier work on the Hugoton (Garlough and Taylor, One of the more striking aspects of the model
1941; Hubbert, 1953; Pippin, 1970; Seimers and Ahr, from a reservoir perspective is the demonstration of
1990; Parham and Campbell, 1993; Fetkovitch et al., lateral continuity in the lithofacies illustrated in
1994; Oberst et al., 1994; Olson et al., 1997; Heyer; 1999; Figure 38. County-scale connected volumes of the
Sorenson, 2005). The full-field geomodel presented more significant reservoir lithofacies are subparal-
here reveals facies and property patterns that could lel to depositional strike and the field margin. Grain-
not be identified at smaller scales. Figures 37 – 39 are supported packstones in the Crouse (Figure 38A, B1_LM,
a series of cross sections and map views of lithofa- Council Grove) are found primarily in the eastern
cies and properties for the six individual models. half of the field, whereas muddier lithofacies are
General trends in thickness and lithofacies distri- dominant in the western, more sheltered shoreward
bution are evident in the 3-D volume: continental part of the shelf. Continental sandstone (eolian) is
rocks are thickest, marine carbonate intervals thin or limited to the northwestern updip margin in the
pinch out at Hugoton’s western updip margin, and Council Grove (Figure 38B), and shallow-marine
the relationship is nearly reciprocal basinward. The and tidal-flat sandstone in the upper Chase is most
important reservoir lithofacies (grain-supported car- abundant in the northwestern half of the field area
bonate, dolomite, and marine and continental sand- (Figure 38C).
stone) are laterally extensive, and the marine carbon- The core to model lithofacies workflow was suffi-
ates, the primary pay zones, are separated by laterally ciently robust to characterize smaller important litho-
continuous continental siltstone with poor vertical facies bodies in the reservoir system (Figure 38D –F).
transmissibility. For example, dolomitized grainstone and packstone
Large-scale sedimentation patterns, interpreted of a relatively thick (30 ft, 10 m) carbonate sand shoal
from lithofacies distribution, are striking when viewed system in the southern half of the field is known
at the scale made possible by the full field-scale model. to be an important contributor to storage and flow

FIGURE 37. Lithofacies in stratigraphic cross sections across the Hugoton shelf (A) for the Chase (B) and Council Grove (C).
Cross sections are 10 – 158 from being dip sections and are hung on the top of the Chase (B) and the Council Grove (C).
Some key observations can be made: (1) In both the Chase and Council Grove, continental half-cycles (yellow-orange
to red lithofacies) are thickest at the west field margin and thin basinward (southeasterly). The pattern for the marine
half-cycles is the opposite and somewhat reciprocal relationship with the continental half-cycles. (2) Backstepping pattern
in lithofacies distribution from one marine cycle to the next in the Chase. (3) Three Council Grove marine half-cycles
pinch out near the west field margin, marking a paleoshoreline that appears to then move northwesterly (landward)
upsection. (4) Trend in carbonate rock texture from mud dominated (landward) to grain dominated (basinward),
especially in the Council Grove. Large-scale sedimentation patterns and distribution of resultant lithofacies (at the cycle
scale) is largely a function of the position on the shelf and reflect the interaction of shelf geometry, sea level, and possibly,
the proximity to siliciclastic sources. Lithofacies distribution and cycle stacking patterns at larger scales may be a function
of lower order cyclicity and a shift from icehouse to greenhouse conditions (upward) during the Lower Permian.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 345

capacity in the Krider (R. Sorenson, 2004, personal Distribution of lithofacies-dependent properties
communication). Modeling of this important lithofa- (porosity, permeability, and water saturation) corre-
cies shows the lateral continuity of a 4  30-mi (6.5  lates well with lithofacies distribution (Figure 39).
50-km) sweet spot (porosity > 18%; Figure 38F). The best porosity and permeability coincide with
346 / Dubois et al.

FIGURE 38. Maps showing lithofacies distribution of selected lithofacies. Illustrated are map views of connected volumes
TM
generated in Schlumberger’s Petrel modeling application. Connected volumes are collections of touching cells in
the cellular model having common properties and, for this model, help to demonstrate the remarkable lateral continuity
of flow units in the Wolfcamp. (A) Thirty largest connected volumes in the Crouse limestone (B1_LM) packstone,
packstone-grainstone (L7, blue), and fine-crystalline dolomite (L6, pink) having porosity greater than 8%. (B) Fifteen
largest connected volumes in the Speiser shale (A1_SH) continental sandstone (L0) with porosity greater than 12%.
(C) Twenty largest connected volumes of marine sandstone (L10) having porosity greater than 15%. (D) Top 20 con-
nected volumes for Krider packstone, packstone-grainstone (L7, blue) and coarse-crystalline dolomite (L9, purple) having
porosity greater than 16%. (E) Enlarged area of (D). (F) Same as (E) except for porosity greater than 18%. Grant and
Stevens counties are outlined in green for reference in (D), (E), and (F).

the main reservoir lithofacies (Figure 37). Laterally west where the free-water level is thought to rise
extensive low-permeability intervals separate the rela- more quickly than the rate of dip.
tively high-permeability pay zones of the layered res- Quantitative measures of lithofacies proportions
ervoir system. Water saturations are high in the con- (Table 3) in core, neural network-predicted lithofa-
fining intervals, and the gas-water contact crosses cies at node wells, upscaled at the node wells and in
stratigraphic boundaries at the east downdip margin the 108-million-cell model, show consistency at the
as the pay intervals dip below the surface and on the different scales, suggesting that the sample rate for
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 347

FIGURE 39. Property distribution in cellular Hugoton model in cross section. (A) Location of cross sections. (B) Chase
Group stratigraphic cross section (datum is top of Chase). Horizontal permeability is shown in west – east section, and
porosity (0 – 30%, yellow is 22%) is in the north – south section. (C) Chase Group through Easly Creek shale (B2_SH,
Council Grove) water saturation. Free-water level is the base of the cross section on the west and east side and the base
of the Easly Creek (B2_SH) in the middle, where the FWL is lower in the stratigraphic column (not able to display all
models simultaneously). Free-water level crosses stratigraphic boundaries in both updip and downdip positions. Highest
permeability (Kxy) and porosity (Phi), and lowest water saturation (Sw) is found in marine carbonates and sandstones.
Continental siltstones separating the marine carbonates are the intervals with Kxy and Phi and higher Sw.

training and lithofacies prediction in each of the are believed to have enabled the success of the neu-
three steps was sufficient. Slight variations in mea- ral networks to predict lithofacies at the node wells
sures are likely to be related to sample distribution and the modeling of lithofacies between nodes. Litho-
in core versus the node wells and the increase in facies classes were chosen to maximize differences
scales from 0.5 ft (0.15 m) to considerably thicker in the signature of wire-line-log variables. Geologic
cells (layers). Several factors related to the nature, constraining variables (e.g., relative position curve
geometry, and distribution of the predicted litho- and marine-nonmarine curve) captured and leveraged
facies and the variables chosen to predict lithofacies geologic information such as the predictable vertical
348 / Dubois et al.

succession of lithofacies in the sedimentary cycles and permeabilities and upscaled plug-scale permeabili-
the primary depositional environment. Extensive lat- ties is interpreted to indicate that production from
eral continuity and lithofacies-body sizes being much many wells is controlled by matrix permeability and
larger than the lateral spacing of node wells helped not fractures.
model the lithofacies between nodes. Significantly, Comparison of petrophysical properties among
there was adequate core control to appropriately sam- lithofacies indicates that permeability increases and
ple the reservoir system for the lithofacies training threshold entry height and transition zone thickness
set. Without core, the model could not have been built. decrease with increasing energy in the depositional
The most common misallocations of lithofacies environment and corresponding decrease in mud and
were prediction of grainstone (L8) as wackestone or silt matrix. Within both the continental and marine
packstone (L5 and L7), predictions of carbonate mud- siliciclastics, permeability increases with increasing
stone (L4) as wackestone (L5), and predictions of the mean grain size at any given porosity. The marine,
continental sandstone (L0) as coarse siltstone (L1). very fine-grained sandstone (L10) exhibits an order
These incorrect lithofacies predictions reflect over- of magnitude greater permeability than marine siltstone
laps of the log characteristics of the lithofacies. For- (L3). However, continental, very fine-grained sand-
tunately, the lithofacies involved are also sufficiently stone (L0) exhibits only approximately 2.5 greater
similar in their petrophysical properties that the re- permeability at any given porosity than continental
sulting distributions of porosity, permeability, and coarse-grained siltstone (L1) and 5 greater perme-
water saturation resulting from random misallo- ability than fine–medium-grained siltstone (L2). This
cations were not significantly different from correct difference is likely because of poorer sorting of the
property distributions. For example, a grainstone (L8) continental sandstone compared to the marine sand-
having 10% porosity and at a position 300 ft (90 m) stone. The poor sorting of the continental siliciclas-
above FWL misclassified as a packstone (L7) would be tics would be expected to result in a wide pore-throat
assigned a permeability of 0.35 md instead of 0.70 md size distribution, which is reflected in the high capillary-
(Figure 24) and a water saturation of 24% instead of pressure transition zones of these rocks.
17% (Figure 28). Limestone permeabilities exhibit subparallel perme-
The majority vote upscaling of lithofacies from ability-porosity trends among lithofacies (Figure 24)
the 0.5-ft (0.15-m) sampling interval to the thick- and increasing mean porosity (Figure 21), upper po-
ness of the model layers intersecting each well did rosity range, and permeability from mudstone through
not significantly alter the lithofacies populations. packstone-grainstone. These changes result in a corre-
Using this methodology, we would not expect to see sponding increase in threshold entry heights and wa-
a significant difference in lithofacies populations ter saturations at any given height above free water.
unless certain lithofacies typically occurred in thin The differences in the capillary-pressure properties
bodies separated by fairly large vertical intervals, re- between lithofacies decrease as porosity and perme-
sulting in a systematic underrepresentation of those ability increase. Both threshold entry heights and
lithofacies in the upscaled results. Finally, the litho- transition zone heights for the limestone are less
facies proportions in the full 3-D model closely re- than for continental siliciclastics of similar porosity.
flect those for the upscaled well cells. This is not
surprising because the stochastic indicator simula-
tion attempts to match the proportions observed in Original Gas in Place
the conditioning data. The distribution of our node Comparison of estimated OGIP by others, production
wells was sufficiently uniform such that the litho- data, and OGIP from calculations in the 108-million-
facies proportions in the conditioning set were very cell model suggests that the geomodel successfully
similar to those in the reservoir system. models the Hugoton, particularly in the center of the
field where control on the free-water level is greatest.
The geomodel OGIP is calculated to be 21.8 tcf
Controls on Petrophysical Properties (0.62 trillion m3) and have a hydrocarbon pore vol-
The petrophysical analysis indicates that accu- ume of 676 bcf (19.1 billion m3, 120 billion bbl) in Grant
rate reservoir properties prediction requires input of and Stevens counties, Kansas, for the Chase (Hugo-
lithofacies, use of properties that represent reservoir ton) and Council Grove (Panoma) intervals (Table 5;
conditions, and filtering of full-diameter data to avoid Figure 40). Cumulative gas production for the area is
microfractured core. The close correspondence of DST 14.1 tcf (0.4 trillion m3) or 65% of calculated OGIP,
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 349

Table 5. Estimated original gas in place (OGIP).*


Group Zone OGIP (bcf) OGIP (109 m3) HCPV (bcf) HCPV (109 m3)

Chase Herington 1227 34.8 38.1 1.08


Krider 2795 79.1 86.7 2.46
Odell 295 8.4 9.2 0.26
Winfield 3215 91.0 99.8 2.83
Gage 807 22.9 25.0 0.71
Towanda 4686 132.7 145.5 4.12
Holmesville 663 18.8 20.6 0.58
Fort Riley 5212 147.6 161.8 4.58
Matfield 127 3.6 3.9 0.11
Wreford 1048 29.7 32.5 0.92
Council Grove A1_SH 331 9.4 10.3 0.29
A1_LM 656 18.6 20.4 0.58
B1_SH 76 2.2 2.4 0.07
B1_LM 143 4.1 4.4 0.13
B2_SH 9 0.3 0.3 0.01
B2_LM 167 4.7 5.2 0.15
B3_SH 56 1.6 1.7 0.05
B3_LM 34 1.0 1.0 0.03
B4_SH 67 1.9 2.1 0.06
B4_LM 22 0.6 0.7 0.02
B5_SH 3 0.1 0.1 0.00
B5_LM 113 3.2 3.5 0.10
C_SH 2 0.1 0.1 0.00
C_LM 20 0.6 0.6 0.02
Total 21,775 589.7 675.9 19.1

*By zone for part of the Hugoton model covered by Grant and Stevens counties, Kansas, for 450 psi (3103 kPa) initial bottom-hole pressure.

slightly low compared to earlier work by others. For It is important to note that estimation of OGIP
the Chase in Kansas, Oberst et al. (1994) estimated using the matrix capillary-pressure method em-
OGIP volumetrically at 31.1 tcf (0.88 trillion m3), ployed in this study is influenced by natural var-
whereas Olson et al. (1997) placed it at 34.5 – 37.8 tcf iance in the capillary-pressure curves and the de-
(0.98 – 1.1 trillion m3). Because their estimates were termination of free-water level. Natural variance in
for different reservoir volumes than ours (Chase in capillary-pressure curves can result in a one stan-
Kansas versus Chase and Council Grove in two coun- dard deviation confidence interval for predicted wa-
ties in Kansas), we cannot compare directly, but assum- ter saturation of greater than 10% of the saturation
ing similar reservoir performance, we can compare value (e.g., Sw = 10% results in 9% < Sw < 11% or,
the estimates on the basis of production efficiency. for S w = 80%, 72% < S w < 88%), which results
The ratio of Chase cumulative production to date in a one standard deviation confidence interval for
(24.8 tcf; 0.7 trillion m3) to OGIP is 79.7% by Oberst predicted OGIP of approximately 3%. Change in
et al. (1994) and 65.6– 71.9% by Olson et al. (1997). the free-water level results in little change in water
Our estimate for the entire Wolfcamp reservoir vol- saturation for intervals greater than 300 ft (91 m)
ume (65%) is closer to the Olson et al. (1997) esti- above FWL, but can have significant influence on
mate. In Kansas, 89.2% of the Hugoton – Panoma intervals that are within their transition zone and
production is attributed to the Chase and 10.8% to have rocks that exhibit transition zones that are
the Council Grove, although we believe the two be- only tens of feet high. For these intervals, shift in
have as one reservoir system. the FWL by a few tens of feet can result in significant
350 / Dubois et al.

FIGURE 40. Map views of original gas


in place (OGIP) for Grant and Stevens
counties, Kansas (A), are the two most
prolific counties in Hugoton. Hydro-
carbon height (HCH) at surface condi-
tions for Chase Group (B), Council Grove
Group (C), and combined, Wolfcamp
(D). Wolfcamp volume map varies little
from the Chase because the Council
Grove volume is small relative to the
Chase (note the smaller scale range
and contour interval for the Council
Grove). Areas of high Chase OGIP are
coincident with Krider ooid shoal
complex illustrated in Figure 38D – F.

characterization and modeling work-


flow for a giant field. Core-based cali-
bration of neural net prediction
of lithofacies using wire-line-log
signatures, coupled with geologic-
constraining variables, provided ac-
curate lithofacies models at well to
field scales. Differences in petrophy-
sical properties among lithofacies and
within a lithofacies among different
porosities illustrate the importance of
integrated lithologic-petrophysical
modeling and of the need for closely
defining these properties and their
relationships. Lithofacies models,
coupled with lithofacies-dependent
petrophysical properties, allowed
the construction of a 3-D geomodel
for the Hugoton that has been ef-
fective at the well, section (1 mi2;
2.6 km2) and multisection scales.
The model provided a tool to pre-
dict lithofacies and petrophysical
properties distribution, water satura-
tions, and OGIP. It will likely provide
a quantitative basis for evaluating
remaining gas in place, particularly
water saturation change and, consequently, OGIP in low-permeability intervals, and help direct field
changes. This was important for the Council Grove management and production practices that will po-
OGIP estimate and is still being refined. tentially enhance ultimate recovery. The reservoir char-
acterization and modeling from pore to field scale
CONCLUSIONS discussed provides a comprehensive lithological and
petrophysical view of a mature giant Permian gas sys-
The more than 100-million-cell, 10,000-mi 2 tem. Both the knowledge gained and the techniques
(26,000-km2), 3-D geologic and petrophysical prop- and workflow employed have implications for un-
erty geomodel of the Hugoton presented in this study derstanding and modeling similar reservoir systems
demonstrates the application of a detailed reservoir worldwide.
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 351

ACKNOWLEDGMENTS Hugoton field, Texas County, Oklahoma, in W. L.


Watney, A. W. Walton, C. G. Caldwell, and M. K.
Dubois, organizers, Midcontinent Core Workshop on
We are thankful for the support from partners
Integrated Studies of Petroleum Reservoirs in the
in the Hugoton Asset Management Project, includ-
Midcontinent: Midcontinent AAPG Section Meeting,
ing Anadarko Petroleum Corporation, BP America Wichita, Kansas, p. 57 – 75.
Production Company, ConocoPhillips Company, Ci- Deutsch, C. V., and A. G. Journel, 1998, Geostatistical soft-
marex Energy Company, EOG Resources Inc., Exxon- ware library and user’s guide: Oxford, Oxford Univer-
Mobil Production Company, Medicine Bow En- sity Press, 369 p.
ergy Corporation, Osborn Heirs Company, OXY Dubois, M. K., and R. H. Goldstein, 2005, Accommodation
U.S.A., Inc., and Pioneer Natural Resources U.S.A., model for Wolfcamp (Permian) redbeds at the updip
margin of North America’s largest onshore gas field
Inc., and from the Kansas Geological Survey. We
(abs.): Proceedings AAPG 2005 Annual Convention,
thank Nathan Winters for his assistance on core de- June 19– 21, Calgary, Alberta, Canada, and Kansas Geo-
scriptions, Raymond Sorenson for insightful discus- logical Survey Open-file Report 2005-25: http://www
sions, Shane Seals in work on earlier models, and .kgs.ku.edu/PRS/AAPG2005/2005-25/index.html
geoPlus Corporation and Schlumberger for provid- (accessed December 31, 2005).
ing the software. Reviewers Eugene Rankey, Sean Dubois, M. K., A. P. Byrnes, G. C. Bohling, S. C. Seals, and
Guidry, and Krishnan Srinivasan provided help- J. H. Doveton, 2003, Statistically-based lithofacies pre-
ful comments and suggestions that improved this dictions for 3-D reservoir modeling: Examples from
the Panoma (Council Grove) field, Hugoton embay-
manuscript.
ment, southwest Kansas (abs.): AAPG Annual Meet-
ing Program, v. 12, p. A44, and Kansas Geological Sur-
REFERENCES CITED vey Open-file Report 2003-30, 3 panels: http://www
.kgs.ku.edu/PRS/publication/2003/ofr2003-30/index
Asquith, G., and D. Krygowski, 2004, Basic well log analy- .html (accessed October 10, 2005).
sis, 2d ed.: AAPG Methods in Exploration, v. 16, 244 p. Dubois, M. K., A. P. Byrnes, and G. C. Bohling, 2005,
Baars, D. L., ed., 1994, Revision of stratigraphic nomen- Geologic model for the giant Hugoton and Panoma
clature in Kansas: Kansas Geological Survey Bulletin, fields (abs): AAPG Midcontinent Section Meeting,
v. 230, 80 p. Oklahoma City, Oklahoma: http://www.kgs.ku.edu
Berg, R. R., 1975, Capillary pressures in stratigraphic traps: /PRS/Poster/2005/MidcontAAPG/index.html (accessed
AAPG Bulletin, v. 59, p. 939 – 956. October 10, 2005).
Boardman II, D. R., and M. K. Nestell, 2000, Outcrop-based Dunham, R. J., 1962, Classification of carbonate rocks accord-
sequence stratigraphy of the Council Grove Group of ing to depositional texture, in W. E. Ham, ed., Classifi-
the midcontinent, in K. S. Johnson, ed., Platform car- cation of carbonate rocks: AAPG Memoir 1, p. 108 –121.
bonates in the southern midcontinent, 1996 Sympo- Dutton, S. P., E. K. Kim, C. L. Broadhead, W. D. Raatz, S. C.
sium: Oklahoma Geological Survey Circular 101, p. 275– Ruppel, and C. Kerans, 2005, Play analysis and digital
306. portfolio of major oil reservoirs in the Permian Basin:
Bohling, G. C., and J. H. Doveton, 2000, Kipling.xla: An Bureau of Economic Geology Reports of Investigations,
Excel add-in for nonparametric regression and classi- RI0271, 302 p.
fication, Kansas Geological Survey: http://www.kgs Fetkovitch, M. J., D. J. J. Ebbs Jr., and J. J. Voelker, 1994,
.ku.edu/software/Kipling/Kipling1.html (accessed Multiwell, multilayer model to evaluate infill-drilling
December 31, 2005). potential in the Oklahoma Hugoton field: 65th So-
Byrnes, A. P., 2003, Aspects of permeability, capillary ciety of Petroleum Engineers Annual Technical Con-
pressure, and relative permeability properties and ference and Exhibition, New Orleans, Louisiana, SPE
distribution in low-permeability rocks important to Paper 20778, p. 162 – 168.
evaluation, damage, and stimulation: Proceedings of Folk, R. L., 1954, The distinction between grain size and
the Rocky Mountain Association of Geologists Petro- mineral composition in sedimentary rock nomencla-
leum Systems and Reservoirs of Southwest Wyoming ture: Journal of Geology, v. 62, p. 344 – 359.
Symposium, Denver, Colorado, September 19, 2003, Garlough, J. L., and G. L. Taylor, 1941, Hugoton gas field,
12 p. Grant, Haskell, Morton, Stevens, and Seward counties,
Byrnes, A. P., M. K. Dubois, M. Magnuson, 2001, Western Kansas, and Texas county, Oklahoma, in A. I. Levorsen,
tight gas carbonates: Comparison of Council Grove ed., Monograph Stratigraphic type oil fields: AAPG,
Group, Panoma field, southwest Kansas and western low p. 78 – 104.
permeability sandstones (abs.): AAPG Annual Meeting Gaymard, R., and A. Poupon, 1968, Response of neutron
Program, v. 10, p. A31. and formation density logs in hydrocarbon bearing
Caldwell, C. D., 1991, Cyclic deposition of the Lower Perm- formations: The Log Analyst, v. 9, no. 5, p. 3 – 12.
ian, Wolfcampian, Chase Group, western Guymon– George, B. K., C. Torres-Verdin, M. Delshad, R. Sigal, F.
352 / Dubois et al.

Zouioueche, and B. Anderson, 2004, Assessment of in- Lower Permian Krider Member, southwest Kansas,
situ hydrocarbon saturation in the presence of deep U.S.A.: Fluid-inclusion, U-Pb, and fission-track evi-
invasion and highly saline connate water: Petrophysics, dences for reflux dolomitization during latest Permian
v. 45, no. 2, p. 141– 156. time: Journal of Sedimentary Research, v. 70, no. 3,
Grammer, G. M., G. P. Eberli, F. S. P. van Buchem, G. M. p. 762 – 773.
Stevenson, and P. Homewood, 1996, Application of Mason, J. W., 1968, Hugoton and Panhandle field, Kansas,
high-resolution sequence stratigraphy to evaluate lat- Oklahoma and Texas, in W. B. Beebe and B. F. Curtis,
eral variability in outcrop and subsurface; Desert Creek eds., Natural gases of North America, v. 2: AAPG
and Ismay intervals, Paradox basin, in M. W. Longman Memoir 9, p. 1539 – 1547.
and M. D. Sonnenfeld, eds., Paleozoic systems of the Mazzullo, S. J., C. S. Teal, and C. A. Burtnett, 1995, Facies
Rocky Mountain region: Rocky Mountain Section So- and stratigraphic analysis of cyclothemic strata in the
ciety for Sedimentary Geology, p. 235 – 266. Chase Group (Permian Wolfcampian), south-central
Hastie, T., R. Tibshirani, and J. Friedman, 2001, The ele- Kansas, in N. J. Hyne, ed., Sequence stratigraphy of
ments of statistical learning: Data mining, inference, the mid-continent: Tulsa Geological Society Special
and prediction: New York, Springer, 533 p. Publication 4, p. 217 – 248.
Hecker, M. T., M. E. Houston, and J. D. Dumas, 1995, McGillivray, J. G., and M. I. Husseini, 1992, The Paleozoic
Improved completion designs in the Hugoton field petroleum geology of central Arabia: AAPG Bulletin,
utilizing multiple gamma emitting tracers: Society of v. 76, p. 1491 – 1506.
Petroleum Engineers Annual Technical Conference and Oberst, R. J., P. P. Bansal, and M. F. Cohen, 1994, 3-D
Exhibition, Dallas, Texas, SPE Paper 30651, p. 223 –235. reservoir simulation results of a 25-square mile study
Hemsell, C. C., 1939, Geology of Hugoton gas field of south- area in Kansas Hugoton gas field: Society of Petroleum
western Kansas: AAPG Bulletin, v. 23, no. 7, p. 1054– Engineers Mid-Continent Gas Symposium, SPE Paper
1067. 27931, p. 137 – 147.
Heyer, J. F., 1999, Reservoir characterization of the Council Olson, T. M., J. A. Babcock, and P. D. Wagner, 1996,
Grove Group, Texas County, Oklahoma, in D. F. Mer- Geologic controls on reservoir complexity, Hugoton
riam, ed., AAPG Midcontinent Section Meeting Trans- giant gas field, Kansas, in D. L. Swindler and C. P.
actions, Geosciences for the 21st Century, p. 71 – 82. Williams, compilers, AAPG Mid-continent Section Meet-
Hough, E. W., M. J. Rzasa, and B. B. Wood, 1951, Interfacial ing Transactions: Mid-continent Section, p. 189–198.
tensions at reservoir pressures and temperatures; Olson, T. M., J. A. Babcock, K. V. K. Prasad, S. D. Boughton,
apparatus and the water-methane system: American P. D. Wagner, M. K. Franklin, and K. A. Thompson,
Institute of Mechanical Engineering Petroleum Trans- 1997, Reservoir characterization of the giant Hugo-
actions, Technical Publication 3019, v. 192, p. 57 – 60. ton gas field, Kansas: AAPG Bulletin, v. 81, p. 1785 –
Hubbert, M. K., 1953, Entrapment of petroleum under hy- 1803.
drodynamic conditions: AAPG Bulletin, v. 37, p. 1954 – Olszewski, T. D., and M. E. Patzkowsky, 2003, From cy-
2026. clothems to sequences: The record of eustacy and cli-
Hubbert, M. K., 1967, Application of hydrodynamics to mate on an icehouse eperic platform (Pennsylvanian –
oil exploration: 7th World Petroleum Congress Pro- Permian), North American Midcontinent: Journal of
ceedings, Mexico City: Amsterdam, Elsevier Publishing Sedimentary Research, v. 73, no. 1, p. 15 – 30.
Co. Ltd, v. 1B, p. 59 – 75. Parham, K. D., and J. A. Campbell, 1993, PM-8. Wolf-
Jennings Jr., H. Y., and G. H. Newman, 1971, The effect of campian shallow shelf carbonate — Hugoton embay-
temperature and pressure on the interfacial tension of ment, Kansas and Oklahoma, in D. G. Bebout, ed., Atlas
water against methane-normal decane mixtures: Trans- of major midcontinent gas reservoirs: Gas Research
actions of the American Institute of Mining Metallur- Institute, p. 9 – 12.
gical and Petroleum Engineers, v. 251, p. 171 – 175. Parrish, J. T., 1995, Geologic evidence of Permian climate,
Kluth, C. F., 1986, Plate tectonics of the ancestral Rocky in P. A. Scholle, T. M. Peryt, and D. S. Ulmer-Scholle,
Mountains, in J. A. Peterson, ed., Paleotectonics and eds., The Permian of northern Pangea: v. I. Paleoge-
sedimentation of the Rocky Mountains, United States: ography, paleoclimate, stratigraphy: Berlin, Germany,
AAPG Memoir 41, p. 353 – 369. Springer-Verlag, p. 53 – 61.
Kluth, C. F., and P. J. Coney, 1981, Plate tectonics of the Parrish, J. T., and E. Peterson, 1988, Wind directions pre-
ancestral Rocky Mountains: Geology, v. 9, no. 1, p. 10 – dicted from global circulation models and wind di-
15. rections determined from eolian sandstones of the
Konnert, G., A. M. Afifi, S. A. Al-Hajri, K. de Groot, A. A. Al western United States — A comparison: Sedimentary
Naim, and H. J. Droste, 2001, Paleozoic stratigraphy Geology, v. 56, p. 261 – 282.
and hydrocarbon habitat of the Arabian plate, in Perry, W. J., 1989, Tectonic evolution of the Anadarko
M. W. Downey, J. C. Threet, and W. A. Morgan, eds., basin region, Oklahoma: U.S. Geological Survey Bulle-
Petroleum provinces of the twenty-first century: tin, v. 1866-A, p. A1 – 16.
AAPG Memoir 74, p. 483 – 515. Pippin, L., 1970, Panhandle–Hugoton field, Texas–Okla-
Luczaj, J. A., and R. H. Goldstein, 2000, Diagenesis of the homa–Kansas— The first fifty years, in M. T. Halbouty,
Geologic and Petrophysical Modeling of the Hugoton Gas Field, Kansas and Oklahoma / 353

ed., Geology of giant petroleum fields: AAPG Memoir 14, Group, Guymon – Hugoton field, Oklahoma, Society
p. 204–222. of Petroleum Engineers Proceedings 65th Annual Tech-
Puckette, G. R., D. R. Boardman II, and Z. Al-Shaieb, 1995, nical Conference and Exhibition, New Orleans, Louisi-
Evidence for sea-level fluctuation and stratigraphic se- ana, September 23 – 26, 1990, SPE Paper 20757, p. 417 –
quences in the Council Grove Group (Lower Permian) 428.
Hugoton embayment, southern mid-continent, in N. J. Soreghan, G. S., 2002, Sedimentologic-magnetic record of
Hyne, ed., Sequence stratigraphy of the mid-continent: western Pangean climate in upper Paleozoic loessites
Tulsa Geological Society Special Publication 4, p. 269– (lower Cutler beds, Utah): Geological Society of Ameri-
290. ca Bulletin, v. 114, no. 8, p. 1019 – 1035.
Purcell, W. R., 1949, Capillary pressure — Their measure- Sorenson, R. P., 2005, A dynamic model for the Permian
ments using mercury and the calculation of perme- Panhandle and Hugoton fields, western Anadarko
ability therefrom: American Institute of Mechanical basin: AAPG Bulletin, v. 89, no. 7, p. 921 – 938.
Engineers Petroleum Transactions, v. 186, p. 39 – 48. Venables, W. N., and B. D. Ripley, 2002, Modern applied
R Development Core Team, 2005, R: A language and en- statistics with S, 4th ed.: New York, Springer, 512 p.
vironment for statistical computing: R Foundation for Washburn, E. W., 1921, A method of determining the dis-
Statistical Computing, Vienna, Austria: http://www tribution of pore sizes in a porous material: Proceed-
.R-project.org (accessed December 31, 2005). ings of the National Academy of Sciences, v. 7, p. 115 –
Rankey, E. C., 1997, Relations between relative changes in 116.
sea level and climate shifts; Pennsylvanian – Permian Weber, L. J., F. M. Wright, J. F. Sarg, E. Shaw, L. P. Harman,
mixed carbonate-siliciclastic strata, western United J. B. Vanderhill, and D. A. Best, 1994, Reservoir de-
States: Geological Society of America Bulletin, v. 109, lineation and performance; application of sequence
no. 9, p. 1089 – 1100. stratigraphy and integration of petrophysics and en-
Rascoe Jr., B., 1968, Permian system in western midconti- gineering data, Aneth field, southeast Utah, U.S.A., in
nent: Mountain Geologist, v. 5, p. 127 – 138. E. L. Stout and P. M. Harris, eds., Hydrocarbon reser-
Rascoe Jr., B., and F. J. Adler, 1983, Permo-Carboniferous voir characterization; geologic framework and flow
hydrocarbon accumulations, midcontinent, U.S.A.: unit modeling: Society of Sedimentary Geology, p. 1 –
AAPG Bulletin, v. 67, p. 979 – 1001. 29.
Ryan, T. C., M. J. Sweeney, and W. H. Jamieson Jr., 1994, Winters, N. D., M. K. Dubois, and T. R. Carr, 2005, Depo-
Individual layer transient tests in low-pressure, multi- sitional model and distribution of marginal marine
layered reservoirs: Society of Petroleum Engineers sands in the Chase Group, Hugoton gas field, southwest
Mid-Continent Gas Symposium, Amarillo, Texas, SPE Kansas and Oklahoma Panhandle (abs.): AAPG Mid-
Paper 27928, p. 99 – 113. continent Section Meeting, Oklahoma City, Oklahoma:
Scotese, C. R., 2004, A continental drift flipbook: Journal http://www.kgs.ku.edu/PRS/Poster/2005/MidcontAAPG
of Geology, v. 112, p. 729 – 741. /index.html (accessed October 10, 2005).
Siemers, W. T., and W. M. Ahr, 1990, Reservoir facies, pore Zeller, D. E., ed., 1968, The stratigraphic succession in Kan-
characteristics, and flow units: Lower Permian Chase sas: Kansas Geological Survey Bulletin, v. 169, 81 p.
10
Ruppel, S. C., and R. H. Jones, 2006, Key role of outcrops and cores
in carbonate reservoir characterization and modeling, Lower Permian
Fullerton field, Permian basin, United States, in P. M. Harris and L. J.
Weber, eds., Giant hydrocarbon reservoirs of the world: From rocks
to reservoir characterization and modeling: AAPG Memoir 88/SEPM
Special Publication, p. 355 – 394.

Key Role of Outcrops and


Cores in Carbonate Reservoir
Characterization and Modeling,
Lower Permian Fullerton Field,
Permian Basin, United States
Stephen C. Ruppel
Bureau of Economic Geology, Jackson School of Geosciences, University of Texas at Austin,
Austin, Texas, U.S.A.

Rebecca H. Jones
Bureau of Economic Geology, Jackson School of Geosciences, University of Texas at Austin,
Austin, Texas, U.S.A.

ABSTRACT

T
he analysis of reservoir sequence and cycle stratigraphy, of depositional
and diagenetic facies, and of the interrelationships between these
attributes and reservoir properties is key to the construction of an accurate
reservoir framework needed for reservoir modeling and improved imaging of
remaining hydrocarbons. Fundamental steps in such a rock-based process of
model construction applied at Fullerton Clear Fork field, a shallow-water car-
bonate platform reservoir of middle Permian age, included (1) creating and ap-
plying an analogous outcrop depositional model; (2) describing and interpreting
subsurface core and log data in terms of this initial model; (3) defining the
sequence-stratigraphic architecture of the reservoir section; (4) developing a cycle-
based reservoir framework; and (5) defining controls, interrelationships, and dis-
tribution of porosity and permeability. Data used in this analysis included cores,
thin sections, three- and two-dimensional seismic data, borehole image logs, and
outcrop models.
Key geological elements addressed and incorporated into the models include
stratal architecture, differential dolomitization, karst fill, mineralogical variations,
and rock-fabric distribution. These components were used to constrain inter-
pretation and definition of flow units, permeability distribution, and saturation.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215882M88698

355
356 / Ruppel and Jones

In addition to resulting in improved and more geologically realistic reservoir


models, the rock-based methods used in this study provide key insights into the
formation, characterization, and interpretation of carbonate platform reservoirs;
these insights have widespread application worldwide.

INTRODUCTION the basis for modeling and interpreting past, present,


and future recovery operations. This report details
The Clear Fork Group (Figure 1) in the Permian approaches used to develop such a framework in the
basin comprises a thick (as much as 800 m [2500 ft]) Fullerton Clear Fork field in west Texas (Figures 2, 3).
succession of dominantly shallow-water platform By any standard, Fullerton is a giant. As of 2002,
carbonates that were deposited across west Texas and cumulative production stood at about 310 million
New Mexico during the Early Permian (Leonardian). bbl, about 18% of the approximately 1.69 billion bbl
Reservoirs developed in these carbonates (Figure 2) of OOIP (Wang and Lucia, 2004). Like many mature
have accounted for more than 3.2 billion bbl of oil Clear Fork reservoirs in the Permian basin, Fullerton
production (Dutton et al., 2005), more than 10% of field has undergone a major decline in oil production
the total recovered from the Permian basin to date. rate and a major increase in water production rates
Despite this substantial production, estimates of origi- for several years. An understanding of the deposi-
nal oil in place (OOIP) indicate that, overall, Leonar- tional and diagenetic facies, the cycle and sequence
dian reservoirs contained more than 14.5 billion bbl stratigraphy, and the architecture of these stratigraph-
of oil at discovery. Recovery efficiency is thus only ic elements is the crucial first step to determining the
about 22%, considerably below the 32% average for probable oil distribution at discovery and defining the
carbonate reservoirs in the Permian basin (Tyler and best strategies for recovering the sizeable remaining
Banta, 1989; Holtz and Garrett, 1990). Recovery from oil volume.
the shallow-water platform reservoirs of the Clear Fork
Group has been even less efficient. Holtz et al. (1992) METHODS
estimated a recovery efficiency of only 18% of OOIP
for these reservoirs. Data and interpretations presented in this report
To recover the remaining oil in Clear Fork res- are based on investigation of 29 cores totaling 4384 m
ervoirs, operators must turn to increasingly sophisti- (14,383 ft), the examination of more than 1700 rock
cated recovery technologies, e.g., water flooding, gas thin sections, and the correlation of approximately
injection, horizontal wells, etc. To effectively deploy 45,000 stratigraphic tops in more than 800 wells.
these technologies, however, it is critical that an accu- The basic procedure followed in this study to develop
rate reservoir framework first be constructed to form a full-field sequence-stratigraphic model and reser-
voir framework is as follows:
(1) identify depositional facies
and vertical facies-stacking
patterns and cycles in cores;
(2) calibrate facies and cycle
patterns to wire-line logs;
(3) construct two-dimensional
(2-D) cross sections of core/
log data sections; (4) define

FIGURE 1. Leonardian strati-


graphic section in the Permian
basin including the Clear Fork
Group and analogous units in
New Mexico and in outcrop.
Productive interval at Fullerton
field is shown in color.
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 357

FIGURE 2. Regional map of the Permian basin showing the location of the Fullerton field and major reservoirs producing
from Leonardian rocks.

2-D cycle and sequence stratigraphy from core and Presley and McGillis (1982; see also Presley, 1987)
log sections; and (5) extrapolate 2-D cycle and se- documented the highly cyclic, predominantly evap-
quence correlations into three-dimensional (3-D) space oritic facies of the upper Leonardian Glorieta and
by correlating well logs. Stratigraphic sequences were Upper Clear Fork units in the Texas Panhandle. Ruppel
tied to available 3-D seismic data to check their ac- (1992, 2002) described the facies, cyclicity, and dia-
curacy and geometry. Conventional core analysis data genesis in the Glorieta and Upper Clear Fork at Mona-
were available for all cores; these data were used to hans Clear Fork field on the Central Basin platform
determine relationships between facies, cyclicity, and and postulated that reservoir development was caused
porosity and permeability development. Figure 3 by cyclic deposition and diagenesis driven by episodic
shows the distribution of cores and 3-D seismic data sea level rise and fall. Atchley et al. (1999) described
in Fullerton field. Clear Fork facies at Robertson field at the north end of
the Central Basin platform and proposed a similar
PREVIOUS WORK model for structural control over facies deposition
and reservoir development. Ruppel et al. (2000) de-
Mazzullo (1982) and Mazzullo and Reid (1989) scribed outcrop equivalents of the producing subsur-
presented overviews of lower Leonardian stratigra- face Clear Fork reservoirs from outcrops in the Sierra
phy and depositional systems in the Midland basin. Diablo Mountains of west Texas. Kerans et al. (2000)
358 / Ruppel and Jones

about 210 m (700 ft) in thickness at a depth of about


2040 m (6700 ft) (Figures 2, 3). In some parts of the
Permian basin, oil production is obtained from the
entire Leonardian section (Figure 1). The productive
reservoir section at Fullerton, however, is essentially
restricted to the Lower Clear Fork, Wichita, and Abo
stratigraphic units (Figure 4), although very minor
production has been reported locally from the Upper
Clear Fork section. By far, the bulk of the oil produc-
tion has come from the lower two-thirds of the Lower
Clear Fork and the Wichita sections of the reservoir.
Even where oil saturated, the Abo has proven difficult
to exploit in part because of its active water drive (as
opposed to the pressure depletion drive that charac-
terizes the overlying parts of the reservoir) and the
fact that it is at or near the oil-water contact through-
out most of the field.
Regionally, the Leonardian is dominated by shallow-
water platform carbonates. Each of the component
stratigraphic units contains updip peritidal tidal-flat
carbonates and downdip subtidal carbonates. Min-
eralogy in each is dominated by dolomite and an-
hydrite. Limestone is relatively uncommon; however,
where present, it is most common in the lower part of
the Leonardian section and in distal, downdip sec-
tions (Ruppel, 2002). The reservoir section at Fuller-
ton is generally consistent with this regional pattern,
but it does contain a higher volume of limestone than
most other Leonardian platform reservoir successions
in the Permian basin.
Structurally, Fullerton field is developed over a
large, compound structural high (Figure 3) that re-
flects deep-seated faulting and differential uplift of
the area that began in the Pennsylvanian (Jones and
Ruppel, 2004). Deeper oil production comes from
FIGURE 3. Structure of the Fullerton Clear Fork field.
Datum is a prominent subtidal flooding event near the
block-faulted Silurian (Wristen Group) carbonates in
base of the Tubb Formation (see Figures 4, 5). the southern and central parts of the fields. The Clear
Fork reservoir seal is provided by evaporite-rich car-
bonates of the Upper Clear Fork and Glorieta.
documented the depositional setting, facies, and archi- The reservoir is currently drilled to well spacings of
tecture of the Abo from outcrops in the Sierra Diablo 16 to 4 ha (40 to 10 ac) and is under active water
(Hudspeth and Culberson counties, Texas) and showed flood. The greatest well density is in the northern
that karsting has had a major effect on both the Abo part of the field (mostly 8- and 4-ha (20- and 10-ac)
and the overlying Lower Clear Fork succession. well spacings); poorer well control exists in the
southern half of the field. In general, these closely
GEOLOGIC SETTING spaced wells provide good control for definition of
stratigraphic horizons and reservoir attributes. In
The Fullerton Clear Fork field is the largest of a many parts of the field, however, especially along the
large number of fields developed in the Leonardian western edge of the field and in the southern half of
series on the Central Basin platform of the Permian the field, wells are represented only by poor-quality
basin (Figure 2). The field covers an area of about (old gamma-ray and neutron) logs and cannot be
14,000 ha (35,000 ac); the reservoir column averages correlated or interpreted with any precision.
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 359

FIGURE 4. Type log of producing res-


ervoir section at Fullerton Clear Fork
field showing cyclicity and general
facies distribution. Shaded areas of the
PE log indicate the local presence of
limestone in the dolomite-dominated
section. Zones of elevated corrected
gamma-ray (CGR) log response indicate
the presence of silt and clay associated
with tidal-flat facies. SGR = spectral
gamma ray; PE = photoelectric; Neu =
neutron; Den = density.

karsted, platform-margin subtidal


carbonates of the Abo (Fitchen et al.,
1995; Kerans et al., 2000). Styles of
depositional architecture and facies
development are very representative
of subsurface succession and, thus,
form excellent, if not essential, mod-
els for interpreting sparser subsur-
face data sets. Key observations es-
tablished from these outcrops are
described below.

Abo
Integrated studies of outcrops and
subsurface data sets indicate that the
Abo represents the basal depositional
sequence (sequence L1) of the Leonar-
dian (Fitchen et al., 1995). Studies of
Abo outcrops in the Sierra Diablo
(Fitchen et al., 1995; Kerans et al.,
2000) have demonstrated three im-
portant aspects of this succession
in the Permian basin: (1) it consists
of dominantly open-marine, outer-
platform facies; (2) it displays clino-
formal architecture; and (3) it is over-
printed by karst features (sinkholes,
caves, cave fill, and collapse features).
OUTCROP ANALOGS The dominance of clinoformal, outer-ramp, fusulinid-
crinoid packstones and wackestones and less common
Studies of analogous reservoir outcrops in the ramp-crest, ooid-peloid, grain-rich packstones and
Sierra Diablo in Hudspeth and Culberson counties, grainstones in the Abo contrasts with the flat-lying,
Texas (Figure 2), provide important insights into the alternating tidal-flat and shallow subtidal wackestone-
geological controls on reservoir development in the packstone successions of the Wichita and Lower Clear
Fullerton reservoir and on the reservoir architecture. Fork. The top-lapping clinoforms of the Abo document
The Leonardian of the Sierra Diablo contains direct rapid basinward progradation, a forced regression prob-
analogs of all major reservoir intervals at Fullerton ably caused by a rapid fall in sea level. Karsting in the
field, including the shallow-water platform carbon- Abo, although initiated at the exposed top of the Abo
ates of the Lower Clear Fork Group (Fitchen et al., during sea level fall, is manifested downsection in
1995; Ruppel et al., 2000) and Wichita units and the the Abo by caves and sinkholes and upsection in the
360 / Ruppel and Jones

overlying Lower Clear Fork as collapse features. Where these names are best considered rock-stratigraphic
karsting and associated thickness variations are de- terms (essentially facies) at the formation level of
veloped, largely in platform-marginal settings, the nomenclature. To place these units in their proper
contact between the Abo and Lower Clear Fork is a perspective, however, it is best to consider both facies
relatively sharp and undulating, unconformable sur- and time interrelationships. In this report, we treat
face. Updip, karsting is less apparent, and the contact these named stratigraphic units (i.e., formations) as
is less pronounced. facies or rock-stratigraphic units but place them in a
sequence-stratigraphic (i.e., time-stratigraphic) frame-
Lower Clear Fork work that has been developed from previous studies
The outcropping Lower Clear Fork succession in (e.g., Fitchen et al.,1995; Kerans et al., 2000; Ruppel
the Sierra Diablo represents a single depositional se- et al., 2000). On the basis of these studies and data
quence (L2). Key elements of this succession are from Fullerton field, we interpret the Abo and the lower
(1) basal backstepping tidal-flat deposits that locally part of the Wichita Formation to be time-equivalent
fill relief on the underlying, karsted Abo surface; facies (or systems tracts) of the earliest composite
(2) an updip succession of amalgamated tidal-flat (third-order) depositional sequence of the Leonardian
facies; (3) a cyclic, downdip succession of alternating (sequence L1). The overlying sequence (L2) comprises
tidal-flat and midramp, subtidal facies; and (4) an over- updip peritidal deposits of the upper Wichita Forma-
all backstepping (upward-deepening) trend (Fitchen tion and downdip, dominantly subtidal deposits of the
et al., 1995; Kerans et al., 2000; Ruppel et al., 2000). Lower Clear Fork Formation (Figure 5). Sequence L3
Lower Clear Fork deposits are characterized by contains basal transgressive systems tract (TST) silic-
alternating peritidal, tidal-flat deposits and subtidal, iclastics of the Tubb Formation and carbonates of part
skeletal wackestones and packstones that document of the overlying Upper Clear Fork (Figure 4).
cyclic rise and fall of sea level at the high-frequency Correlations of usable wire-line-log suites (about
sequence and cycle scale. These cycles, which average 850) show that across most of the area of the Fuller-
about 6 m (20 ft) in thickness, display consistent ton Clear Fork unit (Figure 3), stratigraphic units are
patterns of facies stacking (cycle-base skeletal wacke- relatively isopachous. This is consistent with the de-
stones and overlying peloidal grain-rich packstones) positional setting of the field area on the Central
and appear to be widely continuous (Ruppel et al., Basin platform, a broad, flat carbonate platform not
2000). These midramp, shallow subtidal platform unlike the modern-day Bahamas platform.
rocks pass downdip into clinoformal fusulinid-crinoid The generally isopachous nature of the Wichita and
wackestones and packstones of the outer ramp-slope Lower Clear Fork is confirmed by 3-D and 2-D seismic
in less than 3 km (2 mi) basinward. Updip (landward), data. These data, however, show that the Wichita and
the Lower Clear Fork is characterized by amalgamated, Abo vary considerably in thickness across the field.
inner-platform, peritidal, tidal-flat deposits. These Wichita deposits thin, and Abo deposits thicken to
updip tidal flats are analogous to Wichita facies in the east and southeast (Figure 6). Synthesis of core,
the subsurface. The absence of shallow-water high- wire-line, and seismic data suggests that this recipro-
stand deposits at the top of the Lower Clear Fork cal thickness relationship is the result of facies change;
suggests a rapid, perhaps forced, regression, followed i.e., the Wichita represents the updip, shallower water
by exposure and possible erosion at the top of the (inner-platform) equivalent of the distal, deeper water
L2 sequence. (outer platform to slope) Abo. Support for this con-
clusion is presented in subsequent sections.
FULLERTON RESERVOIR FACIES
AND STRATIGRAPHY Facies and Depositional Setting
Reservoir facies encountered at Fullerton field are
The productive reservoir section at Fullerton field typical of those observed throughout most carbonate
includes parts of the Abo, Wichita, and Lower Clear platform successions of Leonardian and early Gua-
Fork formations (Figures 4, 5). Like many stratigraph- dalupian age in the Permian basin. The character-
ic names used in the subsurface, each of these units is istics of these and similar facies have been document-
commonly considered to display both regional time ed by many authors, including Bebout et al. (1987),
equivalency and facies constancy across the region. Ruppel and Cander (1988a, b), Garber and Harris
In other words, each is considered to be both a rock- (1990), Longacre (1990), Kerans et al. (1994; see also
stratigraphic and a time-stratigraphic unit. In fact, Kerans and Fitchen, 1995; Kerans and Kempter, 2002),
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 361

FIGURE 5. Northwest – southeast cross section (AA0 ) across the Fullerton field area showing sequence architecture and
general facies development based on cored well control. Line of section shown in Figure 3. Depth in feet.

FIGURE 6. Three-dimensional
seismic section (in time) across
the southern part of Fullerton
field showing the seismic defi-
nition of the Clear Fork reservoir
section. Yellow lines are time-
line boundaries; dotted line de-
fines the top of the Abo Forma-
tion. Note that whereas the
upper Wichita and Lower Clear
Fork intervals are essentially iso-
pachous and continuous across
the field, the lower Wichita and
Abo display reciprocal thick-
ness relationships. Dashed lines
are top-lapping Abo clinoforms.
UCF = Upper Clear Fork; LCF =
Lower Clear Fork.
362 / Ruppel and Jones

FIGURE 7. Core and thin-section photo-


graphs of typical Wichita tidal-flat facies.
(A) Slab photo of peritidal mudstone-
wackestone showing weak laminations
and local burrowing. Core is 10 cm
(4 in.) wide. Exxon FCU 5927, depth:
2140 m (7021 ft). (B) Photomicrograph
of peritidal mudstone-wackestone fa-
cies showing abundant intercrystalline
porosity. Exxon FCU 6122, depth: 2162 m
(7092 ft). Porosity: 13.8%; permeabil-
ity: 1.6 md. (C) Slab photo of clay-rich
carbonate mudstone facies in the Wichita.
Core is 10 cm (4 in.) wide. Pan American
FM-1, depth: 2234 m (7329 ft). Note
underlying fenestral mudstone.

on differences in grain preservation,


diagenesis, grain size, etc. Most can
be found in any part of the reservoir
succession.

Peritidal Mudstone-wackestone
These rocks, which are most abun-
dant in the Wichita but also locally
common in the Lower Clear Fork, are
massive to parallel-laminated mud-
rich rocks that only sparsely con-
tain grains other than a few peloids
(Figure 7A, B). They are most typi-
and Ruppel and Bebout (2001). Leonardian facies cally associated with the exposed tidal-flat facies
successions have been documented both in outcrop and the clay-rich carbonate mudstone facies and are
(Ruppel et al., 2000) and in the subsurface (Atchley interpreted to represent peritidal deposition on a very
et al., 1999; Ruppel, 2002). low-energy tidal flat. They are dominantly dolomi-
Facies groupings defined in this study and pre- tized, and where dolomite, they may display very high
sented below were defined with three goals in mind: porosities (as much as 15%). Because this porosity is
(1) to document the characteristics of distinct depo- caused by intercrystalline pores in fine-crystalline
sitional settings; (2) to document widespread and po- dolomite, however, the permeability is generally low.
tentially correlatable sediment packages; and (3) to In limestone, the facies invariably exhibits very low
create a sound geological basis for assessing relation- porosity (typically less than 2%) and permeability.
ships between rock textures and fabrics and petro- Gamma-ray response is typically medium to high and
physics. In many ways, it is the last of these goals that highly variable in these rocks because of the local
is most important because it provides the foundation presence of clay minerals.
for reservoir modeling. However, as a result of this,
facies characteristics used for subdivision are domi-
nantly small-scale matrix properties, e.g., grain size Clay-rich Carbonate Mudstone
and shape, grain type. Larger scale features, such as These rocks are found intimately associated with
fractures, anhydrite nodules, and evidence of burrow- rocks of the peritidal mudstone-wackestone facies in
ing, although noted, are not used for facies subdivision. the Wichita. They are typically dark gray to nearly
Twelve facies can be at least locally identified in black and are found in beds 1– 8 cm (0.4 –3.1 in.) in
the Fullerton reservoir succession. Most of these are thickness (Figure 7C). Where cyclic upward-shallow-
intergradational with one or more others. In many ing successions can be identified, these deposits are
cases, their distinction is somewhat subjective, based found at or near cycle tops in both the Lower Clear Fork
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 363

FIGURE 8. Core, thin-section, and


borehole image log photographs of
typical tidal-flat facies in the Wichita
and Lower Clear Fork. (A) Slab photo of
exposed tidal-flat facies showing typi-
cal parallel laminations and fenestral
pores. Core is 10 cm (4 in.) wide. Lower
Clear Fork Exxon FCU 7322, depth: 2075 m
(6808 ft). (B) Image log of laminated
tidal-flat facies. Lower Clear Fork, HFS
L2.1. Exxon FCU 2564, depth: 2103–2104 m
(6900–6904 ft). Width of image is 63.5 cm
(25 in.). (C) Photomicrograph of exposed
tidal-flat facies showing fenestral vuggy
porosity. Lower Clear Fork Exxon FCU
6122, depth: 2104 m (6903 ft). (D) Slab
photo of cycle top showing fenestral
exposed tidal-flat facies overlain by
subtidal, burrowed peloid wackestone.
Note small lithoclasts above cycle top.
Lower Clear Fork, top HFS L2.2. Core is
10 cm (4 in.) wide. Exxon FCU 7322,
depth: 2074 m (6804 ft).

demonstrate, however, that they were


formed during at least local sea level
fall and exposure. Porosity is locally
very high (at least 30%) and is as-
sociated with a combination of fe-
nestral pores and intercrystalline
and interparticle pores. Permeability
is lower than in grain-dominated
subtidal rocks but can be significant
where high porosities are present.
However, this study and others (e.g.,
and the Wichita. Although they are probably con- Ruppel, 2002) show that effective permeabilities
tributors to the high gamma-ray response observed in (0.1 md) are associated only with porosities above
the Wichita and in Lower Clear Fork peritidal intervals, approximately 12%. Like the peritidal mudstone-
because of their thinness, they cannot be discretely wackestone facies, these rocks typically display some-
defined by logs and, therefore, are not correlatable. what elevated and variable gamma-ray response. This
Their color and clay content and lack of apparent lat- can sometimes be used to distinguish these rocks
eral continuity suggest that they were formed as local from overlying and underlying subtidal facies.
organic-rich ponds or stagnant pools on the tidal flat.
They may act as local baffles to reservoir fluid flow but Peloid Wackestone
are probably very discontinuous laterally. This grain-poor facies differs from the rocks of the
peritidal mudstone-wackestone facies in its associa-
Exposed Tidal Flat tion and sedimentary structures. The rocks are in-
These deposits show obvious evidence of exposure variably associated with other, demonstrably subtidal,
such as fenestral pores, pisolites, mud cracks, insect facies; they are burrowed; and they locally contain
burrows, sheet cracks, tepee structures, and cyano- skeletal debris (Figure 9). These features suggest that
bacterial or microbial laminations (Figure 8). They they were deposited in low-energy subtidal settings.
are locally common in both the Lower Clear Fork and Dominant grains are 80 –150-mm-diameter peloids
the Wichita. In the Lower Clear Fork, they define cycle that are probably fecal pellets created by infaunal bur-
tops (Figure 8D); in the Wichita, their eustatic sig- rowers. The absence of skeletal allochems supports a
nificance is less apparent. Their sedimentary structures low-energy, perhaps restricted, setting.
364 / Ruppel and Jones

FIGURE 9. Core, thin-section, and image


log photo images of nodular peloid
wackestone typical of the Lower Clear
Fork HFS L2.2. (A) Core photo showing
anhydrite nodules surrounded by light-
colored halos of higher porosity. Exxon
FCU 6739, depth: 2130 m (6989 ft). Width
of core is 10 cm (4 in.). (B) Image log
depicting anhydrite burrows and rim-
ming porosity as white (high-resistivity)
patches surrounded by black (low-
resistivity) rims. Exxon FCU 2564, depth:
2097 – 2098 m (6879 – 6883 ft). Width
of image is 63.5 cm (25 in.). (C) Thin-
section photomicrograph of low-porosity,
burrowed, peloid wackestone. Exxon
FCU 6429, depth: 2078 m (6817 ft). (D)
Thin-section photomicrograph of peloid
wackestone with moldic pores. Exxon
FCU 6229, depth: 2085 m (6841 ft).

burg Formation (Ruppel and Bebout,


2001), suggesting that these features
are an expected and perhaps predict-
able feature of late highstand sedi-
mentation on Permian shallow-water
carbonate platforms. Burrowed, nod-
ular fabrics are readily defined on
borehole imaging logs (Figure 9B).
Porosity, except where enhanced
around burrows, is generally moldic
and low (Figure 9C, D).

Peloid Packstone
The distinction between these
rocks and those assigned to the pe-
loid wackestone facies (Figure 10) is
commonly based on apparent pe-
loid abundance and can be subjec-
Anhydrite nodules are locally common in these tive. As with the peloid wackestone facies, peloids are
rocks, probably reflecting the postdepositional entry mostly fecal pellets created by burrowing infauna,
of sulfate-bearing diagenetic fluids into permeability although skeletal debris (chiefly mollusk fragments)
pathways created by burrowers. In many cases, nod- are locally common. It should be noted that the pres-
ules occupy solution-widened vertical burrow path- ervation of these pellets is largely a function of diagen-
ways that are surrounded by alteration halos (Figure 9A). esis, both early and late. Nearly all of the low-energy
These halos (which have been reported from many mud-dominated facies in the Permian contain obvious
Permian platform carbonate successions) commonly pellets, indicating that burrowers were ubiquitous in
display a more grain-rich texture, higher porosity and these deposits. A complete gradation in texture exists
permeability, and a depleted oxygen isotope signature between pelleted mudstones and peloid packstones.
(Major et al., 1990; Ruppel and Bebout, 2001). At Ful- Differences in texture are probably most commonly
lerton, they are most commonly developed in high- caused by differences in early diagenesis. Pellets that
frequency sequence (HFS) 2.2. This stratigraphic po- were early lithified and stabilized are more likely
sition (in the late highstand of the L2 composite to be preserved. Thus, peloid (i.e., pellet) packstones
sequence) is analogous in terms of accommodation to are sediments that underwent significant amounts
similar features documented from the younger Gray- of early diagenesis, whereas peloid wackestones and
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 365

FIGURE 10. Core, thin-section, and


image log photographs of mud-rich
peloid packstone-wackestone. (A) Slab
photo of burrowed Lower Clear Fork
peloid packstone (dolostone). Exxon
FCU 6229, depth: 2133 m (6997 ft). Core
is 10 cm (4 in.) wide. (B) Image log of
peloid wackestone-packstone facies
typical of Lower Clear Fork HFS L2.2.
Exxon FCU 2564, depth: 2105 – 2106 m
(6905–6909 ft). Width of image is 63.5 cm
(25 in.). (C) Photomicrograph of Lower
Clear Fork peloid packstone (limestone)
with moldic porosity. Exxon FCU 5927,
depth: 2104 m (6903 ft). (D) Photomicro-
graph of Lower Clear Fork peloid pack-
stone (limestone) with moldic porosity.
Exxon FCU 6229, depth: 2078 m (6817 ft).

transport. These rocks, which con-


tain pellets (90–120 mm), less com-
mon ooids (150–250 mm), and un-
identifiable spherical grains, are
typically well sorted and contain in-
terparticle pores that are either open
or filled with cements (Figure 11).
The interparticle pores indicate
that these peloids acted as true grains
instead of pelleted mud, in contrast
to the intercrystalline and moldic
pores that typify mud-dominated
facies. As discussed above, in some
instances, the grain-dominated tex-
pelleted mudstones underwent relatively little. The ture is associated with vertical burrows and probably
textural significance of these pelleted facies is thus owes its origin to burrow-related diagenesis. For the
more diagenetic than depositional. Accordingly, ap- most part, this facies is restricted to the subtidal legs of
parent variations in peloid abundance in these mud- Lower Clear Fork depositional sequences. These rocks
dominated facies packstones, wackestones, and mud- are among the highest quality reservoir facies in the
stones are not necessarily an indication of differing field. Highest porosity (as much as 20%) and per-
depositional environment or fluctuations in wave meability are encountered in peloid grain-dominated
energy. Instead, they may be predominantly caused packstone limestones (Figure 11C), although dolo-
by local changes in rates and effects of early diagen- stones also exhibit good porosity and permeability
esis. Reflecting these subtleties, borehole image logs (Figure 11B, D). Where burrowed, porosity and per-
cannot differentiate between peloid wackestones meability can be greatly enhanced in and around
and packstones, imaging them all as granular, rela- burrows (Figure 12A). These high-porosity burrow fills
tively homogeneous sediment (Figure 10B). Porosity commonly account for apparent moldic fabric on
in these rocks is commonly associated with intercrys- image logs (Figure 12B).
talline pores and uncommon skeletal moldic pores
(Figure 10C, D). Ooid-peloid Grain-dominated
Packstone-grainstone
Peloid Grain-dominated Packstone These rocks differ from the peloid grain-dominated
Unlike mud-rich peloid facies discussed above, packstone facies, with which they are commonly close-
grain-dominated (or grain-rich) peloid packstones ly associated, in having recognizable ooids (as much
commonly display evidence of possible wave-related as 250–300 mm in diameter) in addition to pervasive
366 / Ruppel and Jones

FIGURE 11. Core and thin-section


photomicrographs of peloid, grain-rich
packstones. (A) Slab photo of Abo
peloid grain-dominated packstone.
University Consolidated IV-25, depth:
2210 m (7252 ft). Core is 10 cm (4 in.)
wide. (B) Photomicrograph of Lower
Clear Fork peloid dolostone grain-
dominated packstone showing inter-
particle porosity. White areas are poi-
kilotopic anhydrite. Exxon FCU 6122,
depth: 2126 m (6976 ft). (C) Photomi-
crograph of Lower Clear Fork peloid
(ooid?) dolostone grain-dominated
packstone-grainstone with interparticle
porosity. Exxon FCU 6946, depth: 2083 m
(6835 ft). (D) Photomicrograph of
Lower Clear Fork peloid grain-dominated
packstone dolostone showing inter-
particle porosity. Exxon FCU 6229, depth:
2078 m (6816 ft). (E) Slab photo of
Lower Clear Fork (L 2.2) peloid grain-
dominated packstone (dolostone). Exxon
FCU 6229, depth: 2079 m (6822 ft).
Width of photo is 10 cm (4 in.).

Fusulinid Wackestone-packstone
Fusulinid-bearing rocks (Figure 14)
are found in all three formations
(Abo, Wichita, and Lower Clear Fork),
although they are very limited in the
Wichita. In all occurrences, they are
most commonly dolomitized. Fusu-
linids, which constitute as much as
40% of the rock are preserved either
as open (Figure 14C) or anhydrite-
filled (Figure 14A) molds or as well-
preserved fossil tests (Figure 14D).
In the Lower Clear Fork and Wichita,
they are associated with abundant
peloids (probably pellets). In the Abo,
they commonly co-occur with cri-
pellets. Skeletal grains in the form of fusulinids, mol- noids and less commonly with brachiopods. Fusu-
lusks, and crinoids are also common. In most cases, linids are thought to have occupied water depths of
these deposits, which are restricted to subtidal sections 30 m (100 ft) or more and here represent the deepest
of the Lower Clear Fork, are also well sorted, probably water facies observed at Fullerton field. Accordingly,
because of wave action. Grainstones, although relatively their presence is an indicator of platform flooding
uncommon, display excellent size sorting and, in some and relative sea level rise, making them key indicator
cases, possess inclined or cross-laminations (Figure 13B). facies of cycle and sequence boundaries.
These rocks, which occur as both dolostone and lime-
stone, represent the highest energy facies in the reservoir Skeletal Wackestone-packstone
succession and, in many cases, display the best porosity Typically, these rocks contain small volumes of skel-
and permeability. Pores are dominantly interparticle, etal debris (most commonly mollusk fragments, but
although moldic pores are also very abundant, especial- also including crinoids and less common ostracods)
ly in limestone-dominated intervals (Figure 13D). and the ubiquitous peloids. They are gradational into
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 367

FIGURE 12. Core photo (A) and image


log (B) images of burrowed peloid
packstone. Although low-resistivity
(black) features are commonly inter-
preted to be open vugs on image logs,
core photo reveals that vugs are ac-
tually burrow fills (light colored) of
higher porosity. Lower Clear Fork. Core:
Exxon FCU 6739, depth: 2123 m (6966 ft).
Image log: Exxon FCU 2564, depth:
2095 – 2096 m (6873 – 6877 ft). Core is
10 cm (4 in.) wide. Image log is 63.5 cm
(25 in.) wide.

a marker facies indicative of trans-


gression or platform deepening (i.e.,
sea level rise).

Siltstone-sandstone
Quartz silt- and sand-bearing rocks
are common only at the top of the
Lower Clear Fork and in the overly-
ing Tubb Formation. Because the
quartz is associated with potassium-
and thorium-rich clay minerals, its
presence in the Lower Clear Fork and
Tubb is well defined by high cor-
rected gamma-ray (CGR) wire-line-
peloid wackestones and packstones. Evidence of bur- log response (Figures 4, 5). Their maturity, size (fine
rowing is common. The dominance of mollusks and sand to coarse silt), sorting, and shape (generally sub-
the essential absence of more normal-marine organ- angular) suggest that these sediments were origi-
isms in these rocks suggest that they were deposited nally wind blown (Fischer and Sarntheirn, 1988).
in an inner-platform setting. Typically, they exhibit This indicates that most of the quartz in the section
low porosity; pore space is created by skeletal molds was delivered to the area during sea level lowstands
and intercrystalline pores. when the platform was emergent. Quartz silt-sand
occurs in two scenarios: (1) in peritidal tidal-flat fa-
Oncoid Wackestone-packstone cies and (2) in reworked subtidal facies. The former,
Oncoids (or oncolites) are large microbially coated which are probably formed as small volumes of silt,
grains (Figure 15) formed under conditions of fre- sand, and clay, are blown onto intermittently exposed,
quent wave agitation in shallow water. They are abun- slowly accumulating tidal flats and admixed with peri-
dant at the base of the Lower Clear Fork throughout tidal carbonate sediment. The presence of potassium
the entire field area. Invariably, they are associated and thorium in these clastics produces the increased
with fusulinids and other faunas indicative of open- gamma-ray response that is associated with many
marine deposition. This association and their strati- tidal-flat deposits in the Clear Fork and facilitates rec-
graphic position immediately above the top of the tidal- ognition of these cycle-capping deposits in these sedi-
flat–dominated Wichita indicate that they represent ments (Ruppel, 2002). Most of the occurrences of silt,
marine flooding of the platform during sea level rise sand, and clay in the Lower Clear Fork at Fullerton
(transgression). In some downdip wells (e.g., Pan are of this type.
American FM-1; Figure 3), they are common at Lower Subtidal silt, sand, and clay deposits are most com-
Clear Fork cycle bases. Their distribution suggests that mon in the Tubb. These rocks were formed first by
they document relatively high-energy conditions large-volume eolian deposition during extended sea
developed during platform flooding. Like the fusuli- level lowstand and carbonate nondeposition, then re-
nid facies, the oncoid wackestone-packstone facies is worked during the ensuing sea level rise and marine
368 / Ruppel and Jones

FIGURE 13. Core, thin-section, and


image log photographs of Lower Clear
Fork and Abo grainstones. (A) Slab
photo of Lower Clear Fork ooid grain-
stone showing cross-laminations. Exxon
FCU 6122, depth: 2109 m (6920 ft).
Core is 10 cm (4 in.) wide. (B) Image
log of cross-bedded grainstone. Lower
Clear Fork. Exxon FCU 2564, depth:
2134–2135 m (7002–7005 ft). Width of
image is 63.5 cm (25 in.). (C) Photo-
micrograph of Lower Clear Fork ooid
limestone-grainstone showing oomoldic
pores. Exxon FCU 5927, depth: 2107 m
(6913 ft). (D) Photomicrograph of
Lower Clear Fork ooid dolostone grain-
stone wackestone showing interparti-
cle and minor moldic pores. Exxon
FCU 4828, depth: 2175 m (7137 ft).

Depositional Model
The relative distribution of the
facies described above can best be
understood when considered in light
of a conceptual geological model.
Figure 16 portrays idealized 3-D re-
lationships among major facies types
and depositional environments typ-
ical of most middle Permian (Leo-
nardian and Guadalupian) platform
carbonate successions in the Perm-
ian basin. It should be understood
that this model reflects only the
flooding of the platform. These deposits are generally relative interrelationships among facies tracts. The
richer in clastic content, are intermixed with carbonate actual position of individual facies tracts at a given
mud, and show evidence of subtidal conditions (e.g., point in time is a function of many factors, including
burrows, stratification). Although these rocks may accommodation, rates and magnitude of sea level
locally exhibit some minor porosity, they do not ap- rise, rising versus falling sea level trends, climate, tec-
pear to display any significant reservoir permeability. tonics, etc. However, during normal relative rises in
sea level, facies tracts normally step landward,
whereas during falls, they step basinward. Exceptions
Lithoclast Wackestone to this general pattern are common. For example,
Thin intervals (commonly less than 0.3 m [1 ft]) ramp-crest facies may be much better developed dur-
containing scattered lithoclasts are locally encoun- ing sea level highstand and fall than during trans-
tered at the contacts between transgressive marine gression and rise. The basal Leonardian (L1) deposi-
facies and underlying tidal-flat deposits or other facies tional sequence at Fullerton, for example, appears
showing evidence of subaerial exposure (Figure 8D). to be dominated by an updip, inner-ramp succes-
Clasts are variable in composition but most com- sion (Wichita) and a downdip outer-platform succes-
monly consist of fragments or intraclasts derived from sion (Abo) with a very poorly developed middle-ramp
the underlying bed. Clasts typically are 1 cm (0.4 in.) to ramp-crest facies tract. Nevertheless, the vertical
in width or less. Although volumetrically minor, these successions (facies-stacking patterns) formed by sea
rocks are an important indicator facies of lithification level-driven migrations of these facies tracts are the
because of exposure or nondeposition and subsequent key to identifying depositional cycles in cycles from
sea level rise and, thus, of cycle boundaries. cores and logs.
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 369

FIGURE 14. Core, thin-section, and


image log photographs of the fusulinid
wackestone-packstone facies. (A) Slab
photo of fusulinid wackestone with open
and anhydrite-filled fusumoldic pores.
Exxon FCU 7322, depth: 2132 m (6996 ft).
Core is 10 cm (4 in.) wide. (B) Image
log of fusulinid wackestone-packstone
facies. Exxon FCU 2564, depth: 2131 –
2132 m (6990–6994 ft). Width of image
is 63.5 cm (25 in.). (C) Photomicrograph
of fusulinid wackestone showing open
fusumolds. Exxon FCU 7630, depth:
2066 m (6778 ft). (D) Photomicrograph
of fusulinid wackestone showing well-
preserved fusulinids but little porosity.
Exxon FCU 4828, depth: 2139 m (7018 ft).

The ooid-peloid grain-dominated


packstone-grainstone facies in the
Lower Clear Fork at Fullerton is typi-
cal of facies deposited in a platform
ramp crest at the platform margin
where relatively high wave energies
are developed (Figure 16). However,
none of the cored wells in the field
reveal the kind of vertical facies stack-
ing and amalgamation that typically
defines a well-developed ramp crest
(see, for example, Kerans et al., 1994;
Ruppel et al., 2000). This may reflect
insufficient core control or the ab-
sence of a well-developed ramp crest
on the Lower Clear Fork platform. As
Among the identified facies at Fullerton, the peri- previously stated, the Abo also appears to lack a ramp-
tidal mudstone-wackestone facies, the clay-rich carbon- crest facies succession.
ate mudstone facies, and the exposed tidal-flat facies, As in most middle Permian carbonate-platform
for the most part, represent deposition on the inner successions, the fusulinid wackestone-packstone fa-
ramp in a sabkha-tidal-flat setting as low-exposure cies documents outer-platform deposition. Such rocks
index (cf. Hardie and Garrett, 1977) peritidal deposits, are well developed in both the Lower Clear Fork and
tidal-flat ponds, and high-exposure tidal-flat deposits, the Abo. In the Abo, these rocks display classic clino-
respectively (Figure 16). It should be noted, however, formal bedding and the general lack of cyclicity that
that exposure fabrics, like those developed in the inner typifies a distal outer-platform setting (Figure 16). By
platform, can also form on the ramp crest following contrast, fusulinid wackestone-packstone facies rocks
complete aggradation and exposure at any time. in the Lower Clear Fork are horizontally bedded and
The peloid- and pellet-rich facies at Fullerton (peloid interbedded with middle- and inner-platform facies.
wackestone facies, peloid packstone facies, peloid grain- This suggests that these fusulinid-rich rocks were de-
dominated packstone facies, and skeletal wackestone- posited in a much more proximal outer-ramp position.
packstone facies) are also dominantly associated with The oncoid wackestone-packstone facies is closely
the inner ramp but occupy a somewhat more distal, associated with the fusulinid wackestone-packstone
lagoonal to restricted subtidal setting, where sedi- facies in the Lower Clear Fork at Fullerton, indicating
ment formation is dominated by infaunal and epi- that the former was deposited during platform deep-
faunal burrowing activities (Figure 16). ening. This facies has not previously been reported
370 / Ruppel and Jones

FIGURE 15. Core photographs of


oncoid, wackestone-packstone facies.
(A) Small, algally coated fusulinids. Low-
er Clear Fork, Exxon FCU 5927, depth:
2128 m (6980 ft). (B) Large oncoids.
Lower Clear Fork, Pan American FM-1,
depth: 2182 m (7160 ft). Cores are 10 cm
(4 in.) wide.

Sequence Stratigraphy
Figure 17 illustrates diagrammat-
ically the sequence architecture,
basic facies tracts interrelationships,
and rock and time terminology of
the lower Leonardian at Fullerton
field. In this section, we describe the
sequence-stratigraphic characteris-
tics of each of the major Leonardian
units at Fullerton.

Abo Formation
As already discussed, the Abo rep-
resents the distal facies or systems
tract of the earliest (oldest) Leonar-
from middle Permian rocks in the Permian basin. Its dian sequence in the Permian basin: the L1 sequence
occurrence at Fullerton field may indicate the de- (Figures 4, 5, 17). Regionally, the Abo consists of outer
velopment of relatively higher energy conditions on ramp to slope, skeletal crinoid-fusulinid-dominated
the outer platform in the Fullerton area than is gener- subtidal facies and, less commonly, ramp-crest ooid-
ally developed in other regions in the Permian basin. peloidal grainstones (Kerans et al., 2000). In both

FIGURE 16. Depositional model for Permian shallow-water carbonate platforms in the Permian basin. This model is
applicable to most Leonardian and Guadalupian carbonate platform successions, including the reservoir succession at
Fullerton field. Modified from Kerans and Ruppel (1994).
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 371

FIGURE 17. Generalized sequence-stratigraphic model of the lower Leonardian succession at Fullerton field showing
sequence boundaries, primary facies tracts, and stratigraphic nomenclature.

outcrop and the subsurface, the Abo is dominantly apparent either from core sections or from wire-line
characterized by clinoformal, top-lapping geometries logs. In part, this is the result of limited core and well-
(Kerans et al., 2000; Zeng and Kerans, 2003). This log control through the Abo section. But poor cor-
typical Abo architecture is very apparent in 3-D and relatability is typical of outer-platform depositional
2-D seismic data from Fullerton field, where the Abo successions that display clinoformal architectures
is readily defined by its clinoformal reflectors, and its because of the associated dipping bedding surfaces,
top is defined by a prominent toplap surface that is discontinuous facies packages, and poor vertical facies
apparent throughout much of the field (Figure 6). contrasts.
Although the Abo at Fullerton field is penetrated Abo rocks are dominantly dolostones at Fullerton
by relatively few wells and has been cored in even field. Porosity is locally very high, reaching values of
fewer wells, existing cores display facies and bedding as much as 25%. Pore types vary among fusumolds,
characteristics typical of sediments deposited in an intercrystalline pores in coarse-crystalline dolostones,
outer ramp-slope setting (such as open-marine fauna, and interparticle pores in skeletal-peloidal packstones.
inclined bedding, graded beds, slump features, con- The thickness of the Abo is indeterminate largely
glomerates, and lack of shallow-water sediments). for two reasons. First, few wells and no cores pen-
Facies consists of alternating beds of fusulinid-crinoid etrate the complete Abo to Wolfcamp section. Second,
packstones and wackestones and peloidal pack- outcrop studies reveal that the Abo – Wolfcamp con-
stones. No cyclicity is apparent in the Abo, although tact may be lithologically indistinct in many areas:
alternations between skeletal-rich and skeletal-poor in outcrop, both the uppermost Wolfcamp and the
intervals are evident. Inclined beds are locally com- Abo are composed of clinoformal, outer-platform
mon. Additionally, correlations in the Abo are not fusulinid-crinoid wackestones (Kerans et al., 2000).
372 / Ruppel and Jones

A minimum thickness of 90 m (300 ft) is established karsted contact between the outer-ramp subtidal
by the Pan American FM-1 cored well in the south- deposits of the Abo and peritidal tidal-flat rocks of
eastern corner of the field (Figures 3, 5). the overlying Wichita.
The contact between the Abo and the overlying
Wichita varies from sharp to gradational. In some
cores, the contact is marked by an interval (as much as Lower Clear Fork Formation
several feet thick) of intermixed clasts and fragments Rocks assigned to the Lower Clear Fork Formation
of Wichita tidal-flat deposits and Abo fusulinid- represent the subtidal systems tract of the Leonardian
bearing subtidal deposits. The brecciated nature of 2 (L2) sequence. Core studies show that the Lower
this contact interval suggests that it is the result of Clear Fork can be subdivided into three HFSs through-
karst-related dissolution and collapse. These breccias out most of the Fullerton field area: L2.1, L2.2, L2.3
and their distribution are discussed more fully in a (Figures 4, 5). Each of these HFSs can be identified as
later section of this chapter. having a transgressive, dominantly subtidal lower leg
and an overlying, highstand upper leg.

Wichita Formation High-frequency Sequence L2.0


The Wichita at Fullerton field consists of a diverse High-frequency sequence L2.0 documents the
assemblage of peritidal to supratidal tidal-flat depos- initial flooding of the platform following sea level
its. Integration of outcrop and subsurface data sug- fall and exposure at the end of L1 deposition. In most
gests that the Wichita Formation is the updip, proxi- of the field area, HFS 2.0 consists of amalgamated
mal facies equivalent of both the Abo and the Lower tidal-flat deposits of the upper Wichita. Subtidal Lower
Clear Fork subtidal successions. These data also imply Clear Fork deposits of L2.0 are only present at the
that the Wichita actually comprises parts of two depo- margins of the field (Figures 4, 5).
sitional sequences: the highstand leg of sequence L1
and the transgressive leg of sequence L2 (Figure 17). High-frequency Sequence L2.1
The lower Wichita represents the updip, tidal-flat fa- In most of the Fullerton field area, HFS L2.1 forms
cies tract equivalent of the downdip, outer-platform the base of the Lower Clear Fork. As defined for this
facies tract of the Abo in Leonardian sequence L1, study, L2.1 consists of a basal section of transgressive
whereas the upper Wichita represents the updip tidal- to early highstand subtidal platform facies and an
flat facies equivalent of the basal Lower Clear Fork upper section of highstand tidal-flat facies (Figures 4,
subtidal facies in sequence L2 (HFS 2.0; Figures 5, 17). 5, 17). The basal transgressive subtidal facies of the
The Wichita is dominantly composed of dolostone; Lower Clear Fork represent the first marine flooding
however, intervals of limestone are common in the of the platform and a sharp change in depositional
upper Wichita in the northern part of the field. The style from the tidal-flat deposition of the Wichita to
limestone intervals are relatively persistent laterally subtidal deposition of the Lower Clear Fork.
and are generally subparallel to stratigraphic markers The basal transgressive leg of L2.1 consists domi-
in the field (Figure 18a). nantly of fusulinid wackestone-packstone and oncoid
The thickness of the Wichita is relatively consis- wackestone-packstone facies that document landward
tent throughout the northern two thirds of the field, backstepping of outer platform facies across the plat-
ranging from about 75 to 105 m (280 to 350 ft), but form. These rocks are generally overlain by a succes-
thickness decreases markedly to as little as 34 m (110 ft) sion of peloid packstones and grain-dominated pack-
in the southeastern part of the field (Figure 19). This stones that represent late transgression and early
decrease in thickness, which occurs relatively abruptly highstand. Locally, within this succession, there are
along a generally northeast-trending belt, marks the exposed tidal-flat facies indicating periodic exposure.
platform-margin boundary between updip tidal-flat L2.1 is capped throughout most of the field by a suc-
Wichita facies and downdip outer-ramp Abo fusulinid cession of one to three tidal-flat-capped cycles that rep-
facies (Figure 5). resent exposure during late highstand (Figures 5, 17).
The position of the L1-L2 boundary is difficult to Tidal-flat facies in some cases exhibit an elevated
precisely place within the thick succession of Wichita gamma-ray-log response that aids in their recogni-
peritidal deposits. The contact shown in Figure 5 is tion (Figures 4, 5).
extrapolated from downdip core and well control, Like most of the Lower Clear Fork in the Permian
where the L1-L2 contact is well defined by a sharp, basin, HFS 2.1 is dominantly dolostone. However,
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 373

FIGURE 18. Distribution of limestone and dolostone in Wichita and Lower Clear Fork based on cores and wire-line logs.
(a) Cross section (BB0 ) across the northern part of the field showing cycle stratigraphy and mineralogy. (b) Maps
showing limestone in Wichita HFS L2.0, Lower Clear Fork HFS L2.1, and Lower Clear Fork HFS L2.2.

limestone is locally common at Fullerton field, es- differential subsidence along deep-seated faults (Jones
pecially in the transgressive leg of 2.1. Areally, lime- and Ruppel, 2004). This is especially apparent along
stone is most abundant in an arcuate belt through the north-trending fault in the northeastern part of
the field (Figure 18b). the field, where the thickness changes from less than
The thickness of L2.1 is relatively constant across 43 m (140 ft) on the western, upthrown side of the
the Fullerton field area, ranging from about 43 to 46 m fault to more than 49 m (160 ft) on the eastern, down-
(140 to 150 ft) across most of the area. Good evidence thrown side of the fault. Examination of modern struc-
shows that local thickness changes are a function of ture (Figure 3) shows that considerable movement
374 / Ruppel and Jones

High-frequency Sequence L2.2


High-frequency sequence L2.2 is similar to HFS
L2.1 in consisting of a basal transgressive leg com-
posed of backstepping tidal-flat facies, a middle (late
transgressive to early highstand) leg composed domi-
nantly of subtidal facies, and an uppermost (late high-
stand) leg composed of tidal-flat facies (Figures 5, 17).
As is the case with L2.1, there is good evidence that the
marine flooding of the platform following the post-
L2.1 lowstand was progressive. Basal L2.2 tidal-flat
deposits are thickest in the center of the field area but
generally absent along the margins, reflecting greater
accommodation and early flooding of downdip areas.
The lower abundance of outer-ramp fusulinid-rich
facies in the TST of L2.2 relative to L2.1 suggests that
the overall accommodation during the L2.2 sea level
rise was somewhat less than during L2.1. Lower ac-
commodation and attendant lower wave energies are
also suggested by the near absence of the oncoid
wackestone-packstone facies in L2.2. These rocks are
found only in the most downdip core in the field. The
fusulinid wackestone-packstone facies is also largely
restricted to the eastern and southeastern parts of the
area (Figures 5, 17). Throughout most of the field
area, L2.2 is dominated by peloid wackestones and
packstones typical of middle-platform deposition.
The sequence is capped by a thin succession of tidal-
flat cycles (Figures 5, 17).
High-frequency sequence L2.2 is composed dom-
inantly of dolostone, with two important exceptions.
Like L2.1, limestone is dominant in the southern part
of the field (Figure 18b). Dolomite in this area is
largely restricted to mud-rich fusulinid wackestones;
virtually all grain-rich facies in this area are lime-
stone. Limestone is also abundant in a small area in
FIGURE 19. Thickness of the Wichita Formation in the
Fullerton field area. Marked thinning in southeast part the north-central part of the field.
of the field corresponds to facies change from Wichita The thickness of L2.2 is generally significantly less
to Abo. than that of L2.1. L2.2 ranges from about 26 to 27 m
(85 to 90 ft) across most of the central part of the area
to about 100 ft (30 m) along the northern and
on this fault occurred even after Lower Clear Fork southern field margins. Porosity development in L2.2
deposition. is similar to that in L2.1. Porosity is greatest in the
Porosity is most abundant in the subtidal (trans- subtidal (transgressive and early highstand) parts of
gressive and early highstand) parts of the sequence. the sequence and dominated by intercrystalline and
Pore space is dominated by intercrystalline and moldic moldic pores. In addition, like L2.1, porosity is great-
pores, the latter being especially abundant in lime- est in sections containing significant limestone.
stone sections. Overall, these rocks are most porous
in areas where limestone is present. Tidal-flat rocks High-frequency Sequence L2.3
are also locally porous but generally contain low The uppermost Lower Clear Fork HFS (L2.3) is com-
permeabilities. The subtidal rocks of L2.1 probably posed of tidal-flat-capped restricted subtidal cycles
constitute the most productive reservoir interval in throughout most of the field area (Figures 5, 17).
the field. These tidal-flat caps typically contain fenestral and
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 375

fine intercrystalline pore space that is definable on ships among facies to define cycles, (3) the nonuni-
porosity logs. Log correlations suggest that these cycle- form response of wire-line logs to these facies, and
capping tidal-flat facies are relatively continuous across (4) the overprinting effects of diagenesis. Internal,
significant areas of the field. The underlying cycle- cycle-scale correlations are possible, however, in the
base subtidal rocks are dominantly mud-rich pack- upper Wichita in the northern part of the field, where
stones and wackestones. Porosity, as indicated, is limestone beds are relatively correlative. Sedimentary
mostly restricted to tidal-flat caps. Because these rocks structures suggest that these limestones represent un-
are dominated by fine intercrystalline and moldic dolomitized bases of cycles whose tops have been do-
pores, they contain little, if any, reservoir permeabil- lomitized by early diagenesis. The Wichita cycles de-
ity. Accordingly, locally they contain oil stain and fined by these limestone-dolostone couplets average
rarely contribute to oil production. 6 m (20 ft) in thickness.

Tubb Formation Lower Clear Fork Cyclicity: HFS L2.1


The Tubb is characterized by fine-grained silic- Facies-stacking patterns in HFS 2.1 define numer-
iclastics (coarse siltstone and fine sandstone), which ous apparent cycle tops at thicknesses ranging from
are relatively easily defined by high-gamma-ray-log 0.6 to 4.6 m (2 to 15 ft). Cycles are typically char-
response (Figures 4, 5). These clastics are interpreted acterized by mud-rich facies at their bases and grain-
to represent eolian deposits that were deposited during rich facies at their tops. Basal TST cycles in HFS L2.1
the post-L2 lowstand and then reworked during the contain abundant oncoids at cycle bases, along with
ensuing L3 sea level rise (Kerans et al., 2000; Ruppel accompanying fusulinids (Figure 20). These cycles
et al., 2000; Ruppel, 2002). Most intervals are varying are capped by better sorted, peloid-rich facies. High-
mixtures of siltstone-sandstone and carbonate (typ- frequency cycles (typically 1.5–3 m [5–10 ft] in thick-
ically mud-rich, shallow-water facies). The Tubb ness) stack into cycle sets that average 7.5 – 12 m (25 –
locally contributes to oil production in the Permian 40 ft) in thickness (Figure 20). Cycle sets display sim-
Basin but is not part of the reservoir at Fullerton field. ilar facies-stacking patterns to cycles consisting of
fusulinid-rich bases and peloid-rich caps. Locally, tidal-
flat facies cap these cycle sets. Porosity is generally
Cycle-scale Stratigraphy
The fundamental goal of cycle stratigraphy is to
develop a correlation framework based on time-
equivalent surfaces. The basic underlying premise
for this approach is the assumption that widely
correlative depositional cycles are formed by punc-
tuated, allocyclic processes (e.g., sea level rise and
fall) that affected sedimentation over broad areas.
The methodology used to develop a cycle-scale strati-
graphic framework in the Leonardian section at Ful-
lerton field consists of the following: (1) characteriza-
tion of facies-stacking patterns and cycle development
in analogous outcrops; (2) description, interpretation,
and logging of facies, stacking patterns, and possible
cycle tops in cores; (3) integrated log- and core-based
correlation of tentative cycle tops; and (4) definition
of cycle architecture. At Fullerton field, nearly 4570 m
(15,000 ft) of core (from 29 cored wells) was described
in detail to provide the basic data for cycle definition.

Wichita Cyclicity FIGURE 20. Facies stacking and cycle development in


the fusulinid- and oncoid-rich, transgressive systems
Correlations in the thick Wichita succession of
tract of Lower Clear Fork HFS L2.1. Note that porosity
relatively similar tidal-flat deposits are difficult be- is typically developed at or near cycle tops. Note also
cause of (1) the discontinuous nature of tidal-flat facies, the lack of any systematic relationship between gamma-
(2) the lack of systematic vertical stacking relation- ray log and facies or cyclicity.
376 / Ruppel and Jones

facies. Cycles typically average 1.5 – 3 m (5 – 10 ft) in


thickness; cycle sets are commonly 6 – 9 m (20 – 30 ft)
thick (Figure 21).

Lower Clear Fork Cyclicity: HFS L 2.2


Facies stacking and cycle development in HFS
L2.2 are very similar to those in L2.1. Cycles average
1.5– 3 m (5 – 10 ft) in thickness, and porosity is best
developed in cycle tops (Figure 22). L2.1 cycles dif-
fer, however, in the lack of oncoid facies and the
relative scarcity of fusulinid facies. This presumably
reflects decreasing overall accommodation in L2.2
because of platform aggradation and slowing rates
of long-term sea level rise. Additionally, cycle sets
FIGURE 21. Facies stacking and cycle development in are not generally definable in L2.2. This is largely
the grain-rich, highstand systems tract of Lower Clear caused by the general absence of fusulinid facies that,
Fork HFS L2.1. Note that porosity is typically developed in L2.1, define these intermediate-scale sea level-rise
at or near cycle tops. As with other Lower Clear Fork events.
successions, gamma-ray logs do not show a systematic
response to facies or cyclicity. Lower Clear Fork Cyclicity: HFS L2.3
High-frequency cyclicity in the Leonardian at Ful-
highest at both cycle tops and cycle-set tops relative lerton field is most readily definable in L2.3. These
to bases. However, because cycles in the lower parts of rocks, which are characterized by tidal-flat-capped
cycle sets are generally more mud and fusulinid rich, shallow subtidal cycles (Figure 23), appear to be much
overall porosity in these basal TST cycles is commonly more widely correlative than cycles in L2.1 or L2.2.
relatively low. Although neither facies nor cyclicity is defined by
Cores and outcrop studies suggest that Clear Fork gamma-ray logs, both are distinguishable on porosity
platform cycles and cycle sets are correlative across sig- logs because of the typically well-developed porosity
nificant distances. However, where cores are not avail- associated with cycle-capping tidal-flat deposits. The
able, these correlations can be difficult to establish. porosity in these rocks is generally caused by the
Gamma-ray logs display virtually no systematic re- presence of fenestral pores and can be relatively high,
sponse to subtidal facies and cycles (Figures 20, 21) especially relative to the generally low porosity ex-
and, thus, cannot be reliably used for cycle-scale cor- hibited by cycle-base mud-rich wackestones and
relation throughout most of the Lower Clear Fork. In
the absence of gamma-ray logs, the most effective
way of establishing cycle-scale correlations is through
the use of porosity logs. This approach is based on the
observation from outcrops and cores that cycle tops
consistently contain the most grain-rich facies, and
that these rocks are most likely to contain high po-
rosity. Accordingly, porosity logs can be used to cor-
relate both the high-porosity facies in the upper part
of the cycle and the overlying cycle top.
L2.1 highstand cycles are dominated by mud-rich
peloidal facies at their bases and grain-rich peloid-
or ooid-bearing facies at their tops (Figure 21).
Fusulinids are commonly sparse in these highstand
cycles, reflecting the basinward shift in facies tracts. FIGURE 22. Facies stacking and cycle development in
the grain-rich, late transgressive systems tract and early
In general, these cycles are dominated by peloid
highstand systems tract of Lower Clear Fork HFS L2.2.
packstones and grain-rich packstones. Porosity is Note that porosity is typically developed at or near cycle
commonly highest at both cycle tops and cycle-set tops. Note here again that gamma-ray logs do not display
tops because of the abundance of these grain-rich any relationship to facies or cycle development.
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 377

ern part of the field (Figure 18). In all cases, calcite-


rich rocks in the Wichita are peritidal mudstones or
wackestones (the peritidal mudstone-wackestone
facies). These calcite-dominated facies characteristi-
cally exhibit very low porosity (<2%) and essentially
no permeability. Log correlations suggest that lime-
stone intervals are locally correlative (Figure 18a).
Facies-stacking patterns suggest a systematic relation-
ship between mineralogy and cyclicity, and it seems
probable that these limestones are the result of cycle-
FIGURE 23. Facies stacking and cycle development in punctuated diagenesis: limestones representing un-
the nonreservoir Lower Clear Fork HFS L2.3. Porosity is dolomitized cycle bases. Limestone is virtually absent
typically developed in cycle-capping tidal-flat facies, but from the lower Wichita (L1 highstand deposits) but very
little or no permeability is associated with the fenestral
common in the upper Wichita (L2 TST) (Figure 18a).
pores that dominate these caps.
Limestone abundance in the upper (L2) Wichita dis-
plays a systematic trend across the field. The highest
packstones. However, because of the separate-vug abundance of limestone is in the northwestern part
fenestral pores, permeability is generally low, and of the field; essentially, no limestone is present in the
L2.3 sparsely contributes to hydrocarbon production southern part of the field (Figure 18b). Because of the
in the field. Porosity-based wire-line correlations sug- low porosity associated with limestones in the Wichita,
gest that cycle and facies continuity is high. This is porosity logs can be reliably used to define and cor-
somewhat unexpected, considering outcrop observa- relate limestone even where cores, dual porosity, or
tions that suggest that tidal-flat facies are highly dis- photoelectric effect (PE) logs are unavailable.
continuous. The laterally continuous porosity develop- Lower Clear Fork rocks at Fullerton field also con-
ment at these cycle tops probably reflects the effects tain locally abundant limestone (Figure 18a). In HFS
of early diagenesis associated with sea level fall and L2.1, limestone is especially common in a sinuous
subsequent rise at these surfaces. belt that generally follows the margin of the outer
platform (Figure 18b). Essentially, all grain-rich facies
MINERALOGY AND DIAGENESIS in this belt are dominantly calcitic.
Limestone also dominates L2.2 in the southern
The reservoir section at Fullerton field is similar to part of the field. However, outside of this area, lime-
most other platform Leonardian successions in the stone is uncommon, except in a small area in the
Permian basin in showing evidence of significant northwestern part of the field (Figure 18b). Porosity
postdepositional diagenesis. Principal products of this is generally high in limestone-rich intervals in the
diagenesis are matrix-replacive and pore-filling dolo- subtidal Lower Clear Fork. However, porosity is also
mite and anhydrite. However, limestone is locally pres- well developed in dolostones, so except where cores,
ent in the Leonardian section at Fullerton, including dual-porosity log suites, or PE logs are available, it is
the Abo, the Wichita, and the Lower Clear Fork. not possible to differentiate limestone from dolo-
stone in the Lower Clear Fork.
Dolomite and Limestone Distribution
Dolomite is, by far, the dominant mineral in the Stable Isotope Chemistry
reservoir section. The Abo consists entirely of dolo- Stable isotopes can provide insights into the
mite except in the most downdip wells. In the FM-1 timing and spatial distribution of diagenetic events.
well, perhaps the most downdip well in the field, the Because diagenesis is a key factor in porosity preser-
Abo contains alternating zones of limestone and vation and loss, such data can commonly provide in-
dolostone. Dolostone is more commonly associated sights into causes and distribution of porosity trends.
with mud-rich facies (peloid wackestone and fusuli- Stable isotope data collected during this study and
nid wackestone) in the Abo clinoformal succession, those previously gathered by J. Kaufman (1991, per-
whereas limestone intervals are more commonly sonal communication) reveal similarities to other
grain-rich, skeletal facies. Leonardian successions in the Permian basin but also
Limestone is locally common in the Wichita, espe- appear to define spatially and temporally distinct
cially in the upper half of the formation in the north- patterns of diagenesis (Figure 24).
378 / Ruppel and Jones

data reported from the Guadalupian of the Permian


basin (Ruppel and Cander, 1988a, b; Leary and Vogt,
1990; Saller and Henderson, 1998) but similar to
most Leonardian values (Saller and Henderson, 1998;
Ruppel, 2002).
The d18O data obtained from Lower Clear Fork dolo-
stones range from 2.96 to +2.97% PDB (Figure 24).
With the exception of the lightest data, most of these
values are similar to Leonardian data reported by
Ye and Mazzullo (1993), Saller and Henderson (1998),
and Ruppel (2002), but very different from data re-
ported for Guadalupian rocks. Typical d18O values for
Guadalupian (San Andres and Grayburg formations)
platform dolomites are 3 – 6% (Vogt, 1986; Ruppel
and Cander, 1988a, b; Saller and Henderson, 1998;
Ruppel and Bebout, 2001; Ruppel, 2002). The rela-
FIGURE 24. Stable isotope data for the Tubb, Wichita, tively enriched isotopic signatures of Guadalupian
Lower Clear Fork, and Abo at Fullerton field. Solid red dolomites have invariably been interpreted to have
circles and blue triangles are dolomite and calcite samples, been produced during dolomitization by evapora-
respectively, from J. Kaufman (1991, personal communi- tively concentrated seawater brines (Bein and Land,
cation). Open green circles and green triangles are do- 1982; Bebout et al., 1987; Ruppel and Cander, 1988a, b).
lomite and calcite samples, respectively, collected during
Taken collectively, the relatively depleted values for
this study.
the Leonardian dolostones at Fullerton and in other
Leonardian fields suggest that either (1) dolomitiza-
Dolostone samples from the Tubb, Lower Clear tion was caused by brines that were less evaporatively
Fork, Wichita, and Abo at Fullerton field exhibit d13C concentrated than those that caused Guadalupian
values ranging from 1.41 to 5.54% Peedee belemnite dolomitization; (2) dolomitization was caused by fluids
(PDB) (n = 50). These values are generally lighter than of a mixed water origin; or (3) a combination of cases 1
and 2 above. Systematic spa-
tial variations in d18O and
d13C values of dolomites in
the Wichita and Lower Clear
Fork suggest an even more
complex history of dolomi-
tization and diagenesis.
Stable isotope values from
upper Wichita peritidal do-
lomites, for example, define
two spatially distinct trends
(Figure 25a). d18O and d13C
values from high-porosity do-
lomites along the other mar-
gins of the Wichita tidal flat
display relatively light values
(average: 0.0% d18O, +2.6%

FIGURE 25. Maps showing


trends of facies, porosity, and
stable isotope data in the upper
Wichita L1-L2 tidal-flat succes-
sion (a) and the Lower Clear
Fork L2 subtidal succession (b).
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 379

d13C). By contrast, values from low-porosity Wichita related diagenesis and dissolution is common in the
peritidal dolomites elsewhere in the field are distinct- lower part of the reservoir section at Fullerton field;
ly heavier, especially with respect to d18O (average: karst features occur within the Wichita and in the top
+2.8% d18O, +3.0% d13C). These marked differences of the Abo. These rocks are variable in fabric but
suggest that two mechanisms of dolomitization were include four basic types: (1) polymict conglomerates,
involved in the diagenesis of Wichita dolomites. Iso- (2) monomict breccias, (3) fractured and tilted beds,
topically light d18O values along the arcuate belt of and (4) void-filling cement.
higher porosity tidal-flat facies are consistent with The predominant style of polymict conglomerate
early, seawater-dominated dolomitization. The heavier typically consists of rounded clasts of multiple peri-
d18O values exhibited by updip dolomites are typical tidal lithologies (Figure 26). Clasts range in size from
of dolomitization by refluxing brines. Similar heavy a few millimeters to several centimeters in maximum
d18O values are encountered in overlying Lower Clear dimension and are commonly subequant and rounded.
Fork dolomites (average: +2.5% d18O). This isotopic Clasts are commonly enclosed in mudstone or abut
similarity suggests that dolomitization of nonporous one another at stylolitic contacts. Polymict fabrics are
Wichita tidal-flat facies may have been caused by most common in the middle of the Wichita (Figure 26).
evaporatively concentrated brines generated during Intervals of polymict conglomerate of at least 7.5 to as
Lower Clear Fork time. much as 18 m (25 to as much as 60 ft) thick are present
Limestone isotope data from Fullerton come en- in the Exxon FCU (Fullerton Clear Fork Unit) 6122
tirely from the Lower Clear Fork. Values of d13C range core. The multiple facies character of these clasts and
from +1.61 to +5.46% PDB (n = 30), whereas d18O their rounded character indicate that they were formed
data range from 3.89 to 1.55% PDB (Figure 24). by sediment transport. Their discontinuous nature
These data are very different from previously reported and their association with other features indicative of
values for other calcite-bearing samples in the Leo- karst processes suggest that they originated as cave-
nardian and Guadalupian. Most reported d18O values fill deposits.
for Guadalupian calcite have ranged from about 7.6 Although probably not true polymict conglom-
to 10.4 (Leary, 1985; Vogt, 1986; S. Ruppel, 1996, erates, superficially similar deposits that also indi-
unpublished data from the Grayburg Formation at cate karst processes are present at the contact of the
South Cowden field). These values, however, were re- Wichita and the Abo at the downdip edge of the field
corded from replacement calcite interpreted to have area. These rocks are commonly characterized by a
precipitated from meteoric water possibly sourced mixture of two or more facies types, including Wichita
from deep basin fluids. The d13C values from these tidal-flat facies, green silty carbonate, dark-gray silty
replacement calcites are also very depleted (19 to carbonate, and Abo subtidal facies. Instead of being
30% PDB), suggesting precipitation from bacteri- interbedded with one another, the first three sur-
ally mediated sulfate reduction (Leary, 1985; Vogt, round the Abo facies in many cores. This suggests
1986; Ruppel and Cander, 1988b). The more normal that these polymict conglomerates actually represent
d13C values for the Lower Clear Fork limestones at dissolution and/or erosion of the top of the Abo and
Fullerton indicate a very different origin for these subsequent infilling of the irregular surface or differ-
calcites. Lower Clear Fork d18O values are very similar ential compaction of the Abo (L1) and overlying trans-
to the current best estimate for marine calcite pre- gressive Wichita sediments (L2) at the Abo–Wichita
cipitates from seawater during the middle Permian contact. In rare instances, there are also features sug-
(2.8% PDB; Given and Lohmann, 1985; Lohman gestive of collapse brecciation. All of these features
and Walker, 1989). This suggests that Lower Clear have been observed in analogous outcrops of the
Fork limestones contain a preserved record of original Abo –Wichita contact in Apache Canyon in the Sierra
seawater chemistry and by extension that these rocks Diablo (Kerans et al., 2000). The outcrop succession
have undergone relatively little chemical alteration. reveals that karst features (sinkholes and caves) were
formed during post-L1 sea level fall and then filled
Karst Development with transgressive L2 tidal-flat deposits (i.e., Wichita
facies). These tidal-flat facies, in some instances, were
Karst Fabrics brecciated and intermixed with the underlying Abo
Karst diagenesis can be a major factor in reser- in sinkholes and collapsed caves.
voir heterogeneity and compartmentalization in car- Monomict breccias or conglomerates consist of
bonate successions (Loucks, 1999). Evidence of karst- broken and rotated clasts of constant lithology and
380 / Ruppel and Jones

FIGURE 26. Core slab and image log


photographs of polymict conglomer-
ates in the Wichita Formation of prob-
able karst origin. (A) Exxon FCU 6122,
depth: 2195 m (7201 ft). (B) Amoco
University Consolidated IV-25, depth:
2199 m (7213 ft). Cores 10 cm (4 in.)
wide. (C) Image log of conglomerate in
the Wichita. Exxon FCU 2564, depth:
2186– 2187 m (7171– 7174 ft). Image is
63.5 cm (25 in.) wide.

Fractured and tilted beds are also


observed in some cores, especially in
the middle of the Wichita section.
Tilted beds have clearly been formed
by postdepositional collapse. In the
Exxon FCU 5927 core, they form part
of a succession that very much re-
sembles the classic cave fill-cave roof
succession described by Loucks (1999).
The succession in the Exxon FCU
5927 well consists of 6 m (20 ft) of
polymict cave-fill conglomerate com-
posed of mixed tidal-flat facies over-
lain by 20 ft (6 m) of fractured and
locally tilted but apparently gener-
ally in-situ beds of tidal-flat and
subtidal facies probably representa-
tive of a cave roof.
Large zones of void-filling anhy-
drite cement are also strong indica-
tors of karst-related dissolution. Zones
of massive anhydrite as much as
0.5 m (1.5 ft) thick are present in
cores in the Wichita at Fullerton field
(e.g., Exxon FCU 6122, 5927). Smaller
anhydrite-filled voids and fractures
are ubiquitous in the Wichita and at
facies (Figure 27). These rocks are restricted to the the Wichita–Abo contact and also point to late ce-
Wichita, commonly the middle of the section. Com- mentation of dissolution voids by diagenetic fluids as-
monly associated with these deposits are fractures, sociated with reflux dolomitization of the succession.
sediment infill (cracks and fissures), void-filling ce-
ment (chiefly anhydrite), and other evidences of dis- Causes and Timing of Karst
solution. Most of these features can also be formed by Outcrop studies demonstrate that major karsting
nonkarst processes in tidal-flat intervals not exposed of the Leonardian sequence occurred at the sea level
to true karst (e.g., tepee formation is commonly ac- fall and rise event that is defined by the L1-L2 se-
companied by broken and rotated blocks, cracks and quence boundary (Kerans et al., 2000). In downdip
fissures, and cement and sediment infill). Thus, it is areas of Fullerton field, cores demonstrate this same
possible that some of these deposits may not be the relationship. Here, the top surface of the Abo outer
result of true karst processes. However, their thickness, platform facies succession (L1) is karsted and is in-
abundance, and association with other features of karst filled and overlain by brecciated Wichita tidal-flat
formation suggest that many are also karst related. facies (L2). In updip areas, however, establishing a
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 381

FIGURE 27. Core slab photos of mo-


nomict breccias from the Wichita.
(A) Rotated clasts of tidal-flat facies. Exxon
FCU 5927, depth: 2203 m (7230 ft).
(B) Tilted blocks of laminated peritidal
facies. Exxon FCU 5927, depth: 2199 m
(7215 ft). (C) Brecciated clasts of peloidal
wackestone. Amoco University Consolidat-
ed V 15, depth: 2103 m (6899 ft). (D) Clasts
of peritidal mudstone-wackestone. Exxon
FCU 5927, depth: 2204 m (7233 ft).
Cores are 10 cm (4 in.) wide.

top L1 surface (consisting of Abo sub-


tidal sediments downdip and lower
Wichita tidal-flat sediments updip)
developed local caves, sinkholes, and
an irregular topography. During the
subsequent L2 sea level rise, trans-
gressive peritidal deposits of the Wi-
chita filled the irregular karsted sur-
face, including sinkholes and caves.
With continued sedimentation and
compaction, parts of the overlying
L2 (upper Wichita tidal flat) suc-
cession underwent local breccia-
tion and collapse probably because
of stress differences set up over un-
derlying karst features. This scenario
is consistent with outcrops (Kerans
et al., 2000) and with the distribution
of karst features seen at Fullerton
field.

Impact of Karsting on
Reservoir Quality
Some karst-related deposits at Ful-
lerton most certainly exhibit at least
local differences in petrophysical
properties (i.e., porosity, permeabil-
spatial and temporal relationship between karst for- ity, and saturation) from surrounding undisturbed
mation and the L1-L2 sequence boundary is more and unaltered deposits. Polymict conglomerates and
problematic. Most karst features in the Fullerton field anhydrite voids are two obvious examples. However,
area are found in an interval of about 45 m (150 ft) in two factors make quantification of the importance of
the middle of the Wichita section. In general, this these differences difficult. First, most karst deposits
interval correlates approximately to the interpreted do not record significantly different porosity or perme-
L1-L2 boundary (Figure 5). However, karst features ability from that of surrounding nonkarsted deposits
are developed both below and above this horizon, on both wire-line and core data. This is probably be-
indicating that karst-related diagenesis was not lim- cause of the fact that most karst fills are composed of
ited to the L1 sequence. Instead, it appears that two the same facies as nonkarsted intervals. Anhydrite
types of karst-related processes occurred as outlined void fills are an obvious exception to this because
below. they contain no porosity or permeability; but they are
Primary karsting and dissolution probably occurred generally very small and probably of little impact on
during the post-L1 lowstand. At this time, the exposed reservoir flow.
382 / Ruppel and Jones

Second, with the exception of areas in which there but is sufficient to provide necessary information to
are cores, the distribution of karst features in the define major facies successions and cyclicity and to
reservoir section is not definable. Efforts to identify provide a strong basis for accurate correlation and
karst fill using logs and 3-D seismic data are thwarted interpretation of facies and cyclicity to nearby wells.
by the similar lithological and petrophysical proper- Facies recognizable on image logs include (1) tidal-flat
ties these features share with surrounding rocks. It is facies (Figure 8B), (2) peloid wackestone-packstone
tempting to conclude from these observations that (Figure 10B), (3) nodular wackestone-packstone
karst features have no impact of reservoir heteroge- (Figure 9B), (4) cross-bedded grainstone (Figure 13B),
neity or fluid flow. However, there are anomalies in (5) fusulinid wackestone-packstone (Figure 14B),
water production and flow rates in the Wichita that (6) karst breccia (Figure 26C), and (7) clay-rich mud-
cannot be readily explained by matrix petrophysical stone. Additionally, siltstone-sandstone facies can be
properties (T. Anthony, 2003, personal communica- identified from many wire-line logs.
tion). These phenomena may be the result of karst The identification of tidal-flat facies is especially
development. crucial for defining both reservoir architecture and
reservoir quality. Because they generally occupy cy-
RESERVOIR IMAGING cle tops, their definition makes it possible to define
cycle boundaries and thereby facilitates cycle-scale
Accurate definition of reservoir architecture and correlation. Equally important is the identification of
the distribution of rock fabrics in this architecture is tidal-flat facies for their petrophysical significance.
the key to defining improved methods for recovery As discussed previously, tidal-flat facies commonly
of hydrocarbons remaining in these systems. We display relatively high porosities but low permeabil-
used several methods to better image the reservoir at ity. Because of this, it is critical for accurate reservoir-
Fullerton field. Especially important in defining the quality mapping to distinguish tidal-flat rocks from
geologic architecture of the reservoir are (1) the cali- subtidal rocks that are typically higher in permeabil-
bration and use of borehole image logs to aid in the ity. Definition of the fusulinid wackestone facies is
identification and mapping of facies, cyclicity, and also of great importance because these rocks gener-
rock fabrics; and (2) the use of 3-D seismic data to ally represent the deepest water facies in the Leo-
constrain the geologic framework. nardian succession and typically are found at cycle
bases. They are thus important indicators of sea level
Identifying Facies and Cyclicity from rise and guides to cycle definition and correlation.
Borehole Image Logs Using calibrated image log responses, it is possible
Image logs are highly underused in the character- to create a very detailed record of facies in wells
ization of carbonate reservoirs. To most, the principal having good-quality image logs. Vertical resolution is
use of such logs is in the identification of fractures. potentially even better than that attainable from
However, image logs also have the potential to accu- cores (because of uncertainties of core and well depth
rately image many matrix properties that are key to ties). Cycles defined from image logs range in thick-
the proper characterization, modeling, and exploita- ness from 1 to 6 m (3 to 20 ft). Highest facies and cycle
tion of carbonate reservoirs. Traditionally, cores have resolution was obtained in intervals characterized by
been obtained to provide the key data needed for con- alternating subtidal and tidal-flat facies in the Lower
straining the distribution of reservoir facies, cyclicity, Clear Fork because of the marked contrast in image
and rock fabrics. However, because of cost, the num- log character of these two facies types.
ber of cores available is generally far smaller than
needed to accurately constrain these elements. If Imaging Stratigraphic Architecture and Reservoir
properly calibrated with core observations, borehole Development from 3-D Seismic Data
image logs can provide most of the required data to Like image logs, 3-D seismic data are remarkably
construct an accurate reservoir model at a fraction of underused in the characterization of carbonate res-
the cost. ervoirs. A critical need in developing approaches that
Examination, correlation, and comparison of im- lead to improving recovery from reservoirs contain-
age log character with cores show that seven facies ing significant volumes of remaining hydrocarbons
can reliably be identified from the image log in the is a better understanding of the 3-D distribution of
Abo–Wichita–Clear Fork succession at Fullerton. This reservoir attributes. The approaches outlined in this
is fewer than the 12 facies defined from core studies report for assembling and interpreting well data are
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 383

FIGURE 28. Fullerton field 3-D


seismic section (in time) show-
ing general continuity and iso-
pachous nature of Lower Clear
Fork and Wichita reservoir in-
tervals and contrasting clino-
formal nature of the Abo. Yellow
lines are time lines; dotted
line defines the top of Abo
Formation.

wire-line-log data are inap-


propriate for the Abo. In most
cases, it is not possible to re-
solve cycle-scale correlations
of either facies or time sur-
faces in such settings.

the most critical part of this effort. However, because Defining Reservoir Quality
they are limited to well control, they leave important Seismic data can also be robust indicators of po-
gaps in our understanding. Three-dimensional seis- rosity distribution in carbonate reservoirs. Zeng
mic data offer valuable data on interwell and extra- (2004) demonstrated the strong agreement between
well areas (meaning areas of the field where usable 3-D seismic impedance data and reservoir porosity at
well data are absent) and when properly interpreted Fullerton field. Because of this robust relationship,
and applied can greatly improve and, thus, constrain even simple amplitude extractions reveal field-scale
reservoir attribute models. Here, we provide some and fieldwide changes in porosity development that
brief insights into how 3-D data can be used to better are important for understanding controls on porosity
image both reservoir framework and porosity distri- development and for defining interwell and extra-
bution. A far more detailed and rigorous application well distribution of porosity.
of these 3-D data to the construction of the geological Figure 29 is an amplitude extraction map for HFS
model at Fullerton is described by Zeng (2004). 2.1 in a small area of the field created to examine
possible infill drilling locations. Note that whereas
Constraining Reservoir Architecture the northern half of the map displays high negative
Two- and three-dimensional seismic data at Ful- amplitudes indicative of high porosity, the southern
lerton field provide important guides to the stratal half of the area is characterized by markedly lower
architecture of the reservoir succession. Seismic am- amplitude. The porosity distribution revealed by this
plitude sections through the entire reservoir interval map indicates that only those wells in the northern
reveal that much of the section is characterized by half of the target area are likely to encounter good
generally parallel seismic reflectors (Figure 28). This porosity in this zone. It should be pointed out that
is not unexpected, considering the shallow-water- this information on porosity distribution is only ob-
platform depositional setting indicated by the cores tainable from the 3-D data volume; the quality of the
and the apparently subhorizontal correlations sug- well logs in the area is too poor to determine porosity.
gested by wire-line logs. However, 3-D and some 2-D Areas of poor well control like these (caused by
data suggest a very different architecture for the basal either the lack of wells or poor-quality logs) are com-
Leonardian, Abo Formation. Dip-oriented seismic lines mon through the field, and they greatly compromise
reveal sets of clinoformal reflections in the Abo that efforts to develop a fieldwide strategy to identify and
dip generally basinward (toward the east). This appar- target oil resources. This is readily apparent from a
ent clinoformal architecture of the Abo is consistent comparison of a porosity map based on well logs and
with observations of clinoformal fusulinid wacke- a seismic amplitude map for HFS L2.0 (Figure 30). The
stones and packstones in outcropping Abo-equivalent value of the seismic data volume is especially ap-
sections in the Sierra Diablo. These clinoforms dem- parent in the southern end of the field where the
onstrate that conventional horizontal correlations of limited well control does not accurately image the
384 / Ruppel and Jones

and correlated, depositional cycles represent the best


available indicator of original depositional surfaces
or time lines. The procedure for identifying and
using cycle boundaries for flow-unit definition has
been well described by Ruppel and Ariza (2002) and
Lucia and Jennings (2002) for the South Wasson
Clear Fork reservoir.
Ideally, the correlation of depositional cycle tops
should be based on a log that can be directly tied to
facies and one that is independent of diagenesis or
porosity development. The gamma-ray log in certain
ideal settings serves this function. However, through-
out most of the Leonardian carbonate succession (in
fact, throughout most of the Permian carbonate sec-
tion in the Permian basin), the gamma-ray log is not
accurate for detailed correlation because of variable
volumes of uranium, potassium, and thorium. The
spectral gamma-ray log can help to distinguish varia-
tions in these elements and, thus, is useful in sepa-
rating clastic-rich sections (that contain high levels
FIGURE 29. Map of negative amplitude data extracted of potassium and thorium). Because small volumes of
from Fullerton 3-D data showing porosity distribution clastics are commonly associated with tidal-flat facies,
in the L2.1 sequence. Data show that proposed wells in the spectral gamma ray is locally helpful in distin-
the northern half of the area will encounter reservoir guishing tidal-flat facies from subtidal facies (Ruppel,
porosity, whereas those in the southern half will not. 1992, 2002). However, variations in uranium content
are not always systematic, and because of this, gamma-
east – west-trending area of high porosity shown by ray response commonly varies independently of facies.
the amplitude map (compare Figure 30a with b). It is Resistivity logs can also locally be used to separate
apparent that basic amplitude data like these provide tidal-flat facies from subtidal facies on the basis of
a powerful supplement to well control in defining differences in saturation (Ruppel, 2002). However,
and predicting porosity distribution in the field. even under ideal circumstances, neither gamma-ray
logs nor resistivity logs can accurately depict facies in
RESERVOIR FRAMEWORK Leonardian rocks.
Accordingly, we used porosity logs to define and
A critical component of a robust reservoir model is correlate facies and cycle tops in the Lower Clear Fork
a geologically constrained reservoir framework. Geo- section at Fullerton field. The basis for this approach
logically accurate models must be based on correla- comes from integrated studies of cores and wire-line
tions of time-stratigraphic units, the most readily logs at Fullerton that demonstrate two important at-
correlative of which are cycle and sequence bound- tributes of Leonardian cycles. First, cycle-capping fa-
aries. At Fullerton field, we used correlations of cycle cies are either grain-rich subtidal or tidal-flat facies.
tops to construct the reservoir model for the Clear Second, porosity is most commonly associated with
Fork. Although these correlations are ultimately based these facies and, therefore, is most typically developed
on correlations of wire-line logs, the underlying basis at cycle tops. These relationships are key to the use of
for the interpretation and correlation of these logs is wire-line logs for correlation of cycles and of flow units.
a knowledge of 1-D and 2-D facies and cycle-stacking It is important to understand that there are limi-
relationships developed from integrated studies of tations to the accuracy of the porosity-log correlation
cores and outcrops. method. For example, because of diagenesis, all carbon-
ate facies display some variations in porosity. Cycle-
Cycle-based Flow-unit Definition top facies in the Leonardian, whether subtidal grain-
Reservoir flow units are most appropriately based stone or tidal flat, do exhibit lateral changes in porosity.
on the definition and mapping of depositional cycle So locally, cycle tops defined from porosity may be
boundaries. This is because when properly defined slightly mispositioned (e.g., because of a local decrease
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 385

FIGURE 30. Porosity development in HFS 2.0 (upper Wichita) across entire field based on wire-line logs (a) and 3-D
seismic data (b). Note that 3-D amplitude patterns suggest that well-based porosity mapping is inaccurate in some areas.

in porosity in the cycle-top facies or a local increase in reservoir modeling. This is important for two reasons.
porosity in the overlying cycle-base facies). From a First, coarser scale correlations are much more likely
strict chronostratigraphic point of view, this means to be in error. At Fullerton, we found that early cor-
that some cycle tops are incorrect, and that cycle cor- relations made at the sequence or cycle set scale were
relations locally cross time lines. From a reservoir mod- later proven to be off by a cycle or two after we recor-
eling point of view, however, this result is actually related the succession at the cycle scale. From a res-
probably preferable. That is because these Leonardian ervoir modeling point of view, this means that flow
rocks uncommonly contain cycle-base flow barriers, units defined by coarse-scale correlations are more
and flow probably does locally cross cycle boundaries. likely to cross-connect flow layers than those defined
In any case, the porosity-log correlation method, if by finer scale correlations. Second, upscaling (group-
properly constrained by core and outcrop calibration of ing of cycles into thicker flow-unit packages for
facies, porosity, and cyclicity, is the most geologically modeling) is more likely to retain the original geo-
sound basis for constructing cycle correlations and logical architecture if based on fine-scale (i.e., cycle-
establishing the basis for true flow-unit correlations. scale) correlations. Accordingly, we attempted to cor-
It is also important, however, to correlate geologi- relate the reservoir succession at the highest possible
cally defined cycle tops at the highest resolution pos- level of detail supported by outcrop and core obser-
sible. For example, outcrop studies demonstrate that vations and wire-line resolution.
cycles less than 3 m (10 ft) thick can be correlatable
across large areas of the platform (Ruppel et al., 2000). Lower Clear Fork Reservoir Architecture
Where wire-line data permit, an effort should be The robustness of the use of porosity logs for
made to correlate these thin cycles through the res- defining facies and cyclicity is apparent from core
ervoir as well, even if later upscaling is planned for and log relationships in HFS L2.2. A comparison of
386 / Ruppel and Jones

FIGURE 31. Comparison of core-


defined facies and cyclicity with po-
rosity logs in subtidal cycles of HFS
L2.2. Exxon FCU 6122. The systematic
relationship between cycle top facies
and higher porosity permits porosity
logs to be used for cycle and flow unit
correlation.

lateral changes in facies stacking in


L2.1 cycles suggest local variations in
sediment accumulation patterns
that may have been caused by topo-
graphic relief on the platform during
L2.1 flooding. The resultant com-
plex vertical and lateral facies distri-
bution patterns make accurate defi-
nition of cycles difficult. As a result,
cycles defined for L2.1 may actually
reflect combinations of cycles or cy-
cle sets.

Wichita Reservoir Architecture


Rigorous cycle definition is not
possible for the Wichita because
of the preponderance of very low-
accommodation, tidal-flat facies.
Both core and outcrop studies show
that rocks deposited in such settings
uncommonly display systematic
trends in vertical facies stacking and
core data and porosity logs in the Exxon FCU 6122 generally exhibit very low lateral facies continuity.
well (Figure 31) shows that porosity is nearly entirely Accordingly, patterns of depositional facies do not
associated with cycle-top, grain-rich subtidal facies. define extrinsic controls (e.g., sea level rise and fall)
Thus, porosity logs can be used to define both facies but, instead, local controls on sediment accumulation
and cyclicity. Other cored wells in the field exhibit (e.g., paleotopography and climate). Thus, for the
the same core and log relationship. Wichita, it is necessary to use diagenetic features to
Using this relationship, we defined and correlated define the reservoir framework.
15 cycles in the Lower Clear Fork. The average thick- Key diagenetic features used to define Wichita
ness of these cycles is about 5 m (17 ft) for L2.1 and cycles are mineralogy and porosity. In the northern
L2.2 throughout the entire interval. However, cycle part of the field, the upper Wichita contains multiple
thickness varies systematically by sequence in this in- intervals of low-porosity limestone that we have in-
terval. High-frequency sequence L2.2 cycles average terpreted to be cycle bases (Figure 18a). Although in-
about 3.3 m (11 ft) in thickness, whereas L2.1 cycles, terpreted largely from diagenetic relationships, cycles
with the exception of low-accommodation tidal-flat are closely tied to depositional surfaces and, thus, rep-
cycles at the base and top of the sequence, are nearly resent a good approximation of time surfaces. These
twice as thick. This is probably caused by two factors. surfaces are also especially useful from a reservoir
First, L2.1 deposits record the maximum flooding of point of view because they define layers of low and
the platform and probably the development of maxi- high porosity and permeability. The top four cycles
mum accommodation. The overall upward thinning in the Wichita are defined on the basis of these cyclic
of cycles from L2.1 to L2.2 is consistent with an dolostone (high-porosity)-limestone (low-porosity)
overall upward decrease in accommodation. Second, couplets.
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 387

However, no limestone is present in the upper Fork and upper Wichita (L2) surfaces; (2) the effec-
Wichita in the southern part of the field area nor in tive pinch-out of lower Wichita (proximal L1) sur-
the lower part of the Wichita in any part of the field faces at the facies boundary from Wichita tidal flats
(Figure 18a). In these areas, it was necessary to use to Abo subtidal; and (3) the clinoformal nature of
porosity trends alone to construct the reservoir frame- Abo (distal L1).
work. Although it is still likely that porosity variations
in these areas are closely tied to diagenesis associated
with depositional surfaces (i.e., cycle tops), we have Fieldwide Patterns of Porosity Distribution
little independent evidence (i.e., mineralogical varia- Core and log data from wells that penetrate the
tions) with which to demonstrate this. Accordingly, Abo indicate that the Abo locally contains high po-
the framework established for these parts of the Wi- rosity. However, well control is too sparse and in-
chita is far less geologically robust than that defined complete to accurately map porosity distribution.
for other parts of the reservoir. For this part of the Wang and Lucia (2004) presented a realization of the
reservoir, we defined and correlated 12 surfaces on 3-D distribution of Abo porosity using a full-field
the basis of porosity. We have confidence that these model based on available porosity data and a con-
surfaces are subparallel to time surfaces on the basis ceptual geological framework. However, the accura-
of their parallelism to overlying Lower Clear Fork cy of this porosity distribution must be considered
cycle-top surfaces and to a middle Wichita marine relatively low.
flooding surface. However, we cannot tie them rig- Mapping of phih (porosity  thickness) from wire-
orously to cyclicity. line logs demonstrates that the Wichita Formation
contains higher total phih (and higher average po-
Abo Reservoir Architecture rosity) than either the Abo or the Lower Clear Fork. In
The Abo is dominated by clinoformal bedding general, maps show that the highest phih lies along
typical of outer-ramp carbonate deposits; this is dem- the field structural crest, suggesting that porosity de-
onstrated both by outcrop studies (Kerans et al., 2000) velopment is related to structure. However, a more de-
and by 3-D seismic data at Fullerton (Figure 28). It is tailed examination of phih by stratigraphic horizon
therefore certain that the Abo architecture differs shows that phih is related more to platform position
significantly from the generally subparallel character than structure. The upper Wichita (HFS L2.0 TST)
of the depositional surfaces of the platform-top Wi- displays a well-defined, arcuate belt of higher phih
chita and Lower Clear Fork successions. Outcrop (Figure 30) that generally parallels but is slightly
studies and seismic studies at Fullerton and elsewhere landward of the inner-outer platform boundary (the
(Kerans et al., 2000; Ruppel et al., 2000) also dem- Wichita– Lower Clear Fork facies boundary). Decrease
onstrate that clinoformal outer platform successions in phih to the east is a function of the decreasing
like the Abo do not contain readily correlatable cyclic thickness of the Wichita near the facies transition
successions. This is caused by changes in sources and (Figure 19). The decrease in phih to the northwest,
distributional patterns and to extreme variations in however, is not associated with any change in depo-
lateral textures and fabrics of these transported sitional facies; Wichita rocks comprise similar peri-
sediments. Accordingly, it is not possible to establish tidal tidal-flat facies across the entire area. Accordingly,
an accurate internal architecture for these deposits. the cause of porosity development must be more a
For purposes of reservoir modeling, we created a series function of diagenesis than deposition. A model for
of conceptual clinoform surfaces to constrain the this diagenesis is presented in the following section.
reservoir architecture. Although these surfaces do Lower Clear Fork HFS L2.1 also displays an arcuate
not accurately describe the architecture of the Abo, trend in porosity development that is very similar to
they do illustrate the nonparallel and nonhorizontal that seen in the upper Wichita (Figure 33; see also
nature of these deposits and provide more realistic Figure 25b). In contrast with the Wichita, there is
constraints for modeling of reservoir attributes. evidence that this porosity trend is, at least in part,
associated with facies. Areas of highest porosity are
Reservoir Model associated with middle-ramp, grain-rich packstones.
The architecture of the reservoir framework de- Lower porosity areas are associated with outer-ramp
veloped for the Clear Fork reservoir at Fullerton is fusulinid-rich wackestones downdip and mud-rich
depicted by Figure 32. Key aspects of this model are packstones updip. It is probable that the high-phih
(1) the subhorizontal and subparallel nature of Clear belt represents a low-energy ramp crest. However, the
388 / Ruppel and Jones

FIGURE 32. Diagrammatic cross section showing framework used for reservoir model construction and general distribution
of major rock-fabric classes at Fullerton Clear Fork field. Class 1 rock fabrics display the highest permeability relative to
porosity, whereas class 3 fabrics display the lowest. It should be noted that the Lower Clear Fork displays far more complex
lateral variations in rock fabrics than depicted here. Datum is sea level.

change in phih appears also to be significantly re- HFS L2.0 and in the Lower Clear Fork HFS L2.1
lated to diagenesis. Areas of high porosity are also (Figures 30, 33). In each of these cases, belts of high
areas of high limestone abundance. An almost 1:1 porosity are situated immediately up depositional
relationship exists between the presence of limestone dip (landward) from the position of the underlying
and high-porosity areas (compare Figures 18b, 33). inner-outer platform margin (Wichita – Abo facies
No similar trend of high phih is apparent in HFS boundary). However, the mechanism for porosity
L2.2. However, like L2.1, areas of highest porosity are development must differ for the Wichita, which com-
associated with limestone. For L2.2, this is domi- prises peritidal facies that display no systematic varia-
nantly in the southern end of the field; however, a tions across the area, and the Lower Clear Fork, which
small area of limestone also exists in the northwest consists of subtidal rocks of three facies tracts.
corner of the field (Figure 18b). The absence of depositional facies variations in
the Wichita indicates that porosity formation is the
Models for Porosity Development result of diagenetic processes. Stable isotope data
Porosity is developed in all parts of the Abo – suggest that high-porosity dolomites in the Wichita
Wichita– Lower Clear Fork reservoir section. Howev- are the result of seawater-dominated diagenesis
er, there are obvious areal variations in porosity that (Figures 25a, 34a). These rocks display light d18O
reflect a combination of depositional, diagenetic, and values (average: 0.0% PDB) that are consistent with
structural controls on reservoir development. Most seawater dolomitization (based on an approximate
reservoir intervals exhibit spatial variations in poros- 3% fractionation from penecontemporaneous sea-
ity development that parallel platform paleotopog- water calcite; Land, 1980). The restriction of these
raphy. This is particularly apparent in the Wichita high-porosity rocks to the outer margin of the Wichita
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 389

FIGURE 33. Map of wire-line phih showing distribution of


porosity in the Lower Clear Fork HFS 2.1. Areas of high
porosity in HFS L2.1 correlate with areas of abundant
limestone. Compare with Figure 17.

tidal flat suggests that porosity development may be


the result of early seawater-dominated dolomitization
and stabilization along the seaward margin of the tidal
flat. Dolomitization may have been favored along this
seaward margin by the development of somewhat
shallower water conditions, perhaps like those seen in
tidal-flat island complexes (Figure 34a).
The Lower Clear Fork HFS 2.1 also displays a trend
in porosity development that closely parallels the
underlying Wichita– Abo and Wichita– Clear Fork
facies transitions (Figures 25b, 33). In this case, how-
ever, both mineralogy and porosity development are
related to facies. Core descriptions imply that the
trend of high porosity is coincident with a platform FIGURE 34. Models depicting three major stages of
diagenesis and porosity formation in the upper Wichita
ramp crest, an area characterized by more common
and the Lower Clear Fork (L2 sequence). (a) Early seawater
ooid-bearing, grain-rich packstones and grainstones dolomitization: Wichita, L1-L2. Wichita porosity develop-
(e.g., Figure 13C) and by locally more abundant tidal- ment is associated with early seawater dolomitization
flat caps. Ramp-crest development was probably con- along the outer margins of the tidal-flat complex. (b) Early
trolled by inherited paleotopography over the L1 seawater cementation and stabilization: Lower Clear Fork,
L 2.1. Porosity is related to preservation of primary po-
(Wichita –Abo) inner to outer ramp margin and/or
rosity in Lower Clear Fork ramp-crest limestones. (c) Reflux
differential subsidence over deep-seated faults in the dolomitization. Most of the Wichita and Lower Clear Fork
same way that the position of the Wichita– Abo facies was dolomitized by evaporatively concentrated, reflux
transition is most likely controlled by such deep brines. See text for discussion.
390 / Ruppel and Jones

structures. The ramp-crest trend is an area of abun- Rock Fabrics and Permeability Distribution
dant calcite; lower porosity areas to the east (inner In addition to a geologically defined stratigraphic
ramp) and west (outer map) are nearly entirely dolo- framework and an accurate distribution of porosity, a
mite (see Figure 18). Strontium and oxygen isotope robust reservoir model must contain a realistic dis-
data ( J. Kaufman, 1991, personal communication; tribution of permeability. To calculate and distribute
this study) from the calcites indicate that these rocks permeability at Fullerton, we used the rock-fabric
are essentially unaltered original marine precipitates. approach developed by Lucia (1995). This approach
Oxygen isotopes (d18O = 3.0%) are virtually iden- is based on defining and mapping key rock fabrics.
tical to best estimates of marine precipitates of middle Figure 32 displays the stratigraphic distribution of
Permian age (d18O = 2.8%; Given and Lohmann, rock fabrics and associated petrophysical classes
1985). Collectively, these data suggest the HFS L2.1 (groupings of rock fabrics having similar relationships
porosity trend is the result of early marine calcite between porosity and permeability) in the Fullerton
cementation in well-agitated and well-oxygenated reservoir. Abo outer platform skeletal packstones are
conditions along the ramp crest (Figure 34b). Lower dominated by rock fabrics with interparticle pores
porosity in updip dolomites is probably the result of representative of petrophysical class 1 (which exhib-
predolomitization compaction of these muddier sedi- its the highest permeability for a given porosity).
ments and possibly of porosity occlusion by dolomite Wichita tidal-flat facies, by contrast, comprise petro-
and anhydrite cementation. Compaction and dolo- physical class 2 rock fabrics (which have the lowest
mitization, and, thus, porosity loss, were probably permeability for a given porosity). Rocks of the Lower
limited in the ramp crest because of early calcite ce- Clear Fork are more complex than either the Abo or
mentation and stabilization and the grain-rich tex- Wichita in terms of type and distribution of rock
tures that are dominant there. fabrics. As suggested by Figure 32, subtidal Lower
Most Wichita and Lower Clear Fork rocks outside Clear Fork deposits include both petrophysical class 1
these two high-porosity corridors exhibit lower po- and 2 rocks (and locally contain mappable class 3
rosities and a very different stable isotopic character. tidal-flat facies). The distribution of class 1 and 2 rock
These rocks contain much heavier oxygen isotopes fabrics is dominantly a function of diagenesis (prin-
(d18O average 3.0% PDB) that are indicative of do- cipally mineralogy and pore type) and secondarily
lomitization by evaporatively concentrated refluxing related to depositional facies. For example, most ramp-
brines (Saller and Henderson, 1998; Ruppel, 2002). crest grain-rich limestone packstones are class 2 rocks;
Such brines may have been generated at multiple thus, their distribution can be mapped with wire-line
times during the Leonardian. Conditions necessary logs that define mineralogy (e.g., dual-porosity log
for reflux were probably developed at each of the suites, PE log). Other variations in rock fabrics and
documented sequence boundaries (Figure 34c). How- petrophysical class in the Lower Clear Fork are a func-
ever, it seems likely that the greatest potential for tion of dolomite crystal size, anhydrite cementation,
major reflux was at or soon after the L1-L2 (Wichita) development of moldic pores, and depositional tex-
or L2-L3 (Lower Clear Fork– Tubb) sea level lowstands ture. We examined 950 thin sections and obtained
(Figure 34c). The similarity in isotopic character 705 new measurements of core porosity and perme-
throughout both Wichita and Lower Clear Fork re- ability and 30 special core analysis measurements to
flux dolomites suggests that all were dolomitized by a define these properties across the field. A detailed dis-
single massive reflux event, presumably at about the cussion of the methodology and results of the char-
L2-L3 sea level fall and rise. acterization and mapping of these rock fabrics and
Although most reflux dolomites at Fullerton have calculation of permeability at Fullerton is available in
low porosities, grain-rich, Lower Clear Fork dolomite Jones and Lucia (2004).
facies typically exhibit good porosity and perme-
ability (e.g., Figure 13D). This demonstrates the key
function of depositional texture in porosity preser- SUMMARY AND CONCLUSIONS
vation in dolomites. The lack of porosity in updip
Wichita reflux dolomites, as well as in Lower Clear Both the procedures used to develop the reservoir
Fork mud-rich dolomite facies, probably reflects pre- framework at Fullerton field and many of the at-
dolomitization compaction and/or porosity reduc- tributes of this framework offer important guidelines
tion by dolomitization and sulfate emplacement dur- for the characterization of shallow platform carbon-
ing reflux. ate reservoirs and for the development of predictive
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 391

models of the distribution of reservoir properties. unreliable for this purpose because of the presence
Key geological findings from this study include the of highly variable volumes of uranium that show no
following. consistent relationship to facies. We found porosity
The Leonardian reservoir succession in the Perm- logs best suited for high-resolution, cycle-scale corre-
ian basin consists of three formations or facies suc- lations, but only if they have been properly calibrated
cessions, each of which is characterized by distinctive to facies-stacking patterns and facies.
facies, cyclicity, depositional architecture, porosity Wire-line logs do not always provide the best reali-
development, and petrophysical relationships. zation of porosity distribution. Where good-quality
The Abo consists largely of porous outer-ramp, well logs are lacking (e.g., between wells and in areas
open-marine, fusulinid and crinoid facies whose clino- with older logs) simple amplitude extractions from
formal architecture is clearly expressed on seismic 3-D seismic data provide superior resolution of both
data. Lateral facies continuity is poor, and as a result, reservoir architecture and the distribution of reser-
porosity distribution is complex and probably highly voir porosity.
discontinuous. The relationship between permeabil- All of the forgoing findings have significant im-
ity and porosity, however, is among the best in the pact on reservoir heterogeneity, architecture, and fluid
reservoir. flow. It is thus critical that all be incorporated into
The Wichita tidal-flat rocks comprise both high- reservoir models for accurate imaging of reservoir
stand systems tract updip equivalents of the Abo properties and meaningful simulation of fluid flow.
(Leonardian sequence L1) and TST updip equivalents
of the basal Lower Clear Fork (Leonardian sequence
L2). Cyclicity is poorly developed; reservoir architec- ACKNOWLEDGMENTS
ture is more controlled by diagenesis than by deposi-
tional facies. Although these deposits locally contain The results presented in this chapter are part of
high porosity, they exhibit relatively low permeabil- continuing research into the styles and causes of
ity and display poor, small-scale continuity. Areas of heterogeneity in shallow-water platform carbonate
high-porosity development are a result of early dolo- reservoirs in the Permian basin by the Bureau of
mitization and stabilization along the outer margins Economic Geology. Frequent, sometimes vigorous,
of the inner-platform, tidal-flat complex. Cycle-base discussions with colleagues at the Bureau, including
limestones display very low porosity and permeabil- Charlie Kerans, Robert Loucks, Jerry Lucia, and Jim
ity and act as local fluid-flow baffles between higher Jennings, have been especially helpful in formulat-
porosity cycle-top dolostones. ing the interpretations presented herein. Principal
The Lower Clear Fork consists of a dominantly funding for the study was provided by the U.S. De-
subtidal succession of three high-frequency sequences, partment of Energy, ExxonMobil Corporation, and
each of which records sea level rise (transgression) the University of Texas System. Additional support
and fall (regression). Reservoir development is large- was provided by member sponsors of the Bureau’s
ly restricted to subtidal facies (late transgression Carbonate Reservoir Characterization Research Lab-
and early highstand). Highest porosity and perme- oratory, including Anadarko, Aramco, BP, Chevron,
ability are associated with undolomitized grain-rich, ExxonMobil, Great Western Drilling, Kinder Mor-
subtidal packstones and grainstones of the ramp gan, Marathon, Occidental Petroleum, Petroleum De-
crest that have been relatively unaffected by major velopment Oman, Shell International, Statoil, and
diagenesis. TotalFinaElf. Special thanks are extended to David
Karst modification of the reservoir (including in- Smith, Terry Anthony, Steve Krohn, and Amy Po-
clined beds and monomict and polymict cave-fill well of ExxonMobil and Jeff Simmons, Craig Kemp,
breccias) is widespread at and around the L1-L2 se- and John Stout of Oxy for their contributions to
quence boundary. However, karst features are re- organizing the project and providing data. David
solvable only with cores or image logs, and neither Smith and Terry Anthony have been especially help-
wire-line log nor core analysis data reveal definitive ful in proving insights into the field geology and
differences in petrophysical properties. Neverthe- engineering issues. We also express our apprecia-
less, well-performance data suggest that karst devel- tion to Stephen Hartman and Tim Hunt of the
opment may affect fluid flow. University of Texas System West Texas Operations
Wire-line logs must be used for constructing the Office for providing both collaborative funding and
reservoir framework. However, gamma-ray logs are data.
392 / Ruppel and Jones

REFERENCES CITED Holtz, M. H., S. C. Ruppel, and C. R. Hocott, 1992, In-


tegrated geologic and engineering determination of
Atchley, S. C., M. G. Kozar, and L. A. Yose, 1999, A pre- oil-reserve-growth potential in carbonate reservoirs:
dictive model for reservoir characterization in the Journal of Petroleum Technology, v. 44, no. 11, p. 1250–
Permian (Leonardian) Clear Fork and Glorieta forma- 1258.
tions, Robertson field area, west Texas: AAPG Bulletin, Jones, R. H., and F. J. Lucia, 2004, Integration of rock fabric,
v. 83, p. 1031 – 1056. petrophysical class, and stratigraphy for petrophysical
Bebout, D. G., F. J. Lucia, C. R. Hocott, G. E. Fogg, and G. W. quantification of sequence-stratigraphic framework,
Vander Stoep, 1987, Characterization of the Grayburg Fullerton field, Texas, in S. C. Ruppel, ed., principal
reservoir, University Lands Dune field, Crane County, investigator, Multidisciplinary imaging of rock prop-
Texas: University of Texas at Austin, Bureau of Eco- erties in carbonate reservoirs for flow unit targeting:
nomic Geology, Report of Investigations 168, 98 p. University of Texas at Austin, Bureau of Economic Ge-
Bein, A., and L. S. Land, 1982, The San Andres carbonates ology, final contract report to Department of Energy,
in the Texas Panhandle: Sedimentation and diagen- Contract DE-FC26-01BC1535 1, p. 125 – 162.
esis associated with Mg-Cl brines: University of Texas Jones, R. H., and S. C. Ruppel, 2004, Evidence of post-
at Austin, Bureau of Economic Geology, Report of Wolfcampian fault movement and its impact on Clear
Investigations 121, 48 p. Fork reservoir quality: Fullerton field, west Texas, in
Dutton, S. P., E. M. Kim, R. F. Broadhead, C. L. Breton, W. D. R. C. Trentham, ed., Banking on the Permian basin:
Raatz, S. C. Ruppel, and C. Kerans, 2005, Play analysis Plays, field studies, and techniques: West Texas Geo-
and digital portfolio of major oil reservoirs in the logical Society Fall Symposium: West Texas Geological
Permian basin: University of Texas at Austin, Bureau of Society Publication 04-112, p. 207.
Economic Geology Report of Investigations No. 271, Kerans, C., and W. M. Fitchen, 1995, Sequence hierarchy
287 p., CD-ROM. and facies architecture of a carbonate ramp system: San
Fischer, A. G., and M. Sarntheirn, 1988, Airborne silts and Andres Formation of Algerita escarpment and western
dune-derived sands in the Permian of the Delaware Guadalupe Mountains, west Texas and New Mexico:
basin: Journal of Sedimentary Petrology, v. 58, p. 637 – University of Texas at Austin, Bureau of Economic
643. Geology, Report of Investigations 235, 86 p.
Fitchen, W. M., M. A. Starcher, R. T. Buffler, and G. L. Kerans, C., and K. Kempter, 2002, Hierarchical stratigraph-
Wilde, 1995, Sequence stratigraphic framework and ic analysis of a carbonate platform, Permian of the
facies models of the Early Permian platform margins, Guadalupe Mountains: University of Texas at Austin,
Sierra Diablo, west Texas, in R. A. Garber and R. F. Bureau of Economic Geology (AAPG Datapages Dis-
Lindsay, eds., Wolfcampian – Leonardian shelf margin covery Series 5), CD-ROM.
facies of the Sierra Diablo — Seismic scale models for Kerans, C., and S. C. Ruppel, 1994, San Andres sequence
subsurface exploration: West Texas Geological Society framework, Guadalupe Mountains: Implications for
Publication 95-97, p. 23 – 66. San Andres type section and subsurface reservoirs, in
Garber, R. A., and P. M. Harris, 1990, Depositional facies of R. A. Garber and D. R. Keller, eds., Field guide to the
the Grayburg/San Andres dolomite reservoirs; Central Paleozoic section of the San Andres Mountains: Perm-
Basin platform, Permian basin, in D. G. Bebout and ian Basin Section SEPM Publication 94-35, p. 105 – 115.
P. M. Harris, eds., Geologic and engineering ap- Kerans, C., F. J. Lucia, and R. K. Senger, 1994, Integrated
proaches in evaluation of San Andres/Grayburg hydro- characterization of carbonate ramp reservoirs using
carbon reservoirs — Permian basin: University of Texas Permian San Andres Formation outcrop analogs: AAPG
at Austin, Bureau of Economic Geology, p. 1 – 20. Bulletin, v. 78, no. 2, p. 181 – 216.
Given, R. K., and K. C. Lohmann, 1985, Derivation of the Kerans, C., K. Kempter, J. Rush, and W. L. Fisher, 2000,
original isotopic composition of Permian marine car- Facies and stratigraphic controls on a coastal paleo-
bonates: Journal of Sedimentary Petrology, v. 55, p. 430– karst: Lower Permian, Apache Canyon, west Texas, in
439. R. Lindsay, R. Trentham, R. F. Ward, and A. H. Smith,
Hardie, L. A., and P. Garrett, 1977, General environmental eds., Classic Permian geology of west Texas and
setting, in L. A. Hardie, ed., Sedimentation on the southeastern New Mexico, 75 years of Permian basin
modern carbonate tidal flats of the northwest Andros oil and gas exploration and development: West Texas
Island, Bahamas: Baltimore, Johns Hopkins University Geological Society Publication 00-108, p. 55 – 82.
Press, p. 12 – 50. Land, L. S., 1980, The isotopic and trace element geo-
Holtz, M. H., and C. M. Garrett, 1990, Geologic and chemistry of dolomite: The state of the art, in D. H.
engineering characterization of Leonardian carbonate Zenger, J. B. Dunham, and R. L. Ethington, eds., Con-
oil reservoirs: A framework for strategic recovery prac- cepts and models of dolomitization: SEPM Special Pub-
tices in four oil plays (abs.), in J. E. Flis and R. C. Price, lication 28, p. 87– 110.
eds., Permian basin oil and gas fields: Innovative ideas Leary, D. A., 1985, Diagenesis of the Permian (Guadalu-
in exploration and development: West Texas Geolog- pian) San Andres and Grayburg formations, Central
ical Society Publication 90-87, p. 76. Basin platform, Permian basin, west Texas: Master’s
Key Role of Outcrops and Cores in Carbonate Reservoir Characterization and Modeling / 393

thesis, University of Texas at Austin, Austin, Texas, Ruppel, S. C., 1992, Styles of deposition and diagenesis in
125 p. Leonardian carbonate reservoirs in west Texas: Impli-
Leary, D. A., and J. N. Vogt, 1990, Diagenesis of the San cations for improved reservoir characterization: Soci-
Andres Formation (Guadalupian) reservoirs, Univer- ety of Petroleum Engineers Annual Exhibition and
sity Lands, Central Basin platform, in D. G. Bebout and Technical Conference, SPE Paper 24691, p. 313 – 320.
P. M. Harris, eds., Geologic and engineering ap- Ruppel, S. C., 2002, Geological controls on reservoir
proaches in evaluation of San Andres/Grayburg hydro- development in a Leonardian (Lower Permian) car-
carbon reservoirs— Permian basin: University of Texas bonate platform reservoir, Monahans field, west Texas:
at Austin, Bureau of Economic Geology, p. 21 – 28. University of Texas at Austin, Bureau of Economic
Lohman, K. C., and J. C. G. Walker, 1989, The d18O record Geology, Report of Investigations 266, 58 p.
of Phanerozoic abiotic cements: Geophysical Research Ruppel, S. C., and E. E. Ariza, 2002, Cycle and sequence
Letters, v. 16, p. 319 – 322. stratigraphy of the clear fork reservoir at South Wasson
Longacre, S. A., 1990, The Grayburg reservoir, north Mc- field: Gaines County, Texas, in F. J. Lucia, ed., Inte-
Elroy unit, Crane County, Texas, in D. G. Bebout and grated outcrop and subsurface studies of the interwell
P. M. Harris, eds., Geologic and engineering ap- environment of carbonate reservoirs: Clear Fork (Leo-
proaches in evaluation of San Andres/Grayburg hy- nardian-age) reservoirs, west Texas and New Mexico:
drocarbon reservoirs — Permian basin: University of University of Texas at Austin, Bureau of Economic
Texas at Austin, Bureau of Economic Geology, p. 239 – Geology, final technical report to the Department of
273. Energy, Contract DE-AC26-98BC15105, p. 59 – 94.
Loucks, R. G., 1999, Paleocave carbonate reservoirs: Ruppel, S. C., and D. G. Bebout, 2001, Competing effects
Origins, burial-depth modifications, spatial complex- of depositional architecture and diagenesis on car-
ity, and reservoir implications: AAPG Bulletin, v. 83, bonate reservoir development: Grayburg Formation,
p. 1795 – 1834. South Cowden field, west Texas: University of Texas
Lucia, F. J., 1995, Rock-fabric/petrophysical classification at Austin, Bureau of Economic Geology, Report of
of carbonate pore space for reservoir characterization: Investigations 263, 62 p.
AAPG Bulletin, v. 79, no. 9, p. 1275 – 1300. Ruppel, S. C., and H. S. Cander, 1988a, Effects of facies and
Lucia, F. J., and J. W. Jennings Jr., 2002, Calculation and diagenesis on reservoir heterogeneity: Emma San Andres
distribution of petrophysical properties in the South field, west Texas: University of Texas at Austin, Bureau
Wasson Clear Fork field, in F. J. Lucia, ed., Integrated of Economic Geology, Report of Investigations 178,
outcrop and subsurface studies of the interwell envi- 67 p.
ronment of carbonate reservoirs: Clear Fork (Leonar- Ruppel, S. C., and H. S. Cander, 1988b, Dolomitization of
dian-age) reservoirs, west Texas and New Mexico: shallow-water platform carbonates by sea water and
University of Texas at Austin, Bureau of Economic seawater-derived brines: San Andres Formation (Gua-
Geology, final technical report to the Department of dalupian), west Texas, in Sedimentology and geo-
Energy, Contract DE-AC26-98BC15105, p. 95 – 142. chemistry of dolostones: SEPM Special Publication 43,
Major, R. P., G. W. Vander Stoep, and M. H. Holtz, 1990, p. 245 – 262.
Delineation of unrecovered mobile oil in a mature do- Ruppel, S. C., W. B. Ward, E. E. Ariza, and J. W. Jennings Jr.,
lomite reservoir: East Penwell San Andres unit, Univer- 2000, Cycle and sequence stratigraphy of Clear Fork
sity Lands, west Texas: University of Texas at Austin, reservoir-equivalent outcrops: Victorio Peak Forma-
Bureau of Economic Geology, Report of Investiga- tion, Sierra Diablo, Texas, in R. Lindsay, R. Trentham,
tions 194, 52 p. R. F. Ward, and A. H. Smith, eds., Classic Permian
Mazzullo, S. J., 1982, Stratigraphy and depositional mosaics geology of west Texas and southeastern New Mexico,
of lower Clear Fork and Wichita groups (Permian), 75 years of Permian basin oil and gas exploration and
northern Midland basin, Texas: AAPG Bulletin, v. 66, development: West Texas Geological Society Publica-
p. 210 – 227. tion 00-108, p. 109 – 130.
Mazzullo, S. J., and A. Reid, 1989, Lower Permian platform Saller, A. H., and N. Henderson, 1998, Distribution of po-
and basin depositional systems, northern Midland rosity and permeability in platform dolomites: Insights
basin, Texas, in P. D. Crevello, J. J. Wilson, J. F. Sarg, from the Permian of west Texas: AAPG Bulletin, v. 82,
and J. F. Read, eds., Controls on carbonate platform no. 8, p. 1528 – 1550.
and basin development: SEPM Special Publication 44, Tyler, N., and N. J. Banta, 1989, Oil and gas resources re-
p. 305 – 320. maining in the Permian basin: Targets for additional
Presley, M. W., 1987, Evolution of Permian evaporite basin hydrocarbon recovery: University of Texas at Austin,
in Texas Panhandle: AAPG Bulletin, v. 71, p. 167 – 190. Bureau of Economic Geology, Geological Circular 89-4,
Presley, M. W., and K. A. McGillis, 1982, Coastal evaporite 20 p.
and tidal-flat sediments of the upper Clear Fork and Vogt, J. N., 1986, Dolomitization and anhydrite diagenesis
Glorieta formations, Texas Panhandle: University of of the San Andres (Permian) Formation, Gaines
Texas at Austin, Bureau of Economic Geology, Report County, Texas: Master’s thesis, University of Texas at
of Investigations 115, 50 p. Austin, Austin, Texas, 202 p.
394 / Ruppel and Jones

Wang, F., and F. J. Lucia, 2004, Reservoir modeling and Zeng, H., 2004, Construction and analysis of 3-D seismic
simulation of Fullerton Clear Fork field, Andrews porosity inversion models, in S. C. Ruppel, ed., prin-
County, Texas, in S. C. Ruppel, ed., principal investi- cipal investigator, Multidisciplinary imaging of rock
gator, Multidisciplinary imaging of rock properties in properties in carbonate reservoirs for flow unit tar-
carbonate reservoirs for flow unit targeting: University geting: University of Texas at Austin, Bureau of Eco-
of Texas at Austin, Bureau of Economic Geology, final nomic Geology, final contract report to Department
contract report to Department of Energy, Contract DE- of Energy, Contract DE-FC26-01BC1535 1, p. 305 –
FC26-01BC1535 1, p. 219 – 304. 342.
Ye, Q., and S. J. Mazzullo, 1993, Dolomitization of lower Zeng, H., and C. Kerans, 2003, Seismic frequency control
Permian platform facies, Wichita Formation, north plat- on carbonate seismic stratigraphy: A case study of the
form, Midland basin, Texas: Carbonates and Evapo- Kingdom Abo sequence, west Texas: AAPG Bulletin,
rites v. 8, no. 1, p. 55 – 70. v. 87, no. 2, p. 273 – 293.
11
Weissenberger, J. A. W., R. A. Wierzbicki, and N. J. Harland, 2006, Carbonate
sequence stratigraphy and petroleum geology of the Jurassic deep
Panuke field, offshore Nova Scotia, Canada, in P. M. Harris and L. J.
Weber, eds., Giant hydrocarbon reservoirs of the world: From rocks
to reservoir characterization and modeling: AAPG Memoir 88/SEPM
Special Publication, p. 395 – 431.

Carbonate Sequence Stratigraphy


and Petroleum Geology of the
Jurassic Deep Panuke Field,
Offshore Nova Scotia, Canada
John A. W. Weissenberger
Husky Energy Inc., Calgary, Alberta, Canada

Richard A. Wierzbicki
EnCana Corporation, Calgary, Alberta, Canada

Nancy J. Harland
EnCana Corporation, Calgary, Alberta, Canada

ABSTRACT

P
anCanadian Petroleum (now EnCana Corporation) discovered the deep Pa-
nuke field in 1998 with the drilling of the PP-3C well. The well was drilled in
90 m (295 ft) of water, 250 km (155 mi) southeast of Halifax, Canada. Subsequent
delineation and development drilling has proven a significant gas accumulation.
The gas is trapped, by a combined structural-stratigraphic configuration, in
the Upper Jurassic reefal and oolitic limestones and dolomites of the Abenaki
Formation. The Jurassic carbonate platform on the Scotian Shelf was attached to a
metamorphic hinterland, so that the sediments contain varying amounts of silic-
iclastics. Abundant secondary porosity was encountered, ranging from leached
matrix and intercrystalline to vuggy and/or cavernous. Textural, petrographic,
and isotopic evidences suggest that deep burial and hydrothermal diagenetic pro-
cesses caused the porosity. The gas is believed to have been sourced from adjacent
Verrill Canyon Formation shales, whereas small amounts of hydrogen sulfide have
been isotopically linked to synrift evaporites underlying the Abenaki.
The Abenaki is divided into seven third-order depositional sequences, the
Abenaki V being the primary gas zone. These sequences have been regionally cor-
related using geology and a grid of two-dimensional seismic data. Three-dimensional
seismic data was used in delineation drilling and reservoir characterization. Deep
Panuke is the first, and remains at writing, the only significant hydrocarbon dis-
covery in the Mesozoic carbonates of the continental shelf of eastern North America.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215883M883275

395
396 / Weissenberger et al.

INTRODUCTION are Late Triassic to Early Jurassic (Sinemurian–Pliens-


bachian) in age. These comprise synrift siliciclastics
A recent round of exploration drilling for hydro- and evaporites of the Eurydice and Argo formations
carbons on the Scotian Shelf has provided abundant (McIver, 1972; Jansa and Wade, 1975). Overlying these
new data on the Jurassic carbonates and associated are microcrystalline, restricted-marine, occasionally an-
strata of the region. Beginning in 1998, EnCana Cor- hydritic dolomites of the Iroquois Formation (McIver,
poration, then PanCanadian Petroleum Ltd., initiated 1972). The Iroquois is overlain by a siliciclastic succes-
drilling for deep targets below existing production of sion, comprising red beds, immature sands, and asso-
the Panuke field, discovered in 1986. The prospective ciated strata. These are assigned to the Mohican For-
reservoir was the Jurassic carbonates of the Abenaki mation (Given, 1977).
Formation, below oil-bearing siliciclastic sands of the Normal-marine carbonate deposition began with
Cretaceous Panuke Formation. The drilling program the Abenaki Formation (McIver, 1972; revised by Eliuk,
led to the discovery and delineation of the giant deep 1978). It has been lithostratigraphically subdivided
Panuke gas field. This chapter proposes new stratigraph- into several members: the Scatarie, Misaine, Baccaro,
ic, depositional, and paleogeographic interpretations and Artimon (and the Roseway unit; Wade, 1977). For
of the Nova Scotia Jurassic based on the new seismic further discussion, see Eliuk (1978). However, for the
and well data. Misaine Member, which is a marine shale, these com-
prise the bulk of the Jurassic carbonate strata of the
Scotian Shelf. Regionally, these are time equivalent
Regional Geology to shallow-marine and deltaic sediments of the Mic
The structural and stratigraphic development of Mac Formation. The lateral relationships between
the Scotian Shelf Jurassic is strongly influenced by its these lithofacies belts will be discussed below. The
location on the western edge of the North Atlantic corresponding Jurassic deep-water shale succession
(Figure 1). The Nova Scotia Jurassic was at the north- is termed the Verrill Canyon Formation, thought to
ern end of the Bahama –Grand Bank gigaplatform as be time equivalent to sediments from the Mohican
described by Poag (1991). This refers to the carbonate Formation to Early Cretaceous strata. This lithostra-
depositional system that dominated the eastern mar- tigraphy (Figure 4) will be discussed from a sequence-
gin of North America. stratigraphic perspective below.
Extensional and transtensional tectonics, associ- The most comprehensive early work on the Abe-
ated with the opening of the Atlantic Ocean, were naki was presented by Eliuk (1978), with later up-
dominant during the progressive widening of the dates (Eliuk et al., 1986; Eliuk and Levesque, 1989).
Atlantic, from north to south, beginning in the Late Other interpretations of the older wells included Jansa
Triassic (Jansa and Pe-Piper, 1988). The tectonic set- and Wade (1975), Jansa and Wiedmann (1982), and
ting was essentially a Late Triassic to Early Jurassic Ellis (1984). We and our colleagues have presented
rift basin, followed by a passive margin from the Mid- preliminary aspects of this project (e.g., Weissenber-
dle Jurassic through today. ger et al., 2000; Harland et al., 2002; Hogg and Ena-
The basement rocks of the Scotian Shelf are of chescu, 2003), which is augmented and amended
continental origin, consisting of Devonian granites herein.
and late Precambrian to Ordovician metasediments
(Welsink et al., 1989), which were then intruded by
Devonian-aged granites. Figure 2 shows some of the Background and Methodology
main structural elements of the Scotian Shelf, includ- Prior to the drilling of the PP3C discovery well in
ing major, fault-bounded structural highs and inter- 1998, 53 wells had penetrated the Abenaki Formation
vening basinal areas. Three-dimensional (3-D) seismic on the Scotian Shelf. Only 17 of these wells were cored.
data from the deep Panuke area suggest that these The initial stratigraphic interpretation and explora-
larger structural elements are broken into smaller tion program was therefore based on these data; wire-
horsts and grabens. Jurassic strata are present only in line logs, cores, and drill cuttings from the aforemen-
the outer part of the Scotian Shelf in the form of an tioned wells and a regional grid of two-dimensional
onlapping wedge of sediments (Figure 3). seismic data. A 3-D survey, originally shot to image
The gross stratigraphy of this wedge reflects the shallower, oil-prone Cretaceous strata, covered part of
tectonic history, from initial rift to essentially passive- the Panuke area itself. This was used to pick the
margin sedimentation. The early, basin-fill sediments drilling location of the discovery well (PP-3C).
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 397

FIGURE 1. Paleogeography of the Late Jurassic, approximately 150 m.y. Modified from Scotese (1997).

Drilling of the appraisal wells, which followed isotopic analyses of gas samples, and geochemical
PP-3C, provided the additional data presented here- analyses of hydrocarbon and potential source inter-
in. Some additional conventional cores were cut vals, from wells that are no longer confidential, will
(see Appendix 1). The distribution of all these data is be described in detail below.
shown in Figure 5. Because of the prohibitive costs of
cutting conventional core and the additional risk of
encountering lost circulation while coring, a system LITHOFACIES AND
of rotary sidewall coring combined with formation DEPOSITIONAL ENVIRONMENTS
image logs (and other wire-line logs), was devised to
interpret lithofacies and stratigraphy. Sufficient core Deposition on the Jurassic continental margin of
data are now available to confidently construct facies Nova Scotia comprised a mixed carbonate-siliciclastic
models and correlate stratigraphy, as described in de- depositional system. The climate was tropical to sub-
tail below. tropical (Golonka et al., 1994; Scotese, 1997). Deposi-
Special analyses to determine diagenetic history, tional environments and their component lithofacies
reservoir, and source rock quality were also under- (belts) reflect this overall paleogeographic setting.
taken. These include detailed petrographic studies A generalized facies model for the Abenaki For-
of drill cuttings and conventional and sidewall core mation is shown in Figure 6. The depositional envi-
samples; fluid-inclusion and stable isotope analyses, ronments and component lithofacies are described
398 / Weissenberger et al.

FIGURE 2. Tectonic elements of the Scotian Shelf (modified from Welsink et al., 1989).

FIGURE 3. Regional seismic dip line across the Scotian Shelf. Note the wedge-shaped geometry of the Jurassic succession.
Colored lines indicate Abenaki third-order depositional sequences from Abenaki (AB) I at base (red), to AB II (purple),
AB III (brown), AB IV (orange), AB V (yellow), and AB VI/VII (black) at top. Depth is travel time in milliseconds.
FIGURE 4. Stratigraphy of the Scotian Shelf (left) (modified from CNSOPB, 1997), with detailed sequence and chronostratigraphy of the Abenaki Formation and
related strata (right). Chronostratigraphy is from Haq et al. (1987); Abenaki biostratigraphy from van Helden (see Appendix 2).
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 399
400 / Weissenberger et al.

1994; Wenzel and Strasser, 2001; Blomeier and Reij-


mer, 2002, respectively), which were most helpful,
were supplemented with our own field studies in Por-
tugal, the Swiss Jura, and Morocco.
The team was also able to effectively tie formation
image logs to specific lithofacies. This was done by
cutting numerous rotary sidewall cores in the pene-
trated formations. Two complimentary methods were
undertaken: cutting the cores before or after logging.
The former method would pick core points from
sample descriptions. This did not allow precise pick-
ing of sample points, but sample points were clearly
visible on the image logs, giving absolute certainty as
to which log facies was sampled. The latter method
allowed examination of the entire log suite and sub-
sequent sidewall coring of specific formation image
log (Formation MicroImager; FMI) facies. Overall,
more than 200 sidewall core samples were taken in
wells with FMI log suites.
Figure 7 shows the interpreted paleogeography of
the Scotian Shelf during the Late Jurassic. The dis-
FIGURE 5. Location map of the deep Panuke pool showing tribution of the facies belts is by definition general-
Abenaki penetrations and core control. ized, because, as discussed below, facies distribution
changed with changing stands of sea level. That said,
below, from the most landward to the most basin- a carbonate shelf margin to shelf interior coexisted
ward. More detailed facies analysis in the literature with siliciclastic inner shelf deposits. Siliciclastics
includes Eliuk (1978), Ellis (1984), Eliuk and Lev- were also passed into the basin at points along the
esque (1989), and Pratt and Jansa (1989). The most platform margin, notably the large Sable delta (Mic
recent summary, with examples from deep Panuke, Mac Formation) in the northeast part of the study
was presented by Wierzbicki et al. (2002). area. The shelf was generally carbonate dominated to
Because a correct facies model was imperative in the southwest and more siliciclastic rich to the north-
building an accurate sequence-stratigraphic reservoir east of the delta.
model, the exploration and development teams re- A brief description of the depositional environ-
viewed numerous potential analogs for deep Panuke. ments and the constituent lithofacies is given be-
Outcrop studies from similar basins (e.g., Leinfelder, low. Estimates for the relative bathymetries of the

FIGURE 6. Facies model for the Abenaki Formation with basic depositional environments (modified from Weissenberger
et al., 2000).
FIGURE 7. Paleogeography of the upper Abenaki Formation, modified from Wierzbicki and Harland (2004).
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 401
402 / Weissenberger et al.

FIGURE 8. Representative Abenaki lithofacies; (A) burrowed, fine- to medium-grained sandstone with a micritic matrix
(reworked lowstand shelf; I-100 8295.5 ft [2528.4 m]); (B) oncolite packstone (lagoon proximal to margin; K-62, 3384.2 m
[11,103.01 ft]); (C) oolitic grainstone (shoal margin; B-13, 2514 m [8248 ft]; coin is 18 mm [0.8 in.] in diameter).

lithofacies are derived from their position in shoaling- interior can also have a significant siliciclastic com-
upward successions; facies identification in wells that ponent. These are interpreted as deeper lagoon de-
are tied to seismic reflectors traceable from platform to posits, representing as much as 20 m (66 ft) bathym-
basin; and from the literature cited above. etry (the moat of Eliuk, 1978).

Siliciclastic Coastal Plain Platform Interior and Lagoon


Light-gray, very fine to medium quartz sand with A variety of lithofacies characterize this environ-
a micritic matrix is found in association with several ment. Peloidal wackestones to packstones have a mi-
other sediment types on the carbonate platform. nor coral, sponge, and stromatoporoid component
These associations are with ooids and oncoids; with and are commonly burrowed and stylolitic. These are
red and green shales and coal; and with oncolite-ooid- interpreted to represent shallow lagoonal deposits
megalodont packstone to wackestone and cryptalgal proximal to the margin. Oncolitic facies are also com-
mudstones. The sediments vary from pure siliciclastic mon in this setting (Figure 8B) and comprise wacke-
to mixed carbonate-siliciclastic deposits depending stones to packstones containing varying amounts
on local conditions. The respective subenvironments of reefal or oolitic material, depending on the proxi-
of deposition are lowstand sandstones preserved in mity to the margin. Argillaceous facies are observed in
lows or karst features on the platform; shallow la- the platform interior, dominantly dark-brown skeletal
goon; and nearshore biotopes, deltas, shoreface sands, wackestones, shales, and sands. Oncoids, gastropods,
and/or estuaries. These lithologies are all interpreted bivalves, ostracods, and ooids are the most common
to have been deposited in less than 10 m (33 ft) of clasts. These argillaceous lithologies are interpreted
water. Figure 8A shows a bioturbated sandstone in- as deeper shelf deposits (5 – 20 m [16 – 66 ft] water
terpreted as lowstand deposits on the shelf, reworked depth; the moat, Eliuk, 1978, p. 447), deposited during
during the ensuing transgression. transgressions or in parts of the shelf that were pe-
Fossiliferous (gastropod-pelecypod-ostracod) wacke- riodically not filled in a given fourth- or fifth-order
stones and dark-gray shales found in the platform depositional cycle.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 403

FIGURE 9. Representative Abenaki lithofacies (continued from Figure 8); (A) coral-stromatoporoid packstone to
boundstone (reefal platform margin; G-32, 12,710 ft [3874 m]); (B) skeletal packstone with grainy matrix (proximal
foreslope; G-32, 11,839 ft [3608 m]); (C) thrombolitic mudstone (open marine, basin; G-32, 14,414 ft [4393 m]).

Platform Margin margin or as backreef patch reefs, all deposited in from


This environment comprises two dominant com- 2 to 10 m (6.6 to 33 ft) of water. The darker wacke-
ponents, the skeletal or reefal (coral-stromatoporoid) stones are interpreted as shelter cavity deposits be-
and the oolitic. Despite limited core data and rela- cause of the dominantly grainy texture, suggesting
tively sparse well spacing, the platform crest may still overall high-energy conditions.
be interpreted to be dominated by oolitic grainstones Where the margin is dominantly oolitic, the ooids oc-
(Figure 8C), whereas coral-stromatoporoid bound- cur in massive to cross-bedded grainstones (Figure 8C).
stones and packstones (Figure 9A) largely occur in These are light brown in color and have minor oc-
somewhat deeper water. currences of oncoids, solitary corals, pelecypods, and
The reefal component comprises coral-stromatoporoid gastropods. The lithology can vary to include ooid
wackestones to packstones, with common chaetetids, grapestone and oncoid wackestone fabrics. These li-
sponges, bryozoa, pelecypods, gastropods, and cri- thologies are interpreted to represent high-energy
noids. The matrix is typically light brown and grainy. shoal deposits, deposited in less than 5 m of water,
These lithofacies are interpreted as high-energy de- alternating with reefal boundstone lithofacies at the
posits found anywhere from the immediate backreef platform margin or near-margin shoals where waves
through the reef crest to the immediate forereef. These are predominantly wind generated.
would represent water depths from 0 to 10 m (33 ft).
Similar lithologies can be found as much as 50 m (164 ft)
down the foreslope, representing allochthonous ma- Proximal Foreslope
terial shed downslope. Coral-stromatoporoid bound- This environment is characterized by coral-sponge-
stone, with chaetetid and algal components, forms demosponge wackestones to packstones with a light-
the other major reefal lithology. Light to medium gray or buff to dark-gray skeletal or micritic matrix
brown in color with a grainy skeletal matrix and typ- (Figure 9B). Other components include tubiphytes,
ically displaying massive to graded bedding, its minor lithistid sponges, stromatoporoids, chaetetids, cri-
components include bryozoa, pelecypods, gastropods, noids, gastropods, brachiopods, and pelecypods. This
and crinoids. These may be associated with dark-gray lithofacies may be associated with medium- to dark-
skeletal wackestones with the same fauna. These li- gray, laminated siltstones and the aforementioned,
thologies represent bound reef proper, either at the coarser reef-derived material.
404 / Weissenberger et al.

Grey to brown coral-sponge-demosponge pack- demosponges, crinoids oolites, and oncolites, also
stones to boundstones also occur in this setting. occur.
They are typically massive to poorly bedded, with an The boundstones are interpreted as hexatinellid
occasionally silty to sandy, mud-dominated matrix. sponge-microbialite reefs, the shales, wackestones, and
Fauna is deeper water scleractinian corals, sponges, packstones as interreef material deposited in greater
chaetetids, stromatoporoids, and tubiphytes. than 100 m (330 ft) water depth. The stromatoporoid-
The wackestones and packstones are interpreted coral deposits are shelf derived (turbidites) occurring
as coral-dominated deposits deposited in 10 –70 m in depths from 10 to more than 100 m (33 to more
(3.3–230 ft) of water. The packstones to boundstones than 330 ft).
are coral-dominated reefs. This facies is the main reefal
facies seen in core and FMI images at deep Panuke. Al-
though the Abenaki margin is relatively steep, abun- Discussion
dant reefal development occurred there, and the plat- There has been disagreement over whether the
form crest is dominated by grainstones. Abenaki Formation represents a dominantly reef- or
ramp-style depositional system (Eliuk, 1978; versus
Ellis, 1984; Ellis et al., 1985, 1990, respectively). Our
Distal Foreslope new drilling, including new core data and FMI logs
This environment is dominated by sponge coral obtained in the deep Panuke program, supports a
and tubiphytes wackestone, with minor amounts modified reefal model. Wells such as M-79, which
of stromatoporoid, chaetetid, and ooids. Microbia- essentially penetrate the entire Abenaki section, show
lite textures and a cyanobacterial micritic matrix are thick aggradational packages of coral-stromatoporoid
common. Dark-gray to black laminated shales and rudstone to boundstone as stacked third-order se-
medium- to dark-gray siltstones are associated li- quences evidenced on FMI logs and in sidewall cores.
thologies. In addition, lithistid sponge- and coral- These boundstones and rudstones contain signifi-
stromatoporoid packstones to boundstones occur in cant amounts of micrite in the matrix (which has
this setting. These are medium gray in color, with been recrystallized; J. Dravis, 2001, personal com-
a micritic to thrombolitic matrix. Tubiphytes, Micro- munication), suggesting that they were not subjected
selena (and other deep-water scleractinian corals), bryo- to continuous wave action; i.e., they were likely de-
zoans, pelecypods, brachiopods, and crinoids are asso- posited below fair-weather-wave base.
ciated components. Light- to medium-gray skeletal By contrast, the Abenaki (AB) I (Scatarie Member)
wackestone and gray to black shales are associated has a lithofacies composition and depositional pro-
lithologies. file observed on seismic, suggesting more of a ramp,
The wackestone and shaly lithofacies are inter- or open platform setting, with no well-developed reef
preted as background sediments, in-situ sponge, and margin. Eliuk (1978) interpreted the AB I to have
deep-water coral meadows, deposited in 30 –70 m more of a rimmed shelf depositional style south of the
(100 – 229 ft) of water. The boundstones are sponge- G-32 well.
dominated mounds, typically forming in oxygenated Seismic profiles (as discussed below) support the
waters, 30 –100 m (100 – 330 ft) deep. geological evidence. Dramatic platform-to-basin re-
lief is developed from the AB II upward, with steep
depositional dips, to a maximum of approximately
Basin 508, which clearly indicates a steep reef profile. The
Dark-gray to black shales and mudstones with lam- uppermost sequence (AB VII) has a gentler deposi-
inated to thrombolitic textures (Figure 9C), laminated tional profile, suggesting it is less reefal. Limited side-
siltstones, and marls constitute the dominant in- wall core and FMI images, along with drill-cutting
situ sediment. These are interpreted to represent examination, show a deep-water sponge, mound-
deposition in 100 – 200 m (330 – 660 ft) of water. dominated argillaceous environment. Further, our
Sponge material, pelagic foraminifera, radiolaria, observations of facies relationships elsewhere, partic-
and ammonites are the most common fossils. Hexa- ularly of the Kimmeridgian Ota reef in Portugal, sug-
tinellid sponge-tubiphytes boundstones to pack- gest that the ooid shoal and reef environments can
stones, with associated calcareous shales, are also be laterally and vertically adjacent, as is seen in the
observed in this setting. Stromatoporoid and coral Abenaki Formation. This, as well as the presence of
wackestones to packstones, with associated sponges, coarse winnowed matrix in some coral-stromatoporoid
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 405

FIGURE 10. Comparison of


formation image log (FMI) of
a large phaceloid coral (left)
from M-79 (3865 – 3866 m;
12,680 – 12,683 ft) and core of
the same from L-97 (3417 m
[11,210 ft]; coin is 18 mm
[0.8 in.] in diameter). The coral
is in growth position above
(and overlain by) interpreted
stromatoporoid-coral pack-
stone to boundstone.

shows an FMI image from


M-79 with a similar coral,
roughly 1 m (3.3 ft) in height,
also in growth position, in a
succession of nodular fore-
slope mud to wackestones.
Figure 11 compares an FMI
image of a vuggy reefal dolo-
mite from M-79 to core from
the vuggy upper Abenaki res-
ervoir in the H-08 well. The
boundstone intervals, suggests that the reefs did pe- textural similarities of leached and dolomitized
riodically grow above fair-weather-wave base, form- decimeter-size vuggy rock to the unaltered reefal
ing high-energy patch reefs in bathymetries similar to boundstone are clear.
the ooid shoals. They may have formed an armored
reef front at the platform edge but no well has yet
tested this zone. SEQUENCE STRATIGRAPHY
Near the Abenaki platform margin, southwest of
the Sable delta, siliciclastics are found only as rela- The predrill sequence-stratigraphic framework for
tively thin, isolated sands. However, core from the deep Panuke primarily reflects drill-cutting descrip-
lower Abenaki (sequence I; Scatarie Member) from tions and geophysical log interpretations because
the M-79 well has well-rounded gravel- to cobble- of the paucity of conventional core in the central
size metasediments in a lime mud matrix, with abun- Scotian Shelf. Gamma-ray logs commonly showed a
dant marine fauna such as crinoids and bivalves. higher response in muddy lithologies, whereas con-
This suggests that a much more proximal siliciclas- versely, clean gamma readings indicated higher en-
tic source existed during lower Abenaki deposition, ergy, winnowed sediments. Higher gamma response
as no such coarse siliciclastics are known from any was therefore indicative of either muddy foreslope
other cores, nor cuttings examined in this study. or platform interior deposits. Drill cuttings or side-
Eliuk (1978, p. 435) interpreted the Scatarie to be wall cores were typically used to distinguish between
absent (because of nondeposition) from several ba- these environments.
sin ridges penetrated by at least two wells. These Three basic platformal environments were distin-
may have been the source for the coarse siliciclastics guished using older cuttings descriptions, i.e., basic
cored in M-79. descriptions of lithology and fossil content: dark lime-
As discussed above, FMI images were successfully stone representing the deep-water and foreslope;
matched to lithofacies, particularly through the use of light-colored limestones representing shallow-water
sidewall coring. Interpretation of some images was carbonates (typically reef and shoal); and calcareous
relatively straightforward. Figure 10 shows a cored in- sandstones and sandy and silty carbonates repre-
terval with a large phaceloid coral in growth position senting reworked siliciclastic coastal-plain environ-
(foreslope environment) from the L-97 well. It also ments. Alternation of these environments allowed
406 / Weissenberger et al.

FIGURE 11. Comparison of formation image log (FMI) of a vuggy reefal boundstone from the M-79 well (left) to core of
the same lithofacies from PI-1A (4029.4 m; 13,219.8 ft).

the identification of the basic transgressive and re- in the deep Panuke area, carbonate deposition be-
gressive trends in any given well. came more aerially restricted and replaced by argilla-
On the dominantly carbonate part of the shelf, ceous foreslope and basinal deposits (e.g., at the G-32
southwest of the Sable delta, siliciclastics were found well), suggesting that the Abenaki shelf was initially
to be concentrated in fairly thin, discrete beds. Con- terminated by transgression before it was inundated
versely, in the region approaching Sable and farther by Lower Cretaceous siliciclastics. This stratigraphy
northeast on the Banquereau Bank, the dominantly is herein interpreted as a second-order depositional
siliciclastic section was broken by only a few carbon- sequence.
ate intervals. Based on the relative landward position Jansa (1993) discussed the possible causes for the
of the siliciclastic and the seaward position of the termination of the Abenaki platform, favoring con-
carbonate environments in the depositional model, tamination by siliciclastics. This cannot be discounted,
the dominance of siliciclastics or carbonates was in- but, as discussed below, the carbonate platform co-
terpreted to represent relative falling and rising sea existed with siliciclastics throughout the Jurassic. This
level, respectively. suggests that the ultimate termination of upper Abe-
naki deposition in the deep Panuke area likely had
other causes, such as relative sea level change.
Stratigraphic Architecture of the Higher order sequences, herein interpreted as third-
Second-order Sequence order depositional sequences, were initially defined
A schematic illustration of the depositional se- by the presence of thin siliciclastic sands and silts at
quence on the Scotian Shelf near deep Panuke is the sequence boundaries and corresponding deepen-
shown in Figure 12. The gross stratigraphic pattern ings in the middle of the sequences, represented by
shows the Abenaki platform initiated by a relative deeper water carbonates. The latter were interpreted
transgression above the widespread Mohican Forma- as the transgressive maximums of the sequences. These
tion siliciclastic shelf. Following relative aggradation surfaces were initially correlated in the area illustrated
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 407

FIGURE 12. Schematic block diagram of the Jurassic carbonate platform, deep Panuke area. Maximum flooding surface
of the AB V indicated by an asterisk. Time-equivalent basinal (Verrill Canyon Formation) sediments are not shown. For
horizontal scale, refer to Figure 7.

in Figure 12 and later extended along the entire Sco- AB I, which generally corresponds to the Scatarie
tian Shelf. In the region north of the Cohasset L-97 Member, is a broad carbonate ramp. It typically has
well, the upper Abenaki is dominated by siliciclastics. three recognizable shoaling-upward cycles, compris-
There, the sequences were correlated primarily using ing outer-ramp fossil wackestones to packstones and
the flooding surfaces, represented by extensive corre- inner-ramp oolitic and oncolitic facies; identifiable
latable limestone beds in the siliciclastic succession. by cleaning-upward profiles on the gamma-ray log.
New well data largely supported the sequences de- These shoaling patterns are also apparent in the
fined before PP-3C was drilled. Flooding surfaces (and cuttings descriptions, as in the M-79 well. In the deep
sequence boundaries) were better constrained by the Panuke area, the AB I is roughly 140 m (459 ft) thick.
better log data in the new wells tied to sidewall cores, The AB I is interpreted as the early transgressive
as well as detailed drill-cuttings descriptions. Inter- part of the second-order sequence. The base of the
pretation of these new data led to minor adjustments AB I should likely be placed in the upper part of the
of the sequence picks in the regional data set and the Mohican siliciclastics, but lack of deep well penetra-
definition of an additional third-order sequence (at tions precludes the identification of this surface. Sim-
the top of the Abenaki) in the deep Panuke area, the ilarly, part of the Mohican siliciclastic shelf would
AB VII. correspond to the lowstand of the second-order se-
quence, but (again) the lack of deep penetrations and
reduced seismic resolution at depth hampers the in-
Third-order Sequences terpretation of these strata.
The nature of third-order sequences in the defined The AB II comprises two readily identifiable com-
Abenaki second-order depositional sequence was in- ponents, the regionally extensive Misaine Forma-
fluenced by their position in the larger sequence. The tion shale and the progradational reefal platform
408 / Weissenberger et al.

above. The base of the sequence is taken at the top of quence is observed in three of the four penetrations of
the AB I carbonate ramp. If a lowstand systems tract is the sequence in the pool itself. This is interpreted to
present, it may consist of shelf-interior siliciclastics by- reflect the same deepening observed in the siliciclas-
passed into the basin. These have only been pro- tic realm to the northeast.
visionally identified on seismic data and not pene- The AB V has a moderate gamma-ray signature,
trated by wells. similar to the AB III and IV. It is as much as 150 m
The aforementioned, recognizable components (492 ft) thick in the deep Panuke area. The lithofacies
of the AB II are interpreted as the transgressive at the base of the sequence are muddier and have a
and highstand parts of the sequence, respectively. higher gamma response, with a forereef macrofauna.
The Misaine represents a major transgression, with This is interpreted as the transgressive part of the
basinal shale mapped seismically at least 15 km sequence. A lowstand systems tract for the sequence
(9 mi) landward of the underlying AB I shelf mar- is interpreted only from seismic data. Basinal mud-
gin (G. Syhlonyk, 2005, personal communication). stones and shales attributed to the AB V were pene-
It is on the order of 100 – 150 m (330 – 492 ft) thick trated in the M-88 well, but these are interpreted as
near the shelf margin. The furthest landward extent transgressive and highstand deposits. The upper AB V
of the Misaine is also interpreted as the maximum consists of coral stromatoporoid packstones to bound-
flooding surface (MFS) of the Abenaki second-order stones, representing a reef margin environment.
sequence. This transgression is recognized along Figure 12 again shows a correlatable carbonate
the North Atlantic continental margins (Jansa and bed in the lower and middle part of the sequence
Wiedmann, 1982) to the southeast margin of Tethys extending northeast of L-97. This is interpreted as
(Gradstein et al., 1999). representing the MFS of the AB V, which temporar-
The AB II highstand comprises approximately 220 m ily interrupted siliciclastic deposition.
(721 ft) of relatively clean carbonates that prograde The AB VI consists of 150 – 200 m (492 – 660 ft) of
over the Misaine. Seismic data shows downlap of intra- calcareous mudstones and shales with a minor silic-
highstand reflectors onto the Misaine (discussed and iclastic component. They display a moderate to high
illustrated below in Figure 13). Cores and drill cut- gamma-ray response in the deep Panuke area. These
tings show the AB II to consist of skeletal wackestones are interpreted to represent foreslope and basinal de-
immediately above the Misaine, with a gamma-ray posits derived from a backstepped carbonate plat-
reading transitional with the underlying shales. The form. Skeletal-peloidal packstones encountered in
remainder consists of oolitic packstones to grainstones the upper AB VI in the F-09 well are interpreted to be
or reefal boundstones. The AB II platform margin is part of this retreated carbonate platform. Approxi-
typically set back from that of the AB I. mately 150 m (492 ft) of basinal calcareous shale are
The AB III is the thickest of the interpreted third- interpreted as equivalent to the AB VI in the M-88 well.
order sequences. It ranges from 400 to 450 m (1300 Wells penetrating the AB VI on the platform northeast
to 1470 ft) thick in the deep Panuke area. Formation of L-97 encountered an entirely siliciclastic section.
MicroImager, sidewall core, and cuttings data indicate The AB VII is a thin (roughly 50-m [164-ft]-thick)
that it consists of coral-stromatoporoid packstones interval of shales and mudstones with a high gamma-
to boundstones, i.e., reefal strata. This composition, ray response. Where carbonate-rich intervals have
reflected also in a moderate gamma-ray signature, is been encountered in this interval (e.g., G-32 and D-42)
fairly consistent through the sequence, suggesting they consist of sponge-rich, argillaceous limestones.
an aggradational stacking pattern. These are interpreted to represent deeper foreslope to
The northeast part of the schematic (Figure 12) basinal environments (the Artimon Member of Eliuk,
shows the first appearance of abundant siliciclastics 1978). A basinally restricted, 20-m (66-ft) interval of
in the lower part of the AB IV, between the L-30 and sandstone encountered in M-88 is interpreted as the
L-57 wells. This is interpreted as the lowstand and lowstand systems tract of the AB VII.
early transgressive part of the sequence. The MFS is Regional correlations undertaken in this study sug-
interpreted to be represented by a distinctive clean gest the AB VII at deep Panuke correlates with skeletal
carbonate that can be correlated from L-97 to L-30. and oolitic grainstone shoals drilled in the upper-
Southwest of L-97, the AB IV is a platformal carbonate most Abenaki in the southwest part of the Scotian
as much as 150 m (492 ft) thick with a slightly greater Shelf, e.g., in the Acadia K-62 and Bonnet P-23 wells.
gamma response overall than the AB III. A higher gam- These have been termed the Roseway unit (Wade,
ma response in the middle to upper part of the se- 1977) and are of Lower Cretaceous Neocomian age.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 409

FIGURE 13. Two dip-oriented seismic lines extracted from 3-D data in the deep Panuke area. (A) Line 420, approximately
half-way between H-08 and the older well G-32. It shows a strong peak (green) picked as the top of the Abenaki I; the
trough above (blue) represents the Misaine Formation shale. It shows an offlapping Abenaki II succession, the top-lap
geometry suggesting a sequence boundary at the top of the unit. (B) On trace 3239, the Abenaki III displays an offlapping
geometry. A fairly strong trough-peak defines the top of the Abenaki II. Top-lap geometries again support the sequence
boundary at the top of the unit.
410 / Weissenberger et al.

FIGURE 14. Idealized fourth-order depositional sequence, Abenaki Formation. Constituent systems tracts and lithofacies are indicated (modified from Wierzbicki
The AB VII also has a shalier, deep-water
character northeast of L-97, but appears
less calcareous than in the south, suggest-
ing persistent proximity to siliciclastic
input.
An idealized fourth-order depositional
sequence of the Abenaki Formation,
shown in Figure 14, illustrates the approx-
imate thickness and facies distribution.
It assumes some subaerial exposure and
epikarstic dissolution at the basal se-
quence boundary. In this model, sili-
ciclastics were deposited on the shelf and
infilled dissolution pipes during relative
lowstand of sea level. Where favorable
conditions existed, a lowstand reef, limited
in lateral and vertical extent, would grow
on the preexisting slope. During the
transgressive phase, normal-marine car-
bonate deposition resumed on the carbon-
ate platform. The shelf-interior silic-
iclastics would be reworked during the
transgression and commonly preserved
only in topographic lows. A varied suite
of depositional environments as discussed
above would develop, including rela-
tively thick beds of basinal mudstones
and shale. The highstand would develop a
similar suite of facies, shifted somewhat
basinward of the underlying transgressive
deposits.
Figure 15 is a stratigraphic cross sec-
tion correlating the fourth-order parase-
quences and lithofacies of the AB V in
the deep Panuke area. The fourth-order
boundaries (AB Va–Vg) were defined dur-
ing the delineation and predevelopment
phase of the project (see below) based on
log signature and detailed drill-cuttings
descriptions (strip logs in Figure 15).
Immediately evident is the dominance
of reefal lithofacies in most of the se-
quence, except for the two wells drilled
most landward, B-90 and F-09, which are
dominated by oolitic shoal and lagoonal
environments. The reefal intervals dis-
and Harland, 2004).

play well-developed cyclicity from fore-


slope to reef, occasionally shoaling into
oolitic facies (e.g., top AB V in M-79A).
The downward dip of the fourth-order
boundaries from the landward toward the
margin wells (e.g., PI-1B, M-79) suggests
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 411

ongoing subsidence at the margin during AB V first interpreted from well data and correlated to
deposition. The margin wells also show significantly seismic data.
more foreslope lithologies in the lower AB VI, suggest- The two lines also show other noteworthy strati-
ing incipient drowning of the platform margin. graphic relationships. The overall aggradational na-
ture of the carbonate platform is evident on line 420,
although the break in slope of each third-order
Seismic Stratigraphy margin shifts from the AB II to IV. The AB V margin
Predrill geological interpretation of well data estab- is clearly stepped back several hundreds of meters
lished the basic sequence-stratigraphic framework, from that of the AB IV. The lower Abenaki is more
tied to seismic data prior to drilling the deep Panuke difficult to interpret in trace 3239 because con-
discovery. This was followed, after completion of the tinued basin subsidence causes significant basin-
PP-3C well, by a regional seismic-stratigraphic study. ward dips in the reflectors. The AB I margin likely
The objective of this work was to determine whether is somewhat more than 2 km (1.2 mi) basinward of
the interpreted geological sequences could be corre- the younger margins. A strong, steeply dipping peak
lated on seismic data and, further, whether seismic- develops in the foreslope, regionally traceable to the
stratigraphic criteria confirmed their interpretation as top of the AB III. We interpret this as a submarine
(third-order) depositional sequences. hardground representing a depositional hiatus in the
Figure 13 illustrates two arbitrary seismic lines basin during the post-AB III lowstand. Basinal re-
from 3-D seismic data in the deep Panuke area. They flectors onlap this surface. The upper Abenaki plat-
show several aspects of seismic geometry that substan- form margin is slightly stepped back from the AB II
tiate the geologically interpreted sequence boundaries. margin.
Figure 13A is a line (420) from the Panuke 3-D Similar seismic geometries were observed at the
survey. Reflectors above the Misaine are clearly in- other sequence boundaries first interpreted using
clined, basinward dipping, suggesting progradation geological (well) data. This provided some confidence
of AB II carbonate over and downlap onto the un- that regionally correlatable chronostratigraphic units
derlying shale. This is consistent with expected seis- could be identified and form a reliable stratigraphic
mic geometries above an MFS; the Misaine is the framework for exploration and development.
interpreted MFS of the second-order depositional se-
quence. Reflectors higher in the sequence display
flat-lying top sets on the interpreted platform, a dis- Discussion
cernable change in slope and dipping foresets at the Regional seismic data (Figure 3) suggests that the
basin margin. Finally, the most basinward-dipping Abenaki is a depositional wedge on the Scotian Shelf.
reflectors of the AB II do not appear to have any This line suggests increasing coastal onlap in the up-
time-equivalent top sets on the shelf. This offlapping per Abenaki (AB V –VII). Relative sea level data that
seismic geometry suggests an accommodation min- incorporate other regions (Figure 4) suggest, overall, a
imum at the interpreted top AB II surface. This sur- relatively constant onlap. Decreasing onlap in the Mic
face was tied to the geological sequence boundary Mac and lower Missisauga formations, along with the
interpreted from regional well control and regionally change from carbonate to dominantly siliciclastic
correlated on the seismic grid. deposition, is the main criteria for picking a second-
Figure 13B is a dip line (trace 3239) extracted order sequence boundary above the Abenaki.
from the Musquodoboit 3-D survey. It shows similar It is important to note that Figure 12 depicts the
seismic geometries in the AB III to those seen in the dominantly aggradational nature of the shelf in the
AB II on line 420. Reflectors in the lower AB III can deep Panuke area. The time-equivalent shales and
be traced from the platform interior, across the break basinal mudstones of the Verrill Canyon Formation
in slope at the margin, into the basinward-dipping are not shown. When the pool was discovered, po-
foreslope. Higher reflectors in the upper AB III are tential reservoir bodies were thought to include the
inclined near the margin and can be traced upward high-energy platform margin, pinnacle, or possibly
(landward) where they appear to terminate against atoll reefs, and siliciclastic sands bypassed into the
flat reflectors of the overlying sequence. This off- basin during relative sea level lowstand.
lapping geometry suggests a lack of accommoda- Figure 16 shows two isochron maps of the platform
tion on the platform at the interpreted sequence margin south of deep Panuke. They display several
boundary capping the AB III. This surface was also noteworthy features of the platform margin in that
412 / Weissenberger et al.

FIGURE 15. Stratigraphic cross section of the upper Abenaki Formation, deep Panuke area (wells are indicated in Figure 5).
Third-order depositional sequences and constituent fourth-order sequences (parasequence sets) are indicated.
Modified from Wierzbicki and Harland (2004).

part of the shelf. The isochrons at the shelf margin of and G-32 are prominent in both time slices, as is the
the AB I are more widely spaced than at the platform westward turn of the margin at the southern end of the
margin of the AB III, suggesting a gentler, ramp-style survey. This partly reflects the aggradational nature of
depositional profile. The configuration of the margin the platform in this region, but also the inheritance of
in the two different sequences is quite similar. For paleogeography with time, possibly because of the per-
example, the embayments just southwest of the H-08 sistent influence of underlying structural elements.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 413

We interpret the position of embayments and prom- adherence to the scheme of Vail et al. (1977), where
ontories on the platform margin to largely be con- second-order sequences have a duration of 10 –
trolled by the configuration of the underlying (syn- 100 m.y. and third-order sequences are 1–10 m.y. The
rift) horst and graben network. Abenaki sequences also have seismic-stratigraphic
The designation of the Abenaki depositional se- characteristics (as described above), supporting their
quences as second and third order is done in broad designation as depositional sequences. The eustatic
414 / Weissenberger et al.

FIGURE 16. Time structure on Abenaki Formation horizons, deep Panuke, and southwest. (A) Top Abenaki sequence I
time structure. Relatively widely spaced isochrons indicate ramp profile at margin. (B) Top Abenaki III time structure.
Tighter isochrons near the margin indicate a steeper depositional profile. A discernable similarity exists in the position of
the platform margin from the AB I to the AB III (e.g., the embayment near G-32 and south of H-08), showing an
inheritance of paleogeography through the section, likely reflecting deeper structural control.

sea level curve of Haq et al. (1987; partly reproduced sinkholes (G. Syhlonyk, 2005, personal commu-
in Figure 4) defines six supercycles (19 third-order nication). We have examined the similar aged Can-
sequences) throughout a similar time span. Therefore, deiros formation carbonates in central Portugal. These
the Abenaki sequences may more prudently have been have well-developed karst sinkholes, tens of meters
termed ‘‘composite sequences,’’ or supercycles, but deep and wide, as well as bedding-parallel dissolu-
their ultimate importance to the exploration and tion (paleowater table?). These data support the
development program at deep Panuke was their util- model of periodic subaerial exposure of the Abenaki
ity in defining correlatable time lines across the Sco- platform.
tian Shelf and within the pool. The nature of the alternation between siliciclastic-
The sequence-stratigraphic model postulates ex- and carbonate-dominated deposition on the Scotian
posure of the platform at the third-order sequence Shelf is seen in two cored wells to the northeast
boundaries. None of these boundaries are cored, so of deep Panuke: Peskowesk A-99 and Citadel H-52
other data (e.g., FMI images and internal facies archi- (Figure 17). The former displays a change from dom-
tecture) are used to define the third-order surfaces. inantly carbonate to siliciclastic deposition. The base
Eliuk (1978) proposed periodic exposure of the of the core consists of a dark-gray oolitic packstone
platform, with the development of freshwater lenses, with crinoids and shell hash, representing a middle
citing evidence of karst from the G-32 well (which he foreslope setting. Above are fissile calcareous shales
linked to dolomitization). becoming nodular and fossiliferous toward the top
Three-dimensional seismic slices of the carbonate (crinoids and corals), representing a transition from
platform in the deep Panuke area commonly show basin to lower foreslope environments. The top of the
circular depressions, 100 – 700 m (330 –2300 ft) in latter unit is extensively burrowed and mineralized,
diameter, which have been interpreted as karst as is the top of the overlying thin crinoid-bivalve
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 415

FIGURE 17. Mixed carbonate-siliciclastic core from two wells northeast of deep Panuke. Cores show switching from one
dominant sediment type to another after major depositional hiati.
416 / Weissenberger et al.

wackestone. This succession is interpreted as trans- reworked during the subsequent transgression. This is
gressive, with the borrowed and bored contacts at the overlain by 30 cm (12 in.) of oolitic grainstone coral-
tops of the two shoaling-upward cycles representing stromatoporoid-sponge packstone, reflecting trans-
significant depositional hiati at two marine flooding gression above the boundary.
surfaces. The base of the overlying succession is a
black, fissile, occasionally sideritic shale. This is over- DIAGENESIS
lain by a thin, fine-grained sandstone, followed by shales,
silts, and finally, fine- to medium-grained sand. This Understanding diagenetic processes is critical for
is interpreted as a basinal to lower shoreface-offshore the correct evaluation of the deep Panuke reservoir.
transition. This is sharply overlain by planar to tabular, Examination of drill cuttings from PP-3C, immedi-
parallel-laminated, medium- to coarse-grained sand- ately above the cavernous porosity zone, revealed the
stone with common Skolithos, interpreted as upper first Abenaki dolomite in coral-stromatoporoid pack-
shoreface deposits above a sequence boundary. stones, suggesting that dolomitization may have
The succession in Citadel H-52 begins with silic- been involved in vuggy porosity creation. However,
iclastics, shales, and silts passing upward into fine- to logs from PP-3C indicated that the cavernous po-
medium-grained, parallel-laminated sandstone, with rosity was mostly in limestone. Detailed petrogra-
a burrowed upper surface. Above is a dark-gray to phy, isotope, and fluid-inclusion analyses were
black, very fine silt to sandstone, with Planolites and conducted on the subsequent wells and presented
Terebellina. The top of the unit is irregular and miner- by R. A. Wierzbicki, J. J. Dravis, I. Al-Aasm, and N. J.
alized. This succession is interpreted as an offshore Harland (2005, personal communication). This work
transition to middle (and finally, lower) shoreface and some additional data are discussed below.
setting with a major submarine hardground at the Early diagenesis is characterized by grain suturing
top. Above the hardground is a fossil lag with abraded in grainstones and grainy packstones, suggesting lit-
coral fragments, and bivalve fragments, overlain by tle cementation before compaction. Stylolites and
fissile to nodular calcareous shale. This is, in turn, wispy microstylolites are common. However, abun-
overlain by a thick branching coral-abraded stro- dant secondary limestone porosity is preserved in the
matoporoid wackestone, with crinoids, bivalves, and Abenaki at relatively deep burial depths. This poros-
sponges, capped by a massive, medium gray-brown, ity was observed to cut stylolites. The porosity is pre-
peloidal wackestone to packstone, with crinoids and served along stylolites, and stable skeletal grains (e.g.,
bivalve fragments. The youngest interval is inter- echinoids, calcispheres, corals, stromatoporoids, etc.)
preted to represent basinal to middle-foreslope en- are partially dissolved, and calcite cement in healed
vironments, initiated by an initial marine flooding fractures does not extend into adjacent secondary mi-
event. croporosity. These fabrics suggest the secondary lime-
These two cores demonstrate that carbonate and stone porosity formed after pressure solution and at
siliciclastic environments coexisted on the Abenaki least some of the fracturing. The observed porosity was
shelf. However, although there are certainly mixed dominantly vuggy and moldic, the latter typically
lithologies (e.g., ooids with quartz nuclei, calcareous reflecting dissolution of micritic peloids or micritized
sandstones), there is commonly a switching of dom- skeletal grains.
inant sediment type at a major event, such as a depo- Dolomites are dominantly fine to medium crys-
sitional hiatus, significant transgression, or drop in talline, nonferroan, and subhedral to anhedral in
sea level. Prather (1991) describes in detail similar shape. This dolomite appears to be mostly fabric selec-
mixed carbonate-siliciclastic environments from the tive, replacing grainy limestone textures (packstones-
Jurassic of the Baltimore Canyon. grainstones) proximal to the platform margin. Sad-
Such a surface is interpreted at a fourth-order cycle dle dolomite was observed as a replacive mineral and
(likely parasequence set, sensu Van Wagoner et al., pore-filling cement. The matrix dolomites overlie
1988) from the M-79 well, illustrated in Figure 18. pressure-solution seams and massively replace relict
An irregular surface, observed at the top of a coral- sutured grain fabrics. Porosity in the dolomite section
stromatoporoid packstone, is interpreted to represent is vuggy and intercrystalline. Vuggy porosity cross-
a platform margin environment. This is abruptly over- cuts pressure solution seams in dolomites, with many
lain by 25 cm (10 in.) of fine-grained quartz sand- pores appearing to be molds or partial molds. Saddle
stone, interpreted to have been deposited during ex- dolomites and dolomitized echinoids were also dis-
posure of the platform at a sea level lowstand and solved. This suggests that the pores may have formed
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 417

FIGURE 18. Formation image log (FMI) of a fourth-order sequence boundary from the Abenaki V of the M-79 well.
Irregular surface capping grainy coral-stromatoporoid packstone (3884.9 m; 12,745.7 ft) is overlain by sandstone lag,
oolitic grainstone, and finally, argillaceous coral-stromatoporoid-sponge packstone.

as a result of the dissolution of dolomitized relict fractures and swarms of short macrofractures), whereas
grains or cement, instead of relict limestone. horizontal natural and crosscutting fractures sug-
Fracturing is common in the Abenaki, normal ver- gest active wrench faulting. Vertical fractures com-
tical burial and tectonic fractures (including micro- monly cut stylolites, suggesting their burial origin.
418 / Weissenberger et al.

FIGURE 19. Thin-section photomicrographs of Abenaki Formation reservoir, porosity, and diagenetic fabrics. (A) Secondary
vuggy porosity associated with emplacement of replacement dolomite, as commonly observed in the Abenaki dolomitic
limestones (H-08; 3448.6 m [11,314.3 ft]; plane polarized light, 5.5 mm [0.2 in.] field of view); (B) Well-developed
micromoldic porosity in ooids and peloids, interpreted to result from deep burial conditions (F-09, 3520 m [11,548 ft];
diffused plane polarized light, 3.0 mm [0.12 in.] field of view); (C) Partial dissolution of saddle dolomite (replacing
recrystallized limestone, with partial occlusion of porosity in saddles by nonferroan calcite (arrows). Such partial dis-
solution of dolomites is common in Abenaki dolomitic limestones (H-08, 3448.6 m [11,314.3 ft]; plane polarized light,
3.0 mm [0.12 in.] field of view); (D) A curved (saddle) dolomite rhomb adjacent to a pinpoint vug (PI-1A, 4029.5 m
[13,220.1 ft], 3.5-mm [0.13-in.]-wide field of view). Original descriptions by J. Dravis.

Dissolution of dolomite crystals along fracture planes zation and cementation at depth caused by higher
suggests that the fractures delivered diagenetic fluids. burial temperatures. Matrix dolomites were enriched
Well tests from deep Panuke suggest an effective frac- in Sr87, whereas saddle dolomites were nonradiogenic,
ture network, with fractured dolomites in the AB V as implying mixing of a radiogenic (juvenile) and non-
much as 1000 m (3300 ft) wide. radiogenic (basinal) source. Fluid inclusions in cal-
Figure 19 illustrates some of the key diagenetic cite cements showed homogenization temperatures
textures of the Abenaki in a series of photomicro- ranging from 60 to 1818C, with salinities of less than
graphs. The samples show evidence for secondary po- 0 to 13.9 wt.% NaCl, i.e., saline to nonsaline fluids
rosity development related to partial dolomitization, acting in a warm to hot regime. Dolomite cement
small moldic porosity in limestones, and saddle do- (saddles and others) inclusions reveal temperatures
lomites (19A, B, and C/D, respectively). from 85 to 1478C, with salinities from 7.6 to 12.0 wt.%
Geochemical analyses (R. A. Wierzbicki, J. J. Dravis, NaCl. This suggests that warm to hot, moderately
I. Al-Aasm, and N. J. Harland, 2005, personal commu- saline waters were involved in dolomitization.
nication) of the dolomites revealed carbon isotopes, Evidence for faulting during diagenesis includes
which were not depleted, and oxygen isotopes that vertical stylolites, natural horizontal fractures, and
were relatively highly depleted. This suggests a lack of twinned calcite. Helium is also present (0.02–0.03%)
organic influence on dolomitization and recrystalli- in the deep Panuke gas, and minor sphalerite is observed
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 419

FIGURE 20. Map of limestone and dolomite porosity trends in the central part of the deep Panuke field. Interpretation
made from amplitudes extracted from the 3-D seismic volume.

in the reservoir. As mentioned above, the presence of PETROLEUM GEOLOGY


juvenile fluids is suggested by the strontium isotope
data. Vitrinite reflectance data from the M-79 well The PP-3C discovery well was drilled on the basis
also showed a marked increase from 0.8 to 1.10 at of the aforementioned predrill stratigraphic model;
about 4260-m (13,976-ft) drill depth, suggesting a the location picked from a 1990 3-D seismic survey. An
thermal anomaly. Regionally, evidence for the pres- arbitrary dip line from that survey, near the discovery
ence of hot, fault-borne fluids includes zebra dolo- location, is shown in Figure 22A, with a schematic in-
mite textures in the core of the Iroquois Formation terpretation in Figure 22B. The apparent topography
from the P-23 well. in the basement is interpreted as a basement horst
These data suggest that deep burial and hydro- and associated half graben to the west, the latter
thermal fluids were responsible for limestone and filled by synrift deposits. These are overlain by Mohi-
dolomite porosity at deep Panuke. High-porosity can Formation siliciclastics. Three basic seismic facies
vuggy limestones in the PP-3C and H-08 wells occur are discernable from the data: stratified platform-
as linear trends behind the platform margin, ap- interior (lagoon), chaotic reflectors at the platform
parently influenced by wrench faulting. Vuggy do- margin, and basinward-dipping reflectors in the
lomites occur as curvilinear trends subparallel to the foreslope.
platform margin (Figure 20). Figure 21 shows the The chaotic margin facies is between 500 and
stratigraphic distribution of these same porous lime- 1200 m (1640 and 3900 ft) wide in a dip sense. Seismic
stone and dolomite intervals. modeling indicated that some of the reflectors in the
420 / Weissenberger et al.

FIGURE 21. Cross section of the deep Panuke wells, showing main occurrences of vuggy limestone and dolomite porosity. Color legend: light blue is porous
limestone; purple is porous dolomite; dark blue is tight limestone; orange stippling reflects no significant vuggy porosity. Diagram is modified from Wierzbicki
and Harland (2004).
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 421

FIGURE 22. Arbitrary dip line from the Panuke 3-D seismic survey (A) and schematic interpretation of the same,
illustrating the end of the Abenaki sequence V (B). Interpretation after G. Syhlonyk.
422 / Weissenberger et al.

margin facies represent a porosity anomaly. This is


corroborated by a velocity sag in reflectors below
the margin, representing the effect of slower velocity
through the overlying porosity. Bright reflectors in
the adjacent foreslope were (later) interpreted as sands,
which had bypassed the platform during sea level
lowstands.

Exploration Drilling Phase


The PP-3C well was intended to test such a
velocity anomaly. Beneath the Cretaceous (Mic
Mac–Missisauga Formation) siliciclastics, the well-
encountered gas shows in calcareous shales and mud-
stones of the upper Abenaki. Upon penetrating less
than 20 m (66 ft) of clean limestone, a lost circu-
lation zone occurred at 3934 m (12,906 ft) drill depth.
Although only 0.1 m (0.3 ft) of porosity had been
drilled, the formation continued to take drilling fluid,
including two plugs of lost circulation material (LCM).
Gas and condensate were flared. Introduction of a
third LCM plug sealed off the formation, and the
well was suspended.
Given the likelihood of penetrating further cav-
ernous porosity, it was decided to drill out using the
annular velocity control (AVC) drilling method, which
had not previously been attempted on the Scotian
Shelf. This entailed pumping sufficient sea water
into the wellbore while drilling to control the for-
mation. No drill cuttings can be recovered using
this method. Once drilling recommenced, several bit
drops of decimeters to more than 1 m (3.3 ft) were
encountered, suggesting further significant porosi-
ty. More than 200 m (660 ft) of Abenaki carbonate
was drilled in this manner, to a final total depth of
4163 m (13,658 ft).
Logging revealed a porosity zone 44 m (144 ft)
thick in the AB V, which was interpreted from logs to
represent predominantly vuggy dolomitic limestone
porosity. As mentioned above, dolomite had been
observed in the drill cuttings immediately above the
lost circulation zone (a coral-stromatoporoid pack-
stone; Weissenberger et al., 2000). The variability of
the porosity seen on logs (Figure 23) was interpreted
to reflect a complex pore system, likely including
vugs, caverns, and matrix.
Testing yielded a maximum flow rate of 69 mmcf/
day, limited by the equipment employed. Wet gas
was produced, with 600 ppm of hydrogen sulfide
(H2S). Also observed was a relatively slow buildup that
may reflect formation damage and possible reser- FIGURE 23. Well log of the Panuke PP-3C discovery well
voir phase changes caused by the introduction of showing cavernous porosity in the Abenaki V.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 423

Table 1. Abenaki Formation penetrations, deep Panuke.*


Well Name and Spud Top AB V Total Depth Net Pay Maximum mmcf/
Number Total Vertical (Depth and Test Rate day
Depth, Subsea Formation) (106 m3/day)
Measured Depth

Panuke F-09 August 23, 2000 3255.1 m 3815 m Abenaki 4 0m Flowed gas, TSTM 0
Panuke H-08 May 24, 2000 3348.5 m 3682 m Abenaki 4 104.7 m 1.55 55
Panuke M-79 July 10, 2000 3351.5 m 4597 m Abenaki 1 0m No test NT
Panuke M-79A October 11, 2000 3331.2 m 3934.2 m Abenaki 5 11.4 m TVD 1.75 62
Panuke PI1A-J99 August 27, 1999 3323.5 m 4033 m Abenaki 4 1.5 m No test NT
Panuke PI1B-J99 November 11, 1999 3322.2 m 4046.4 m Abenaki 4 13.9 m 1.49 53
Panuke PP3C J-99 July 17, 1998 3315 m 4163 m Abenaki 4 44.6 m 1.96 69
Margaree F-70 May 21, 2003 3330 m 3677 m Abenaki 4 70 m 1.5 53
Marcoh D-41 August 28, 2003 3306 m 3625 m Abenaki 4 74.2 m No test NT
Queensland M-88 December 14, 2001 3737.4 m (eq) 4443 m Abenaki 2 0m No test NT

*Well information summary table. Porosity cutoff for net pay calculation is 3%.

abundant LCM into the reservoir, as well as more M-79 wellbore did encounter significant, wet porosity
than 300,000 bbl of seawater pumped into the for- lower in the Abenaki.
mation during AVC drilling. Reservoir temperature Two further wells investigated the northeast ex-
dropped simultaneously, from 130 to 808C in the tent of the pool: F-70 is another gas well, analogous
wellbore. to M-79A; and D-41, which was drilled in 2004, was
still confidential at the time of writing. Two other
wells, F-09 and M-88, were drilled off the main AB V
Delineation and Predevelopment Drilling Phase bank margin during the delineation phase of the proj-
The PI-1 well tested the extent of the deep Panuke ect. F-09 drilled 70 m (230 ft) of interbedded carbon-
discovery to the northeast (Figure 5). It was drilled ate muds and siliciclastics in the AB VI (Figure 24).
on a seismic porosity anomaly analogous to PP3-C The underlying oolitic AB V interval had some gas
interpreted in the AB V. The well (PI-1A) encountered shows, but was too tight to be economic. The M-88 well
significant gas shows, some dolomite, but no cavern- was drilled basinward of the Abenaki platform mar-
ous porosity. Instead of testing the AB V, the seismic gin. It encountered bypassed siliciclastic sands in the
data was reinterpreted, and a sidetrack wellbore (PI-IB) interpreted lowstand of the AB VII (Figure 26).
drilled 70 m (230 ft) to the northeast. PI-1B penetrated A new 3-D seismic survey, 450 km2 (173 mi2) in
moderate, dolomitic limestone porosity in the AB V. area, was shot across the deep Panuke area in 2002.
The well was completed and tested gas at significant These data were used to locate the subsequent wells
rates (Table 1). and to model the porosity in the pool for reserves
The dimensions of the pool to the southwest were determination and depletion planning.
then tested by the H-08 well. The well encountered a Predevelopment geoscience and engineering
thick section of cavernous porosity in the AB V. An- work on deep Panuke was presented in Brown et al.
nular velocity control drilling was again necessary to (2004). To summarize, shear sonic logs from eight
control the formation. Ultimately, logs showed 104 m wells, vertical seismic profile data, high-quality seis-
(341 ft) of pay at a slightly lower structural eleva- mic processing of the new 3-D data set, including
tion than PP-3C and PI-1A (Figure 24) and tested gas Post-stack Time Migration, Post-stack Depth Migra-
(Table 1). A further delineation well at M-79 had a tion (PSDM), and Compressional + Shear Wave imped-
similar result to PI-1A (no significant AB V porosity in ances and Lambda-Mu-Rho volumes, were used to
the initial wellbore). However, M-79A, drilled 600 m image porosity in the pool. Seismic reservoir charac-
(1968 ft) to the southwest, encountered extremely terization used a supervised neural network approach
productive gas pay (Figure 24). The seismic section with seismic attributes from PSDM and amplitude-
(Figure 25) shows M-79 drilled behind the interpreted versus-offset cubes as input nodes to analyze basin,
platform margin, M-79A penetrating a significant backreef, and reefal parts of the pool. Accuracy of
interval of porous dolomite closer to the margin. The porosity prediction was then tested by matching well
424 / Weissenberger et al.

FIGURE 24. Structural cross section, deep Panuke pool, indicating distribution of conventional and sidewall core data.
Diagram modified from Wierzbicki and Harland (2004).
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 425

FIGURE 25. Seismic dip line through the M-79 and M-79a wells, deep Panuke pool. The platform margin is aggradational
from the AB II upward.

porosity to seismic data and defining a set of sig- used to history match the different reservoir models,
nificant seismic attributes. The attributes were first thereby reducing uncertainty, particularly in perme-
extracted from a 3  3 trace area around each well- ability estimates for the pool. Figure 27 shows one
bore and then from the entire volume and fed into such 3-D porosity model for the deep Panuke pool,
the trained neural net. with an extracted seismic profile.
Numerous field-scale porosity-permeability mod- A geostatistical approach was also used to estimate
els were developed by integrating geological and geo- the likely reserve size distribution of the pool (Deutsch
physical interpretations in a geostatistical framework. et al., 2003). Geological and geophysical inputs (the
These data were integrated with engineering infor- latter from the 2002 3-D data set) were used to real-
mation into a ‘‘shared earth model, containing the istically represent the reservoir heterogeneity.
full 3D representation of the reservoir’’ (Brown et al.,
2004, p. 3). Numerous property models, generated Critical Elements
using seismic attributes, well log, and core data, were The deep Panuke pool is interpreted as a com-
simulated and compared to well test data. This was an bined stratigraphic-structural trap. A common gas-
iterative process, testing the validity of each reservoir water contact exists at 3504 m (11,496 ft) subsea
model. The five wells in the pool with production (Figure 24). Top seal is provided by argillaceous car-
tests provide large-scale permeability data, which was bonates and shales in the upper Abenaki (Abenaki
426 / Weissenberger et al.

FIGURE 26. Geological dip


cross section interpreting plat-
form to basin transition be-
tween the M-79 and M-88
wells. Note bypassed quartz
sands in M-88, just below
3600 m (11,811 ft).

the Abenaki platform. Two


hypotheses for the origin of
the gas were contemplated:
biogenic production caused
by the prolonged interac-
tion of sea water with the
reservoir (while suspended
and AVC drilling) and mi-
gration from anhydrite-rich
strata where thermochemi-
cal sulfate reduction (TSR)
had occurred. Sulfur isotope
analyses were undertaken
on a sample of the gas and
on a sample of anhydrite
from the Eurydice Forma-
tion of the Mohican I-100
well (Figure 4). The analyses
(I. Hutcheon, 1999, personal
communication) indicate
that the sulfur in the deep
sequence VI and VII) and Mic Mac–lower Missisauga Panuke gas matches the isotopic signature in the
formations. Lateral seal is provided by these same Eurydice Formation anhydrite. Subsequent gas anal-
formations and shale- and mud-filled reentrants in ysis of wells where there was no significant fluid
the platform at the northeast end of the pool, as well influx showed the same H2S concentrations. It is
as tight, platform interior oolitic limestones. A re- likely that the observed sulfur concentrations are
gional structural doming also exists along the Abe- related to TSR processes.
naki carbonate margin, which supports the east–west
closure of the pool.
The source of the gas is believed to be organic CONCLUSIONS
shales in the Verrill Canyon Formation (Williamson
and Des Roches, 1993). The principal reservoir zones Figure 28 illustrates the tectonostratigraphic evo-
are reefal boundstones and associated sediments, hav- lution of the Jurassic carbonate hydrocarbon system
ing both intercrystalline and vuggy porosity in dolo- on the Scotian Shelf. During the Early Jurassic, sea
mites and secondary and vuggy porosity in limestones. level rise allowed the deposition of the Iroquois For-
The diagenesis responsible for this porosity was dis- mation, a broad, shallow-water carbonate shelf over
cussed above. synrift deposits of the Argo and Eurydice formations
(Figure 28A). The latter was largely restricted to topo-
graphic lows (half grabens) between tilted fault blocks.
Discussion Overlying the Iroquois, a thick section of siliciclastics
When small amounts of H2S were encountered in was deposited in the transition from the Lower to
tests of the PP-3C well, it raised the immediate ques- Middle Jurassic (Mohican Formation; Toarcian to Ba-
tion of provenance and possible extent of sour gas jocian). These strata are provisionally defined as a
in any further wells drilled or discoveries made on second-order depositional sequence.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 427

FIGURE 27. Derived porosity distribution, Abenaki sequence V, deep Panuke pool. Seismic section shown is taken just
north of the M-79 well. The red and black colors indicate amplitudes interpreted as the highest porosity zones in the
porosity model. Interpretation by R. Tonn.

Later in the Middle Jurassic (Bajocian – Callovian; in the AB VI and VII. Northeast of deep Panuke, sedi-
Figure 28B) the first Abenaki third-order deposition- mentation was dominated by siliciclastics.
al sequence (AB I) was deposited above the Mohican From the Early to middle Cretaceous, sedimen-
Formation with continued relative sea level rise. The tation was dominated by siliciclastics (Figure 28E).
second Abenaki sequence (AB II) comprises two parts: Tectonic activity in the region began in the Aptian
the basinal Misaine Formation (transgressive systems (Jansa and Pe-Piper, 1988). This caused the reacti-
tract) and the overlying foreslope and shallow-water vation of basement faults in a strike-slip sense and
carbonates (highstand systems tract). The Misaine rep- the movement of hydrothermal fluids into the Abe-
resents the MFS of both the third-order AB II and the naki carbonate platform. These fluids completed the
second-order depositional sequence. Figure 28B il- creation of secondary limestone and dolomite po-
lustrates some of the stratal geometries that are ap- rosity begun earlier in the burial history (R. A. Wierz-
parent on seismic data, and the possibility of reefs bicki, J. J. Dravis, I. Al-Aasm, and N. J. Harland, 2005,
initiating on the drowned AB I. personal communication). The Verrill Canyon source
Beginning in the Upper Jurassic, the stratal archi- rock entered the gas-generation window in the Ap-
tecture of the Abenaki platform was essentially ag- tian (Williamson and Des Roches, 1993), likely filling
gradational. The coral-stromatoporoid reef margin the deep Panuke trap by the Albian.
became well established (Figure 28C). During rela- Drilling since 1998 has proven a large accumula-
tive lowstands in sea level, siliciclastics from the plat- tion of natural gas at deep Panuke, extending 26 km
form interior bypassed into the basin. Minor epikarst (16 mi) along the Jurassic carbonate platform margin.
developed during these lowstands. Carbonate plat- EnCana has not published a revised reserves estimate
form deposition continued through the lowermost since 2002, despite the drilling of several delineation
Cretaceous (Berriasian; Figure 28D). The large-scale wells. It remains the first and only significant hydro-
stacking pattern was aggradational to the top of the carbon discovery in the Mesozoic carbonates on the
AB V in the deep Panuke area, then retrogradational entire continental shelf of eastern North America.
428 / Weissenberger et al.

FIGURE 28. Tectonostratigraphic


evolution of the carbonate plat-
form, deep Panuke area. (A) Early
Jurassic; Iroquois Formation car-
bonate platform deposited during
transgression over Triassic synrift
deposits. Early dolomite formation
in semiarid climate. (B) Bathonian–
Callovian; maximum flooding
surface of second-order deposi-
tional sequence (Misaine Forma-
tion) and progradation of Abenaki
II highstand carbonate. (C) Oxfor-
dian; exposure during a relative
sea level lowstand at the base of
an Abenaki third-order sequences
(e.g., AB IV) yields meteoric karst
and cross-bank transport (and
foreslope deposition) of siliciclas-
tics. (D) Kimmeridgian–Berriasian;
relative transgression in Abenaki VI
and VII sequences. Platform ar-
chitecture changes from domi-
nantly aggradational to retrogra-
dational in the deep Panuke area.
(E) Valanginian – Cenomanian;
Missisauga to Logan Canyon sili-
ciclastics deposited; subsidence of
forereef; regional wrench tecto-
nism moves hot diagenetic fluids
up reactivated normal faults.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 429

APPENDIX 1: NEW CORE AND SIDEWALL CORE DATA, DEEP PANUKE

Location Core Type Core Interval Number Recovery Analysis Done


of Side-wall (Whole Core)
Core
PP-3C None

PI-1A Whole core Whole core 4029.28 – 4040 m 45 1.15 m (3.77 ft) Routine
and sidewall (13,219.42 – 13,254 ft)

Side-wall core 3895 – 4034 m Routine


(12,778 – 13,234 ft)

PI-1B None

M-79 Whole core Whole core 4532.7 – 4538.7 m 99 5 m (16 ft) None
and sidewall (14,871.06 – 14,890.7 ft)

Side-wall core 3512.5 – 4591 m


(11,523.9 – 15,062 ft)

M-79A None

H-08 Whole core Whole core 3446 – 3460 m 3.2 m (10.4 ft) Routine and
(11,305 – 11,351 ft) special core

F-09 Sidewall Side-wall core 3264 – 3798 m 31 Routine


(10,708 – 12,460 ft)

F-70 Whole core Whole core 3434 – 3461 m 45 24.65 m (80.8 ft) Routine
and sidewall (11,266 – 11,354 ft)

Side-wall core 3350 – 3636 m Special core


(10,990 – 11,929 ft) analysis

APPENDIX 2: BIOSTRATIGRAPHIC SUMMARY, M-79 WELL, DEEP PANUKE

Depth (m) Age Assignment Cycle


3180 (top sample) to 3220 Early Cretaceous, Neocomian AB 7

3240 – 3360 Portlandian, anguiformis or older AB 6

3380 – 3440 Late Kimmeridgian, fittoni or older AB 6

3460 – 3820 Late to early Kimmeridgian, elegans or mutabilis or older AB 5/4/3

3840 – 4020 Late Oxfordian, rozenkrantzi or older AB 3

4040 – 4240 Middle Oxfordian, glosense to Callovian AB 2

4260 – 4460 Middle Callovian, coronatum and jason AB 2/Misaine

4480 – 4600 (total depth) Callovian AB 1 Scatarie

B. G. T. van Helden
430 / Weissenberger et al.

ACKNOWLEDGMENTS dian Society of Petroleum Geologists Memoir 13,


p. 713 – 720.
Eliuk, L. S., S. C. Cearley, and R. Levesque, 1986, West
We thank EnCana Ltd. for giving permission to
Atlantic Mesozoic carbonates: Comparison of Balti-
publish our work on deep Panuke and Husky Energy
more Canyon and offshore Nova Scotian basins: AAPG
for additional support in completion of this publica- Bulletin v. 70, p. 586 – 587.
tion. This paper would not have been written without Ellis, P. M., 1984, Upper Jurassic carbonates from the
the contribution of our many colleagues and associ- Lusitanian basin, Portugal and their subsurface coun-
ates who worked on various aspects of the project, and terparts in the Nova Scotian Shelf: Ph.D. thesis, Open
we beg the pardon of those we do not mention. J. A. University, England, 283 p.
W. Weissenberger thanks John Hogg and Ian DeLong Ellis, P. M., P. D. Crevello, and L. S. Eliuk, 1985, Upper Ju-
rassic and Lower Cretaceous deep-water buildups, Abe-
for the opportunity to work with them on the project
naki Formation, Nova Scotian Shelf, in P. D. Crevello
from the beginning. We must single out the contri- and P. M. Harris, eds., Deep-water carbonates: SEPM
butions of our geophysicists, Garth Syhlonyk, Robert Core Workshop 6, p. 212 – 248.
Riddy, and Rainer Tonn, who worked through the Ellis, P. M., R. C. L. Willson, and R. R. Leinfelder, 1990,
delineation and development phases, as well as en- Controls on Upper Jurassic carbonate buildup devel-
gineers Tom Craig and John Slade. We must also ac- opment in the Lusitanian basin, Portugal: Interna-
knowledge several colleagues who supported this proj- tional Association of Sedimentologists Special Publi-
ect through their specialized work: Ihsan Al-Aasm cation 9, p. 169 – 202.
Given, M. M., 1977, Mesozoic and early Cenozoic geology
(isotopes and fluid inclusions), Jeff Dravis (petrogra-
of offshore Nova Scotia: Bulletin of Canadian Petro-
phy and diagenesis), Les Eliuk and Bev Harris (drill- leum Geology, v. 25, p. 63 – 91.
cuttings description), Bert van Helden (micropaleon- Golonka, J., M. I. Ross, and C. R. Scotese, 1994, Phanerozoic
tology), and Ian Hutcheon (sulfur isotopes). Thanks paleogeographic and paleoclimatic modeling maps, in
again to John Hogg for critically reading the manu- A. F. Embry, B. Beauchamp, and D. J. Glass, eds., Pangaea,
script and to Phil Argatoff of Penta Graphix Ltd., who global environments and resources: Canadian Society
prepared many of the diagrams. The authors also ap- of Petroleum Geologists Memoir 17, p. 1 – 47.
Gradstein, F. M., M. A. Kaminski, and F. P. Agterberg, 1999,
preciate the helpful comments of reviewers Sherry
Biostratigraphy and paleoceanography of the Creta-
Becker, Niall Toomey, and Mitch Harris, which greatly
ceous seaway between Norway and Greenland, in F. M.
improved the manuscript. Please note that the refer- Gradstein and G. J. van der Zwaan, eds., Ordering the
ence to proven reserves in the AAPG 2006 conference fossil record; challenges in stratigraphy and paleontol-
abstracts is withdrawn because it was included in error. ogy: Selected papers from a symposium held in honor
of the 75th birthday of Cor Drooger: Earth Science Re-
REFERENCES CITED views, v. 46, no. 1 – 4, p. 27 – 98.
Haq, B. U., J. Hardenbol, and P. R. Vail, 1987, Chronology
Blomeier, D. P. G., and J. J. G. Reijmer, 2002, Facies archi- of fluctuating sea levels since the Triassic: Science,
tecture of an Early Jurassic carbonate platform slope v. 235, p. 1156 – 1167.
(Jbel Bou Dahar, High Atlas, Morocco): Journal of Sedi- Harland, N. J., J. R. Hogg, R. Riddy, G. Syhlonyk, G.
mentary Research, v. 72, no. 4, p. 462 – 475. Uswak, J. A. W. Weissenberger, and R. A. Wierzbicki,
Brown, S., R. Riddy, R. Tonn, and R. A. Wierzbicki, 2004, 2002, A major gas discovery at the Panuke field, Ju-
The integration of geology, geophysics and reservoir rassic Abenaki Formation, offshore Nova Scotia (abs.):
engineering for field appraisal (abs.): Canadian Society Canadian Society of Petroleum Geologists, Annual
of Exploration Geophysics Annual Meeting Abstracts Meeting, Calgary, p. 154.
and Program, 4 p. Hogg, J. R., and M. E. Enachescu, 2003, An overview of the
Deutsch, C., R. A. Wierzbicki, R. Riddy, and J. Slade, 2003, Grand Banks and Scotian Shelf basins development
Quantification of uncertainty in gas resources of deep and exploration, offshore Canada (abs.): AAPG Interna-
Panuke (abs.): AAPG Annual Meeting Program, v. 12, tional Conference and Exhibition, Barcelona, Spain,
6 p., CD-ROM. CD-ROM.
Eliuk, L. S., 1978, The Abenaki Formation, Nova Scotia Jansa, L. F., 1993, Early Cretaceous carbonate platforms of
Shelf, Canada — Depositional and diagenetic model the northeastern North American margin, in J. A. T.
for a Mesozoic carbonate platform: Bulletin of Cana- Simo, R. W. Scott, and J. P. Masse, eds., Cretaceous car-
dian Petroleum Geology, v. 26, p. 424 – 514. bonate platforms: AAPG Memoir 56, p. 111– 126.
Eliuk, L. S., and R. Levesque, 1989, Earliest Cretaceous Jansa, L. F., and G. Pe-Piper, 1988, Middle Jurassic to Early
sponge reef mound, Nova Scotia shelf (Shell Demas- Cretaceous igneous rocks along eastern North Amer-
cota G-32), in H. H. J. Geldsetzer, N. P. James, and G. E. ican continental margin: AAPG Bulletin, v. 72, no. 3,
Tebbutt, eds., Reefs, Canada and adjacent areas: Cana- p. 347 – 366.
Sequence Stratigraphy and Petroleum Geology of the Jurassic Deep Panuke Field / 431

Jansa, L. F., and J. A. Wade, 1975, Paleogeography and sedi- An overview of the fundamentals of sequence stratigra-
mentation in the Mesozoic and Cenozoic, southeastern phy and key definitions, in C. K. Wilgus, B. S. Hastings,
Canada, in W. J. M. van der Linden and J. A. Wade, eds., C. G. St. C. Kendall, H. W. Posamentier, C. A. Ross, and
Offshore geology of eastern Canada: Geological Survey J. C. Van Wagoner, eds., Sea level changes: An inte-
of Canada Paper 74-30, p. 51– 105. grated approach: SEPM Special Publication 42, p. 125 –
Jansa, L. F., and J. Wiedmann, 1982, Mesozoic – Cenozoic 154.
development of the eastern North American and north- Wade, J. A., 1977, Stratigraphy of the George’s Bank basin—
west African continental margins — A comparison, in Interpreted from seismic correlation to the western Sco-
U. von Rad, K. Hinz, M. Sarnthein, and E. Seibold, eds., tian Shelf: Canadian Journal of Earth Sciences, v. 14,
Geology of the northwest African continental margin: p. 2274 – 2283.
New York, Springer-Verlag, p. 215 – 269. Weissenberger, J. A. W., N. J. Harland, J. R. Hogg, and G.
Leinfelder, R. R., 1994, Distribution of Jurassic reef types: Syhlonyk, 2000, Sequence stratigraphy of Mesozoic
A mirror of structural and environmental changes carbonates, Scotian Shelf, Canada (abs.): GeoCanada
during the breakup of Pangaea, in A. F. Embry, B. Beau- 2000 Convention ( Joint Meeting of the Canadian
champ, and D. F. Glass, eds., Pangaea: Global environ- Geophysical Union, Canadian Society of Exploration
ments and resources: Canadian Society of Petroleum Geophysicists, Canadian Society of Petroleum Geolo-
Geologists Memoir 17, p. 677 – 700. gists, Canadian Well Logging Society, Geological
McIver, N. L., 1972, Cenozoic and Mesozoic stratigraphy Association of Canada, and the Minerological Associ-
of the Nova Scotia Shelf: Canadian Journal of Earth ation of Canada) CD disk (5 p., 4 figures).
Sciences, v. 9, p. 54 – 70. Welsink, H. J., J. D. Dwyer, and R. J. Knight, 1989,
Poag, C. W., 1991, Rise and demise of the Bahama – Grand Tectono-stratigraphy of the passive margin off Nova
Bank gigaplatform, northern margin of the Jurassic Scotia, in A. J. Tankard and H. R. Balkwill, eds., Exten-
proto-Atlantic seaway: Marine Geology, v. 102, p. 63 – sional tectonics and stratigraphy of the North Atlantic
103. margin: AAPG Memoir 46, p. 215 – 231.
Prather, B. E., 1991, Petroleum geology of the Upper Juras- Wenzel, A., and A. Strasser, 2001, Sedimentology, paleo-
sic and Lower Cretaceous, Baltimore Canyon Trough, ecology and high resolution sequence stratigraphy of
western North Atlantic Ocean: AAPG Bulletin, v. 75, a carbonate-siliciclastic shelf (Oxfordian, Swiss Jura
p. 258 – 277. Mountains): Field trip guidebook A3, International Asso-
Pratt, B. R., and L. F. Jansa, 1989, Late Jurassic shallow wa- ciation of Sedimentologists, Davos, Switzerland, 19 p.
ter reefs of offshore Nova Scotia, in H. H. J. Geldsetzer, Wierzbicki, R. A., and N. J. Harland, 2004, Diagenetic mod-
N. P. James, and G. E. Tebbutt, eds., Reefs, Canada and el: deep Panuke reservoir, offshore Nova Scotia, Canada
adjacent areas: Canadian Society of Petroleum Geolo- (abs.): AAPG Annual Convention Abstracts and Pro-
gists Memoir 13, p. 741 – 747. gram, CD-ROM.
Scotese, C. R., 1997, Paleogeographic atlas: PALEOMAP Wierzbicki, R. A., N. J. Harland, and L. S. Eliuk, 2002, Deep
Progress Report 90-0497, University of Texas at Arling- Panuke and Demascota core from the Jurassic Abenaki
ton, 20 p. Formation, Nova Scotia: Facies models, deep Panuke,
Vail, P. R., R. M. Mitchum Jr., and S. Thompson, 1977, Abenaki Formation (abs.): Canadian Society of Petro-
Seismic stratigraphy and global changes in sea level: leum Geologists Annual Meeting Core Convention
Part 4. Global cycles of relative changes of sea level, in Extended Abstracts, p. 71 – 101.
Seismic stratigraphy — Applications to hydrocarbon Williamson, M. A., and K. Des Roches, 1993, A maturation
exploration: AAPG Memoir 26, p. 83 – 97. framework for Jurassic sediments in the Sable subba-
Van Wagoner, J. C., H. W. Posamentier, R. M. Mitchum, sin, offshore Nova Scotia: Bulletin of Canadian Petro-
P. R. Vail, J. F. Sarg, T. S. Loutit, and J. Hardenbol, 1988, leum Geology, v. 41, p. 244 – 257.
12
Ramon, J. C., and A. Fajardo, 2006, Sedimentology, sequence stratigraphy,
and reservoir architecture of the Eocene Mirador Formation, Cupiagua
field, Llanos Foothills, Colombia, in P. M. Harris and L. J. Weber, eds.,
Giant hydrocarbon reservoirs of the world: From rocks to reservoir
characterization and modeling: AAPG Memoir 88/SEPM Special
Publication, p. 433 – 469.

Sedimentology, Sequence
Stratigraphy, and Reservoir
Architecture of the Eocene
Mirador Formation, Cupiagua
Field, Llanos Foothills, Colombia
Juan Carlos Ramon
BP Colombia, Bogotá, Colombia

Andres Fajardo
Chevron-Texaco, Bogotá, Colombia

ABSTRACT

T
he stratigraphic architecture and facies distributions in a high-resolution
time-space framework define the three-dimensional (3-D) reservoir zonation
of the Mirador Formation in the Cupiagua field. A high-resolution genetic
sequence stratigraphy study, using more than 731 m (2400 ft) of core and 40 well logs,
is integrated with petrophysical information to populate a static structural model
based on the interpretation of a 312-km2 (120-mi2) 3-D seismic volume. Dynamic data
(pressure, gas tracers, gas-oil ratio behavior) are integrated with the geological model
to better define sandstone bodies’ lateral continuity. Production logs (production
logging tools) complement petrophysical data in the definition of fluid-flow units.
Three scales of stratigraphic cycles are recognized in the Mirador Formation.
Short-term (high-frequency) cycles correspond to progradational-aggradational
units. Six intermediate-term cycles are identified by the stacking patterns of their
component short-term cycles and by the general trend of facies successions, in-
dicating increasing or decreasing accommodation-to-sediment supply (A/S) ratios.
Two long-term cycles are defined from the stacking pattern of the intermediate-
term cycles and by the general trend of facies successions.
The lower half of the Mirador Formation consists of coastal-plain facies tracts
and is composed of channel, crevasse splay, and swamp and flood-plain facies
successions. A bay facies tract occurs in the upper half of the Mirador Formation
and is composed of bay-fill, bay-head delta, and channel facies successions.

Copyright n2006 by The American Association of Petroleum Geologists.


DOI:10.1306/1215884M883276

433
434 / Ramon and Fajardo

The lower Mirador Unit was deposited over a wide flood-plain sequence.
Each intermediate-term cycle is composed of aggradational channel deposits, pro-
gradational and aggradational crevasse splay bodies, and aggradational swamp
and flood-plain facies successions. The first two intermediate-scale cycles (I, II)
show a seaward-stepping pattern, and then the next cycle (III) shows a landward-
stepping stacking pattern. The fall-to-rise turnaround is located at the base of
cycle II. The upper Mirador shows the continuous landward-stepping pattern and
places prograding bay-head delta and bay-fill facies successions over the alluvial-
plain setting of the Lower Mirador. This upper unit consists of three onlapping
cycles composed of a succession of aggradational channel deposits, prograda-
tional bay-head delta, and bay-fill deposits with a landward-stepping stacking
pattern. During this cycle, deepening-up bay-fill facies successions were deposited
in the area as a consequence of the increasing accommodation conditions that
prevailed in the area. Finally, the Mirador is capped by restricted-marine shales of
the Carbonera Formation.
The Cupiagua structure is a large, east-verging, asymmetric anticlinal fold
that trends north-northeast in the hanging wall of the frontal fault. Average
length and width of the Cupiagua structure are 25 and 3 km (15 and 1.8 mi),
respectively. The original oil in place in the Cupiagua field is estimated between
1000 and 1100 MMSTB of oil and 3000 to 4500 mmcf of gas. The Mirador For-
mation accounts for approximately 51% of the recoverable oil in the Cupiagua
field.

INTRODUCTION The Cupiagua field is a rich, near-critical gas conden-


sate reservoir over a very high (1800-m; 6000-ft)
The Cusiana and Cupiagua fields, discovered in hydrocarbon column. The fluid composition changes
1988 and 1992, respectively, are located about 150 km slightly with depth. The condensate yield is about
(93 mi) northeast of Bogota, Colombia, South Amer- 250 bbl/MMSCF of gas, and the average API gravity
ica (Figure 1). The fields lie in the foothills trend on is 408. The initial reservoir pressure was close to
the edge of the eastern Cordillera. The first oil pro- 6500 psia. The dew point pressure is 5350 psia.
duction from Cusiana began in September 1992 as The main drive mechanism is gas reinjection and
part of an early production scheme via the existing gas expansion. An efficient revaporization recovery
pipeline systems. Commerciality of Cusiana and a small mechanism exists. At mid-2005, the pore volume re-
central area of Cupiagua were declared in June 1993. placement is about 70%. A predicted decline of close
Phase I of the development of Cusiana allowed pro- to 30% yr1 is exacerbated by reservoir complexity,
duction to increase to 185 MBOPD during 1995. A scale, and fluid issues like gas recycling. Recent op-
second phase of the project was sanctioned during timization of the gas injection resulted in an average
1995. This consisted of an expansion of the Cusiana decline of about 23% yr1.
facilities, the construction of an 800-km (497-mi) pipe-
line between Cusiana and the oil terminal at the
Atlantic Coast (Coveñas) and the construction of the REGIONAL SETTING
Cupiagua central processing facilities. This second de-
velopment phase allowed production to progressively The Llanos Foothills are located between the un-
ramp up from 185 MBOPD to reach a peak through- deformed Llanos foreland grasslands and the high-
put in the first half of 1998 of 341 MBOPD, with gas elevation, highly deformed eastern Cordillera. The
handling rates in excess of 1.4 MMSCFD. During 1997, Llanos Foothills involves a zone of frontal deforma-
the Cupiagua South accumulation was discovered. tion running northeast for hundreds of kilometers
The Cusiana and Cupiagua field developments are and about 20 km (12 mi) wide (Cooper et al., 1995).
now close to complete, and the fields have entered The foothills are limited by the Guaicaramo and the
their decline phase. At mid-2005, the field has 48 wells, frontal-Yopal fault systems. The main outcropping
36 producers, and 12 gas-injector wells (Figure 2). structural feature is the Nunchia syncline.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 435

FIGURE 1. Location map of the


Cupiagua and Cusiana fields
in the Llanos Foothills. The
map shows the different fields
and their producing fluids.

Cretaceous strata are domi-


nated by shallow-marine, del-
taic, and basal transgressive
sandstones. Interbedded shales
increase in thickness and fre-
quency basinward. The Cre-
taceous cycle (Figure 3) rep-
resents a sedimentary wedge
with a central shale unit (Ga-
cheta Formation) enclosed
by transgressive sands (the
Une and Guadalupe forma-
tions, respectively).
The Tertiary sequence be-
gan with deposition of the
laterally amalgamated chan-
nel-belt systems of the Barco
Formation. The overlying Los
Cuervos Formation consists
of aggradational flood-plain
deposits and isolated single-
story channel belts. Paleocene
deposits are preserved in a
The stratigraphic column (Figure 3) may be divid- relatively narrow region along the foothills but have
ed into three major depositional cycles, which, in turn, been mostly removed by subsequent erosion in most
may be further divided. The first cycle covers the Pa- of the adjacent Llanos basin. The Eocene Mirador
leozoic, the second begins in the Albian–Cenomanian Formation transitionally overlies the Los Cuervos
and continues until the Paleocene (mainly Cretaceous mudstones. The Carbonera Formation comprises four
succession), and the third extends from the Eocene to regressive sandy units (labeled C1, C3, C5, and C7)
the present. The rock successions corresponding to intercalated with four transgressive shale units (C2,
these cycles are separated by the regionally significant C4, C6, and C8). The shale units can be correlated
sub-Cretaceous and sub-Paleocene unconformities. regionally, particularly the uppermost unit. The sandy
Early Cretaceous extension created a graben where units are interpreted to be nearshore, coastal plain,
the Cordillera Oriental mountain range is today, which and predominantly deltaic. The Carbonera Forma-
was filled with as much as 3 km (1.8 mi) of Lower tion, as with all previous formations, was deposited
Cretaceous marine sediments (Dengo and Covey, in a basin that extended to the west far beyond the
1993). The main Cretaceous transgression, which be- present-day Llanos (Villamil, 1999). The Carbonera
gan in the Aptian, entered from the north and spread Formation thickens steadily westward toward the
rapidly across the basin. These sediments were pri- mountain front, reaching a thickness of more than
marily derived from the Guyana shield to the east. 5000 ft (1500 m). The shaly Miocene Leon Forma-
Early Tertiary uplift and erosion have cut back the tion, which overlies the Carbonera Formation, pro-
present edge of Cretaceous sediments well west of vides the first indication of the uplift of the Cordillera
their original depositional limit. From west to east, Oriental and the isolation of the Llanos basin in the
progressively younger Cretaceous strata overlie pro- east from the Magdalena basin in the west. The thick-
gressively older Paleozoic rocks. est Leon Formation sediments are located east of
436 / Ramon and Fajardo

FIGURE 2. Well base map of


the Cupiagua field. Main struc-
tural features are shown. Pro-
ducing wells are in green, and
injector wells are in red.

more than 4000 m (13,000 ft).


These sediments are mainly
coarse-grained red beds of
continental origin.
The hydrocarbons were
generated toward the west
from the world-class source
rock La Luna – Gachetá For-
mation, buried more than
6000 m (20,000 ft) below the
main thick skin thrusting of
the eastern Cordillera.

CUPIAGUA STRUCTURE

The Cupiagua structure is


an elongated asymmetrical
anticline running north-
northeast for approximately
30 km (18 mi) (Figure 4). This
structure is an east-southeast-
propagating hanging-block
anticline associated with the
east-southeast-verging fron-
tal and core faults (Figure 5).
The regional detachments for
these faults are in the mud-
stone intervals of the Creta-
ceous Gacheta Formation.
the present basin edge, where they reach as much as The Cupiagua structure lies below the Yopal fault,
1000 m (3300 ft) in thickness. Farther westward, a which separates the Nunchia syncline from the un-
thinning trend (facies change into sandstones) is derlying structural deformation (Coral and Rathke,
apparent before the formation reaches the mountain 1997). The west flank is 2.5 – 3 km (1.5– 1.8 mi) wide
front. The Leon Formation consists of homogeneous with average dip of 35–388. Locally, it can get as much
shale across much of the central Llanos basin but is as 4 km (2.5 mi) wide. The forelimb is quite variable
rich in sandstone layers along the foothills. Coarser from south to north along the structure (Figures 4, 5).
clastics present in the west and northwest are in- In the south part of the Cupiagua structure, the fore-
terpreted as having been derived from the eastward limb is very high angle to inverted and is almost as
reworking of Carbonera sands exposed in island ridges high as the west flank (Figure 4B). To the north, the
ancestral to the Cordillera Oriental (Cooper et al., forelimb is structurally simpler, and finally, it disap-
1995; Villamil, 1999). The Miocene –Holocene Gua- pears in the middle part of the field (Figure 4C). Far-
yabo and Necesidad formations represent very thick ther north along the structure, the detachment position
mollasse deposits associated to the uplifting of the rises stratigraphically, and the whole structure bends
eastern Cordillera. This last cycle is a thick, easterly strongly toward the east. Figure 6 is a time slice of the
thinning wedge of sands and shales, which reaches Cupiagua three-dimensional (3-D) volume showing
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 437

FIGURE 3. Generalized stratigraphic column of the Llanos Foothills. The main reservoir sandstones of the Mirador, Barco,
and Guadalupe formations are shown.
438 / Ramon and Fajardo

FIGURE 4. Details of the Cu-


piagua structure in (A) 3-D
view. Note the northeast bend-
ing in the Cupiagua structure.
This change in strike is associ-
ated to a vertical feature that
affects the whole sedimentary
sequence and in geological
maps is called the Golconda
wrench. Structural cross sec-
tions across (B) the southern
and (C) central-northern parts
of the Cupiagua field show the
change in structural style along
strike (after Martinez, 2003).
In the south, the structure has
a high-angle to inverted im-
bricate. Toward the north, this
imbricate disappears, and the
structure becomes a simple
hanging-block anticline.

the east, and they formed


where the frontal thrust
bends (Figure 4C). Structural
mapping, fault-seal analysis,
the north–south trend of the syncline over the Cu- well-test data, production data, and gas tracer results
piagua structure. Toward the northern portion, the demonstrate that the faults in the anticline are not
syncline bends toward the northeast (Figure 6). Along compartmentalizing the field. A few exceptions oc-
the area of strike change, a lateral ramp or a later strike- cur where these faults act as baffles.
slip fault is interpreted. This lateral ramp corresponds Associated with the Golconda lateral ramp or wrench
to a surface lineament called the Golconda wrench in the central part of the structure are a series of east-
fault (Coral and Rathke, 1997). This feature affects to southeast-trending lineaments. These very high-
the whole sequence from surface to the Cretaceous angle faults cause minor displacements of the Car-
sequence, and the Cupiagua structure bends to the bonera strata and seem to extend down and cut the
northeast and then continues to the north (Figure 4). reservoir section. Cross-line 1510 (Figure 7) running
The western limit of the structure is a west-verging north-northeast shows several of these almost vertical
series of faults. In the southern half (south of the faults. The Golconda lineament dipping to the north
lateral ramp) of the field, there is a major backfault is the lateral ramp feature that causes the bend in the
that seems to cut completely the east-southeast- Cupiagua structure. South of this feature, an opposite
verging system of thrusts (Figure 4B). As seen in thrust exists that breaks the Cupiagua structure.
Figure 5A, the seismic reflectors below the frontal Several secondary faults occur both at the north and
fault are continuous, and the frontal fault looks broken south but close to the lateral ramp. Seismic quality
in two by the west-verging backthrust. This fault is in the reservoir section is poor, and it is hard to de-
interpreted to be a backthrust of a deeper thrust (deeper termine if the vertical faults cut all the way down.
than the frontal thrust) and to cut all overlying thrust Locally, dynamic data support the presence of some
faults. This backthrust seems to be genetically asso- of these faults and evidence of their seal or baffle
ciated with the Golconda lateral ramp, as it dies and character.
does not exist north from the ramp. On the contrary,
in the northern part of the field (north of the lateral RESERVOIR STRATIGRAPHY
ramp), the western limit of the field is defined by a
series of backthrusts of the frontal fault (Figure 5B). In the Cupiagua and Cusiana fields, the reservoir
These west-verging faults are genetically related to sandstones comprise Late Cretaceous to late Eocene
the propagation of the Cupiagua structure toward shallow-marine to alluvial sandstones, sourced from
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 439

FIGURE 5. Dip seismic sections


across the Cupiagua field.
Cupiagua wells close to each
line are shown. Inline 1200
(A) runs along the southern
part of the Cupiagua structure.
Note that the dominant struc-
tural feature is the Nunchia
syncline. The Yopal fault sepa-
rates two different structural
styles of deformation. In this
seismic line, it appears that the
back-main west-verging thrust
breaks the Cupiagua structure
and the frontal thrust. This
would suggest that the back-
fault is a younger event. This
back-main thrust extends only
south of the Golconda linea-
ment. Inline 1900 (B) is located
in the northern part of the
Cupiagua structure. Note that
the Nunchia syncline axis and
the crest of the Cupiagua
structure move eastward. The
west-verging faults are, in
this case, associated to the
frontal fault.

This chapter concentrates


on the sedimentology and
the stratigraphy of the Mir-
ador Formation.

FACIES, FACIES
SUCCESSIONS, AND
FACIES TRACTS IN THE
MIRADOR FORMATION

Two facies tracts exist in


the Mirador Formation in
the Cupiagua field. The
coastal-plain facies tract is
the Guyana field to the east. The reservoirs display located mostly in the lower half of the Mirador and
sheetlike packages of shoreface bodies; laterally and is composed of channel, crevasse splay, and swamp
vertically amalgamated channel fills and overbank and flood-plain facies successions. The bay facies
deposits and bay-head delta and bay-fill bodies. tract occurs in the upper half of the Mirador For-
The Mirador Formation is the main reservoir zone mation and is composed of bay-fill, bay-head delta,
and contains 51% of the hydrocarbons initially in and channel facies successions (Figure 8). Each facies
place. Barco and Guadalupe formations have 28 and succession consists of a spectrum of facies that are
21% hydrocarbons initially in place, respectively. arranged in regular vertical and lateral successions.
Cupiagua field has moderate core coverage of the These regular facies successions reflect lateral facies
reservoirs, with six cores taken from Mirador, four transitions along the depositional profile and their
from Barco, and four from Guadalupe Formation, for vertical superposition through progradation and
a total of about 1100 m (3800 ft). aggradation.
440 / Ramon and Fajardo

FIGURE 6. Time slice at 2236 ms


of the Cupiagua 3-D volume
showing the attitude of the
Carbonera strata. The axis of
the Nunchia syncline runs along
the left. Most strata over the
time slice belong to the east
flank of the syncline. Leon (yel-
low) and Carbonera (blue, red)
markers are shown. Note the
bending of the syncline and the
displacement of strata along
the Golconda lineament.

ture are followed by initials


of the texture of the rock.
Trough cross-stratification is
abbreviated as tx, ripples as
rp, sandstone as Ss, mud-
stone as Md, and granule
sandstone as gSs. For exam-
ple, a facies of trough cross-
stratified granule sandstone
is txgSs, sandstone with hor-
izontal burrows is hbSs, and
burrowed and irregular lam-
This section presents the results of the facies inated mudstone is blMd. Figure 9 shows conven-
analysis and descriptions and interpretations of the tions of sedimentary structures, bed contacts, and
facies, facies successions, and facies tracts recognized other symbols that are used in figures and core de-
in this study. Interpretations are discussed in terms of scription logs.
their inferred hydrodynamic and environmental pa- Core logs for the Cupiagua A-1, C-3, and H-11
rameters in a stratigraphic context. cored wells are presented in Figures 10 –12. These
Facies abbreviations are constructed from the logs include a graphic representation of the core de-
dominant sedimentary structure and the texture of scription, facies interpretation, inferred stratigraphic
the rock. Initials of the dominant sedimentary struc- cycles, gamma-ray response, and important remarks.

FIGURE 7. Strike line (cross-


line 1510) along the crest of the
Cupiagua structure. North is
to the right. The main feature in
the center of the figure is the
Golconda wrench (magenta
fault). Several vertical faults oc-
cur associated to this wrench.
North of this feature, the Cupi-
agua structure and the Nun-
chia syncline bend toward the
northeast. Mirador and intra-
Carbonera horizons are shown.
Note the change in thickness
between the C5 and Mirador
markers caused by the Yopal
thrust fault. This fault separates
the Cupiagua structure from
the overriding Nunchia syncline.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 441

FIGURE 8. Gamma-ray (GR) log and


interpreted stratigraphic cycles and fa-
cies tracts in the Mirador Formation of
the Cupiagua A-1 well. The coastal-plain
facies tract composes the lower half of
the Mirador and the bay facies tract
constitutes the upper half of the Mira-
dor Formation. Intermediate-term and
long-term stratigraphic cycles are shown.

Coastal-plain Facies Tract


The coastal-plain facies tract oc-
curs in the lower half of the Mirador
Formation. Three facies successions
were recognized: channel, crevasse
splay, and swamp and flood-plain
facies successions. The channel facies
succession occurs at the base of the
intermediate-term cycles and is com-
posed of medium- to coarse-grained
sandstones and sandstones with
floating granule-size clasts. The cre-
vasse splay facies succession overlies
the interval of channel sandstones
and is composed mainly of very fine-
to fine-grained sandstones. The
swamp and flood-plain facies succes-
sion occurs associated with the cre-
vasse splay complex and consists of
massive, laminated, and burrowed
mudstones intercalated with flood-
plain deposits and soils. The change
from channel to crevasse splay to
swamp facies succession defines an
overall increase in accommodation
space. Flood-plain deposits represent
rock deposited during a period of de-
creased accommodation.

CHANNEL FACIES SUCCESSION

The channel facies succession


contains the main reservoir rock in
the coastal-plain facies tract. Eight
facies were identified in the channel
facies succession; however, a single
channel does not necessarily exhib-
it this complete spectrum of facies.
The facies of this facies succession
are pebbly conglomerate (pbCg), mas-
sive granule sandstone (mSs), granule
442 / Ramon and Fajardo

FIGURE 9. Conventions of sed-


imentary structures used in
the core descriptions (used in
the following figures).

Pebbly Conglomerate
(pbCg)
Facies pbCg appears very
sparsely at the base of the ag-
gradational channel succes-
sion in relatively thick beds
(as much as 1.2 m [4 ft]) or in
thin lag deposits (as much as
7.6 cm [3 in.]). It is overlain
by facies mSs, uppSs, txgSs, or
txSs. The lower contact of
this facies is a scour surface,
and the upper contact is tran-
sitional or sharp with the
overlying facies. Bed thick-
ness normally is between
5 cm (2 in.) and 0.6 m (2 ft);
however, beds as much as
1.2 m (4 ft) occur. Coarse
clasts are well rounded and
generally pebble size, but
some are granule size. Quartz
clasts are the most common
(more than 95%), and the re-
mainder are lithic clasts. Sedi-
mentary structures are un-
common. Figure 14A shows
a photograph of this facies.
In a stratigraphic context,
sandstone with upper plane bed stratification (uppSs), this facies is common in relatively low accommoda-
trough cross-stratified granule sandstone (txgSs), trough tion conditions, where it constitutes a minor propor-
cross-stratified sandstone (txSs), burrowed sandstone tion of the aggradational channel facies and occurs
with relict trough cross-stratification (btxSs), and in the thickest beds. In contrast, in high accommo-
ripple-laminated sandstone (rpSs). dation settings, beds of this facies are uncommon.
The facies present in a single, aggradational chan- Two hydrodynamic interpretations are possible.
nel succession are strongly dependent on temporal When this facies contains either normal or inverse
and spatial changes in accommodation space and grading, it is interpreted as deposited by a hypercon-
base level. In general, under overall increasing ac- centrated flow. When grading is absent, it is inter-
commodation conditions, there is a transition from preted as rapidly deposited by a turbulent current in
facies pbCg, mSs, and/or uppSs into txgSs and/or txSs the upper flow regime.
and then into btxSs, cvSs, and/or rpSs.
A facies substitution diagram for the channel facies Massive Granule Sandstone (mSs)
succession is presented in Figure 13. This diagram This facies occupies the basal part of an aggra-
condenses the natural succession of facies during an dational channel succession. This facies is overlain
overall increase in accommodation space for the by facies uppSs, txgSs, or txSs. The lower contact, in
channel facies succession cored in the coastal-plain some cases, is a scour surface; in other cases, the con-
facies tract of the Mirador Formation. tact is sharp or transitional with facies pbCg. The
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 443

FIGURE 10. Core description showing interpreted facies tracts and cycles of the Cupiagua A-1 well. The Mirador
Formation is divided into six intermediate-term stratigraphic cycles and two long-term cycles. (See Figure 9 for legend.)
444 / Ramon and Fajardo

FIGURE 11. Core description showing interpreted facies tracts and cycles of the Cupiagua H-11 well. (See Figure 9 for legend.)
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 445

FIGURE 12. Core description showing interpreted facies tracts and cycles of the Cupiagua C-3 well. (See Figure 9 for legend.)
446 / Ramon and Fajardo

FIGURE 13. (A) Channel facies


succession, example from Cupia-
gua H-11 well. Facies that com-
pose these aggradational units
define an overall increase in ac-
commodation space. (B) Facies
substitution diagram of channel
facies succession. The horizontal
axis represents the probability of
the substitution of facies. The
area occupied by each facies is
directly proportional to the fre-
quency of the facies in the chan-
nel facies succession. (See Figure 9
for legend.)

The quartz arenite is upper


medium to coarse grained with
a few scattered granules. Sedi-
mentary structures are absent,
but normal or inverse coarse-tail
grading uncommonly occurs.
In a stratigraphic context,
this facies is present only in
low accommodation condi-
tions. Two hydrodynamic in-
terpretations are considered.
When normal or inverse coarse-
tail grading is present, this fa-
cies is interpreted as a hypercon-
centrated flow deposit. When
grading does not occur, it is in-
terpreted as rapidly deposited
by a turbulent current in the up-
per flow regime.

Granule Sandstone with


Upper Plane Bed
Stratification (uppSs)
This facies occurs in the
lower part of the aggradational
channel successions and is
overlain by facies txgSs or txSs.
The lower boundary is either
sharp or transitional. The up-
per boundary is always sharp.
Bed thickness commonly ranges
between 0.15 and 0.6 m (0.5
and 2 ft). The quartz arenite is
coarse to very coarse grained,
with scattered quartz and un-
upper contact normally is sharp. Bed thickness com- common lithic granules. The planar stratification is
monly ranges between 7.6 and 20 cm (3 and 8 in.), defined by alignment of pebble-size quartz clasts in
but, in exceptional cases, are as much as 0.6 m (2 ft). horizontal or low-angle (less than 58) laminae.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 447

In a stratigraphic context, this facies occurs in rela- Burrowed Sandstone with Relict
tively low accommodation conditions. Facies uppSs Trough Cross-stratification (btxSs)
was deposited by a turbulent current in the upper flow This facies occupies the upper part of aggrada-
regime. tional channel successions. This facies overlies facies
txSs and substitutes with facies rpSs. Bed thickness is
Trough Cross-stratified Granule Sandstone (txgSs) sometimes difficult to observe, but, in general, is less
This facies constitutes as much as 30% of the channel than 0.3 m (1 ft). The quartz arenite is medium to coarse
facies succession cored in the Mirador Formation. It is grained. The most characteristic feature is the burrow-
commonly overlain by facies txSs and, in very few cases, ing. More than 70% of the original trough cross-
by facies btxSs. The quartz arenite is coarse to very stratification is destroyed, and relict foresets with vari-
coarse grained with scattered pebbles. Sometimes, the able angles of inclination are visible through the
granule-size grains are arranged along the foresets. burrowing overprint. In some cases, the rock is com-
Thickness of trough cross-stratification sets is variable, pletely burrowed, and some irregular laminations occur.
commonly from 0.3 to 1.2 m (1 to 4 ft). Sets are com- In a stratigraphic context, this facies occurs in high
monly of constant grain size, but some sets exhibit a accommodation conditions. When the burrow inten-
fining-upward trend. The angle of the foreset lami- sity is constant between two scour surfaces, this facies
nae range between 20 and 308, and the foreset lami- was deposited under low sedimentation rates. When
nae thickness ranges between 5 and 40 mm (0.2 and burrow intensity decreases from top to bottom in
1.5 in.). Pebbles are aligned parallel to the foreset sandstones between two scour surfaces, it is inter-
laminae. preted as fluctuations in rates of accumulation; periods
This facies occurs in higher accommodation con- of bed migration were interrupted by a time of non-
ditions than the previous facies. The bed set thickness deposition when organisms reworked the sediment.
changes with changes in accommodation space; as
accommodation increases, bed set thickness increases. Ripple-laminated Sandstone (rpSs)
It was deposited by a turbulent current in the lower This facies is uncommon in the channel facies suc-
flow regime. Facies txgSs is the stratigraphic record of cession cored in the Mirador Formation. Facies rpSs,
migrating 3-D dunes on the channel floor. when present, appears on top of the aggradational
channel successions. Bed thickness is between 5 cm
Trough Cross-stratified Sandstone (txSs) (2 in.) and 0.3 m (1 ft). The quartz arenite is lower to
Facies txSs is the most common (as much as 60%) upper medium grained. Occasionally, minor burrow-
facies of the channel facies succession. It is overlain ing that partially destroyed the ripple laminae is typical.
by facies btxSs, cvSs, and rpSs. The quartz arenite is In a stratigraphic context, this facies occurs in high
medium to coarse grained. Fining-upward trends accommodation settings. It is interpreted as a waning
occur in some vertical successions of sets. Thickness flow cap of the hydrodynamic regime responsible
of trough cross-stratification sets is variable, common- for the deposition of the channel succession.
ly between 0.15 and 0.6 m (0.5 and 2 ft); however,
some sets are strongly amalgamated and range be- CREVASSE SPLAY FACIES SUCCESSION
tween 5 and 12.7 cm (2 and 5 in.) thick, whereas other
sets are as much as 1.2 m (4 ft) thick. Foreset lami- This facies succession was preserved during rela-
nae are inclined between 15 and 258 and are between tively higher accommodation conditions than the
3 and 30 mm (0.118 and 1.18 in.) thick. Foreset lami- channel facies succession. Thin muddy beds of the
nae commonly are locally broken by burrowing. swamp and flood-plain facies succession are com-
Figure 14B is a photograph of this facies. monly interlayered with this facies succession. Five
In a stratigraphic context, this facies was deposited facies are identified: sandstone with mudstone rip-up
in higher or the same accommodation conditions as clasts (rupSs), small-scale trough cross-stratified sand-
facies txgSs. The bed set thickness changes as a func- stone (stxSs), ripple-laminated sandstone (rpSs),
tion of the accommodation space; as accommoda- convolute sandstone (cvSs), and burrowed and irreg-
tion space increases, the bed set thickness increases. ular laminated sandstone (blSs). The crevasse splay fa-
Facies txSs was deposited by turbulent currents in cies succession generally does not contain good res-
the lower flow regime. The trough cross-stratified ervoir rock. For similar porosities, the crevasse splay
sandstone is the stratigraphic record of migrating 3-D sandstones have much lower permeability distribu-
dunes on the channel floor. tions as compared to channel sandstones.
448 / Ramon and Fajardo

FIGURE 14. Photos of the different facies present in the coastal-plain facies tract in the Lower Mirador in the Cupiagua
field. (A) Pebbly conglomerate. Clast are well rounded and composed of white quartz and some cherts; (B) trough
cross-stratified sandstone; (C) sandstone with rip-up clasts; (D) small-scale trough cross-stratified sandstone; (E) ripple-
laminated sandstone; (F) convolute sandstone (3805 m [12,484 ft] measured depth); (G) burrowed, laminated
sandstone; (H) laminated mudstone with starved ripples; (I) mottled mudstone; (J) rooted and irregular laminated
mudstone; (K) brecciated mudstone. All cores are 10cm (4 in.) wide.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 449

FIGURE 14. (cont.).


450 / Ramon and Fajardo

Crevasse splay intervals normally occur between


channel facies successions and swamp facies succes-
sions. This succession of facies tracts records increas-
ing wetness and increasing preservation through
time, beginning with a subaerially exposed surface of
unconformity at the channel base.
Facies successions in crevasse splay complexes
constitute a series of commonly asymmetrical, short-
term, base-level fall hemicycles that contain progres-
sively less crevasse-channel facies and progressively
more distal crevasse splay facies (Figure 15). This
means that initial crevasse splay progradations onto
the flood-plain or ephemeral flood-plain swamps were
replaced by crevasse splay progradations into wetter
flood plains and permanent swamps. Thus, individual
short-term stratigraphic cycles represent episodes of
progradation, but each episode brought progressively
more distal facies into the short-term facies succes-
sion. The short-term stratigraphic cycles are stacked
to form an intermediate-term base-level rise cycle re-
flecting increasing accommodation through time.

Sandstone with Rip-up Clasts (rupSs)


Facies rupSs occurs uncommonly and normally
overlies scour surfaces; in some cases, it is interlay-
ered with facies blSs. The upper contact is either tran-
sitional or sharp. Bed thickness commonly ranges
between 12 and 20 cm (5 and 8 in.); however, beds as
much as 0.6 m (2 ft) occur. This quartz arenite is fine
to lower medium grained. Mudstone rip-up clasts are
as much as 3 cm (1.2 in.) long in dimension and oc-
casionally so abundant that the rock is almost a rip-
up clast conglomerate (Figure 14C). Variable degrees
of burrowing and irregular lamination occur in fa-
cies rupSs.
Where this facies overlies a scour surface, it is in-
terpreted as a crevasse channel lag deposit. The mud
clasts were either ripped up because of channel scour-
ing and incorporated in the flow or were derived from
slumps into the channel and incorporated in the flow.
When facies rupSs does not overlie a scour surface, it
is interpreted as related to an upstream scour event.
Facies rupSs occurs in the proximal parts of a cre-
vasse splay. In a stratigraphic context, it represents
relatively lower accommodation conditions in a cre-
vasse splay succession.

Small-scale Trough Cross-stratified


Sandstone (stxSs)
This facies constitutes as much as 30% of the cre-
vasse splay intervals cored in the Mirador Forma-
FIGURE 14. (cont.). tion. It occurs in proximal portions in a crevasse splay
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 451

FIGURE 15. Crevasse-splay facies suc-


cession in cycle I of the Cupiagua A-1
well. Individual short-term strati-
graphic cycles represent episodes of
progradation, but each episode brings
progressively more distal facies into
the short-term facies succession. The
short-term stratigraphic cycles are
stacked to form an intermediate-term
base-level-rise cycle reflecting in-
creasing accommodation through
time. (See Figure 9 for legend.)

are thicker than 0.3 m (1 ft). The


quartz arenite is lower to middle
fine grained and well sorted. Small-
scale trough cross-bedding is the
most common sedimentary struc-
ture; however, different degrees of
burrowing partially destroy the fore-
set laminae. Foreset laminae are
inclined 10–208 and are 1–2 mm
(0.04–0.08 in.) thick. The type of
burrowing was not identified, but
there is a low diversity of ichnofos-
sils. Figure 14D shows a photograph
of this facies.
The small-scale trough cross-
stratification is the record of mi-
grating 3-D dunes in both channel-
ized and unconfined unidirectional
flow in the lower flow regime. Bur-
rowing is evidence of fluctuations in
flow strength that allowed rework-
ing of sediments by organisms.
Facies stxSs is interpreted as part of
a crevasse channel deposit where
there is a succession from a scour
surface into facies stxSs, then into
rpSS, and finally, into facies blSs.
When the succession is from facies
blSs into rpSs and into stxSs, this
facies is interpreted as a crevasse
splay prograding into a swamp and
flood plain.
Facies stxSs occupies proximal
portions in a crevasse splay com-
complex. The lower contact is either a scour surface plex. This facies commonly appears in the lower
or a transitional surface from underlying ripple- parts of the crevasse splay intervals cored in the
laminated sandstones. The upper contact is sharp Mirador Formation. In a stratigraphic context, facies
or transitional. Bed thickness commonly ranges be- stxSs represents relatively lower accommodation
tween 5 and 15 cm (2 and 6 in.); uncommonly, beds conditions in a crevasse splay deposit.
452 / Ramon and Fajardo

Ripple-laminated Sandstone (rpSs) exhibit carbonaceous or muddy irregular lamina-


Facies rpSs constitutes as much as 40% of the cre- tions and relict foreset laminae of short-term trough
vasse splay facies succession. Upper and lower contacts cross-stratification (Figure 14G).
of the ripple-laminated sandstone are either sharp or Facies blSs was deposited in an oxygenated envi-
transitional. Bed thickness generally ranges between ronment where benthic organisms reworked the sedi-
2.5 and 15 cm (1 and 6 in.); however, beds as much as ments that were deposited. It is found in distal parts of
0.45 m (1.5 ft) thick are found. The quartz arenite is prograding crevasse splays and on top of crevasse chan-
fine to very fine grained and well sorted (Figure 14E). nels. This facies was deposited in relatively higher
Differential degrees of burrowing partially destroy the accommodation conditions than facies stxSs or rpSs.
ripple laminae.
This facies represents a migration of ripples in the SWAMP AND FLOOD-PLAIN
lower flow regime in either channelized or uncon- FACIES SUCCESSIONS
fined flow. This facies appears in two different suc-
cessions in a crevasse complex. It occurs in crevasse The swamp and flood-plain facies successions are
channel deposits, where it is interpreted as a waning dominated by mudstones, and they generally occur
flow cap of the crevasse channel succession. It also adjacent to the crevasse splay facies succession. They
appears in the lower half of shallowing-upward suc- are more common toward the top of the intermediate
cessions of prograding crevasse splays and is inter- cycles of the Lower Mirador (Figure 10). Most mud-
preted as deposition in unconfined flow. stones correspond to wet flood plain and swamps.
Facies rpSs is better developed in more distal parts Locally, these are replaced transitionally by increas-
of a crevasse splay facies succession than facies stxSs ing thickness of soils. An unconformity is interpreted
and rupSs. In a stratigraphic context, those intervals toward the top of the Lower Mirador in the Cusiana
with major proportion of facies rpSs are interpreted field. This unconformity is underlain by soils, caliche,
as deposited during higher accommodation condi- and brecciated mudstones and is overlain by strata,
tions than intervals dominated by facies stxSs. which accumulated in wet flood-plain, marsh, and
swamp to brackish bay environments (Fajardo, 1995).
Convolute Sandstone (cvSs) Three facies were differentiated in the swamp fa-
This facies is uncommon in the crevasse splay fa- cies succession: massive mudstone (mMd), laminated
cies succession. Bed thickness ranges between 5 and mudstone with starved ripples (lsrMd), and burrowed
25 cm (2 and 10 in.) (Figure 14F). The quartz arenite and irregular laminated mudstone (blMd). Three fa-
is very fine grained and well sorted. Relicts of ripple cies where identified in the flood-plain facies succes-
laminae are common. sion: mottled mudstone (mtMd), rooted and irregular
This facies is interpreted as rapidly deposited sedi- laminated mudstone (rtMd), and brecciated mud-
ments that collapsed because of water escape pro- stone (brMd).
cesses. This facies is associated with facies rpSs or blSs. A facies substitution diagram for swamp and flood-
It was deposited in similar accommodation condi- plain facies succession is presented in Figure 16. This
tions as facies rpSs. diagram represents the succession of facies under
decreasing accommodation conditions from swamp
Burrowed, Laminated Sandstone (blSs) to flood-plain facies succession. The left part of the
This facies constitutes as much as 10% of the cre- diagram represents drier conditions. The right part of
vasse splay facies succession cored in the Mirador the diagram represents wetter conditions or even
Formation. It occupies the distal part of a crevasse brackish conditions.
splay complex. Upper and lower contacts are either
transitional or sharp. Bed thickness is most common- Massive Mudstone (mMd)
ly between 5 cm (2 in.) and 0.6 m (2 ft). The quartz This facies occurs between crevasse splay deposits
arenite is very fine to fine grained and well sorted. or with other swamp facies. Lower and upper con-
Clay matrix is present in some intervals. The most tacts are commonly sharp, where it is interbedded
distinctive feature of this facies is the irregular ar- with crevasse splay facies successions, but where it is
rangement of the lamination caused by a high degree interbedded with other swamp facies associations,
of burrowing. Sometimes, the rock is completely or the contacts are generally transitional. Bed thickness
almost completely homogenized by organism rework- is variable. Where this facies is interbedded with cre-
ing. Some intervals with lower degree of burrowing vasse splay sandstones, the bed thickness ranges from
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 453

2.5 to 12 cm (1 to 5 in.). The color is


commonly dark gray and medium
gray. Some mudstones have high or-
ganic matter content.
Facies mMd was deposited from
suspension in open-swamp and open-
lake conditions. This facies is inter-
preted as the deepest swamp facies.

Laminated Mudstone with


Starved Ripples (lsrMd)
This facies consist of laminated
gray mudstones intercalated with
variable amounts of silt and few very
fine sandstone ripples. Upper and
lower contacts are either sharp or tran-
sitional. When it is interbedded with
crevasse splay sandstones, contacts are
sharp, and beds range between 2.5 and
15 cm (1 and 6 in.) thick. Beds of this
facies in open-swamp intervals range
between 0.3 and 1 m (1 and 3 ft) thick;
in exceptional cases, beds as much as
1.8 m (6 ft) thick occur. Texturally,
this facies varies from silty mud to
muddy silt. It is commonly dark to
middle gray and occasionally green-
ish gray or black. Plane parallel hori-
zontal lamination, starved ripples,
and lenticular and wavy laminations
are the characteristic sedimentary
structures (Figure 14H).
This facies was deposited from
suspension in open-swamp environ-
ments. It is interpreted as having
been deposited at the same depth or
even a shallower depth than facies
mMd. The starved ripples and lentic-
ular and wavy lamination are evi-
dence of weak currents.

Burrowed and Irregular


Laminated Mudstone (blMd)
Lower and upper contacts are
either transitional or sharp. Bed thick-
ness ranges from 0.3 to 1 m (1 to 3 ft).
It is generally composed of muddy
siltstone but may be sandy siltstone.
It exhibits different colors, including
FIGURE 16. Swamp and flood-plain facies successions in the Cupiagua A-1
well. The succession from swamp to flood plain defines an overall increase dark to light gray, greenish gray, and
in accommodation space in intermediate-scale cycle III. Cycle IV is composed beige. Burrowing is the most distinc-
of wetter mudstones of marsh and brackish ponds. (See Figure 9 for legend.) tive feature of this facies. Irregular
454 / Ramon and Fajardo

laminations are frequently present. On top of the


lower Mirador, facies blMd changes transitionally to
mottled and rooted mudstones (mtMd or rtMd).
This facies was deposited from suspension in a
swamp or in ephemeral flood-plain ponds. In a strati-
graphic context, where this facies is interbedded with
facies mtMd or rtMd, it is interpreted as the result
of an overall decrease in accommodation space. In
contrast, where it is interbedded with facies mMd
and lsrMd, it is interpreted as having been deposited
during an overall increase in accommodation space.

Mottled Mudstone (mtMd)


This facies’ most distinctive feature is its varicol-
ored aspect. Upper and lower contacts are transi-
tional. Bed thickness ranges between 0.6 and 2.7 m
FIGURE 17. Facies succession and substitution diagram
(2 and 9 ft). The mudstones of some intervals have a
of the bay-fill and the channel and bay-head facies tract.
dark-gray background with irregular reddish-brown Bay-fill and bay-head cycles tend to have a base-level-fall
patches; other mudstones are light gray or beige asymmetry with a thin base-level-rise cap. (See Figure 9
with reddish and yellowish patches, and others are for legend and text for abbreviations.)
green or beige with blue and purple patches. A com-
mon feature is the vertical distribution and orienta- Bed thickness ranges between 1 and 1.5 m (3 and 5 ft).
tion of patches and irregular veins (Figure 14I). This facies appears brecciated, and it is composed
This facies is interpreted as deposited from sus- of angular, irregular masses of brownish color with
pension in a flood-plain environment where there chaotic distribution (Figure 14K). In some intervals,
are alternating dry and wet conditions. In a strati- ferruginous pseudonodules of less than 2 mm (0.08 in.)
graphic context, this facies represents a time of de- are present.
creasing accommodation. This facies results from modification of previously
deposited mudstones by soil processes. The facies
Rooted and Irregular Laminated has strong evidence of subaerial exposure.
Mudstone (rtMd)
This facies is beige and light gray; it contains Bay Facies Tract
black to brown root traces composed of carbona- The bay facies tract occurs in the upper Mirador
ceous material and is interbedded with facies mtMd. Formation. In this facies tract are channel, bay-head
The basal and upper contacts are commonly transi- delta, and bay-fill facies successions (Figures 10 –12).
tional. Bed thickness ranges between 0.3 and 2.1 m A transition from channel to bay-head delta to bay-fill
(1 and 7 ft). Burrows are another feature present in facies successions represents a continuous increase in
this facies (Figure 14J). accommodation space. A transition from bay-fill to
This facies represents the soil alteration of a pre- bay-head delta to channel facies succession defines
vious facies deposited from suspension. Two inter- a continuous decrease in accommodation space.
pretations of this facies are possible. Where it is in- A facies substitution diagram of the bay facies tract
terbedded with facies mtMd, it is interpreted as is presented in Figure 17. This diagram was construct-
deposited in a flood-plain setting where drier condi- ed from the bay facies tract intervals cored in three
tions predominated. Where it contains organic lami- wells in the Mirador Formation in the Cupiagua field.
nations and it is interbedded with facies mMd, it is
interpreted as deposited in a wetter environment like CHANNEL AND BAY-HEAD DELTA
an ephemeral pond. FACIES SUCCESSIONS

Brecciated Mudstone (brMd) Channel and bay-head delta facies successions con-
This facies constitutes a small percentage of the tain the main reservoir rock in the bay facies tract.
swamp and flood-plain facies successions. The lower These successions are dominant in the lower half of
contact is transitional, and the upper contact is sharp. the upper Mirador in the Cupiagua field area. Three
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 455

facies are recognized in these successions: trough Ripple-laminated Sandstone (rpSs)


cross-stratified sandstone (txSs); burrowed sandstone This facies is uncommon (2%) in the bay facies tract.
with relict trough cross-stratification (btxSs); and It sometimes occurs on top of channel facies suc-
ripple-laminated sandstone (rpSs). Figures 10–12 show cessions and uncommonly within bay-fill facies suc-
examples of these facies successions. cessions. In channel successions, the thickness is as
much as 0.6 m (2 ft); in bay-fill successions, it is less
Trough Cross-stratified Sandstone (txSs) than 20 cm (8 in.). The quartz arenite is fine grained
Facies txSs is the most common facies of the bay and commonly contains clay matrix. It exhibits sev-
facies tract. It occurs both in channel and bay-head eral degrees of burrowing, but relict ripples can al-
delta facies successions. Upper and lower contacts of ways be easily identified.
trough cross-stratification sets are commonly sharp; This facies was deposited in the lower flow re-
however, when burrows are present, the contacts are gime, and it represents the migration of 3-D rip-
indistinct. The quartz arenite normally is medium ples. The intensive burrowing suggests low rates of
grained, but coarse-grained intervals occur. Fining- deposition.
upward or coarsening-upward trends are recognized
in a vertical succession of sets. Thickness of trough BAY-FILL FACIES SUCCESSION
cross-stratification sets is variable; it ranges from few
inches to as much as 1.2 m (4 ft). Foreset laminae are This facies succession is composed of burrowed
inclined between 15 and 258 and are between 2 and sandstones and mudstones. It is more common in
8 mm (0.08 and 0.3 in.) thick. Minor burrowing oc- the upper half of the upper Mirador. Four facies are
curs in this facies. With the exception of uncommon defined in this facies succession. From deeper to shal-
burrows, this facies is virtually identical to facies lower, those facies are laminated mudstone (lMd),
txSs of the coastal-plain facies tract (see Figure 14B). burrowed and irregular laminated mudstone (blMd),
Figure 18A is a photograph of this facies. burrowed and irregular laminated sandstone (blSs),
Facies txSs was deposited by turbulent currents in and vertical burrowed sandstone (vbSs) (see Figure 17).
the lower flow regime. The trough cross-stratified sand- A transition from lMd to blMd to blSs and, later, into
stone is the stratigraphic record of migrating 3-D dunes vbSs represents a continuous decrease in accommo-
on the channel floor or in bay-head deltas. dation space (Figure 17). In some cases, facies vbSs is
overlain by facies hbSs; this upward transition rep-
Bioturbated Sandstone with Relict Trough resents an increase in accommodation space.
Cross-stratification (btxSs)
This facies is the second most common facies of Laminated Mudstone (lMd)
the bay facies tract. It occurs in the upper part of This facies is uncommon in the bay facies tract. It
aggradational channel successions and in the lower occurs at the base of shallowing-upward bay-fill suc-
part of progradational bay-head delta successions. cessions. Upper and lower contacts are normally sharp.
Trough cross-stratification set thickness is some- Bed thickness is less than 1 m (3 ft). The mudstone is
times difficult to observe but, in general, is less than dark gray to black and exhibits discontinuous silt lami-
0.45 m (1.5 ft). The quartz arenite is medium to coarse nae or lenses (Figure 18C).
grained. The most characteristic feature is the bur- This facies is interpreted as having been deposited
rowing. More than 60% of the original trough cross- from suspension. The lamination is evidence of weak
stratification is destroyed, and relicts of foresets with currents. The absence of burrowing indicates that
variable angle of inclination are visible throughout dysaerobic or anaerobic conditions occurred at the
the burrowing overprint (Figure 18B). Macaronichnus, sediment-water interface.
Ophiomorpha, Gyrolithes, Arenicolites, and crab bur-
rows are recognized. Burrowed and Irregularly
Where the burrow intensity is constant between Laminated Mudstone (blMd)
two scour surfaces, this facies was deposited under This facies constitutes less than 4% of the bay facies
low sedimentation rates. Where burrow intensity tract. It occupies the lower part of the shallowing-
decreases from top to bottom, it is interpreted as upward bay-fill successions. Upper and lower con-
fluctuations in rates of accumulation; periods of bed tacts are either sharp or transitional. Bed thickness is
migration are interrupted by times of nondeposition variable, ranging from few inches to as much as 1.2 m
when organisms reworked the sediment. (4 ft). This facies is composed of dark-gray or black
456 / Ramon and Fajardo

FIGURE 18. Photos of the different facies present in the bay facies tract in the upper Mirador in the Cupiagua field.
(A) Trough cross-stratified sandstone; (B) bioturbated sandstone with relict trough cross-stratification; (C) laminated
mudstone; (D) bioturbated and irregularly laminated mudstone (3797 m [12,457 ft] measured depth); (E) bioturbated
and irregularly laminated sandstone; (F) vertically burrowed sandstone. All cores are 10 cm (4 in.) wide.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 457

silty mudstone. The more distinctive characteristic is


the intensive burrowing. Burrows are mainly hori-
zontal. In some intervals, Thalassinoides burrows are
recognized (Figure 18D). Irregular and discontinuous
laminations normally occur.
This facies was deposited from suspension in open-
bay conditions. The intensive burrowing implies
optimal substrate conditions for organisms to live.

Bioturbated and Irregularly


Laminated Sandstone (blSs)
This facies constitutes 30% of the bay-fill facies
tract. Bed thickness normally is between 15 cm (6 in.)
and 0.3 m (1 ft); beds as much as 1 m (3 ft) thick occur.
It is composed of fine and very fine sandstone with
variables amounts of clay matrix. It has carbonaceous
or muddy irregular laminations and variable degrees
of burrowing (Figure 18E). The major difference with
facies vbSs is that single ichnofossils are not easily
distinguished. Mudstone rip-up clasts occur in some
intervals. This facies was deposited in open-bay en-
vironments at slow rates, which allowed intense re-
working by organisms.

Vertically Burrowed Sandstone (vbSs)


This facies is the most common of the bay facies
tract. It occupies the upper part of shallowing-upward
bay-fill successions. Upper and lower contacts are
normally transitional with overlying and underlying
facies. Bed thickness is variable, ranging from a few
inches to 2 m (6 ft). The quartz arenite is fine and
sparsely very fine or medium grained. The distinctive
feature is the dominance of vertical burrows, such as
Ophiomorpha, Skolithos sp., and Arenicolithes (Figure 18F).
Horizontal burrows are uncommon and include
Teichichnus, Planolites, and Paleophycus. This facies
lacks ripple lamination or cross-stratification, but ex-
hibits carbonaceous or muddy horizontal wispy lami-
nations. This facies was deposited at higher rates than
facies blMd and blss in open-bay settings.
Facies vbSs was deposited in open-bay environ-
ments at slow rates, which allowed reworking by
organisms.

FOUR-DIMENSIONAL STRATIGRAPHIC
ARCHITECTURE OF THE MIRADOR FORMATION

Three scales of stratigraphic cycles are recognized


in the Mirador Formation. Short-term (high-frequency)
cycles correspond to progradational and aggradational
units. Six intermediate-term cycles were identified by
FIGURE 18. (cont.). the stacking patterns of their component short-term
458 / Ramon and Fajardo

FIGURE 19. Correlation of the cored wells in the Cupiagua field showing the long-term stratigraphic cycles defined in
the Mirador Formation. Depth in feet.

cycles and by the general trend of facies successions, by deposition on an almost horizontal surface; exam-
indicating increasing or decreasing accommodation- ples of aggradational units are channel, lake, and
to-sediment supply (A/S) ratios. Two long-term cy- flood-plain deposits. Progradation occurs by deposi-
cles were defined from the stacking pattern of the tion on an inclined surface; examples of prograda-
intermediate-term cycles and by the general trend of tional units are crevasse splay, bay-head delta, and
facies successions (Figure 19). bay-fill deposits. A single short-term cycle generally
Stratigraphic cycles, regardless of scale, record a is composed of both aggradational and prograda-
complete base-level cycle (Barrell, 1917; Wheeler, tional components because of changes in inclination
1964). During a base-level cycle, accommodation-to- of depositional surfaces along the geomorphic pro-
sediment supply ratio increases (as base-level rises) to file. The definition of short-term cycles is based on fa-
a maximum limit and then decreases (as base-level cies successions, occurrence of surfaces of stratigraphic
falls) to a minimum limit along the entire geomorphic discontinuity, and measurements of sedimentologic
profile connecting linked depositional environments. attributes like bed set thickness or burrow orientation,
diversity, and density.
Short-term Stratigraphic Cycles Short-term stratigraphic cycles in channel facies
Short-term stratigraphic cycles are progradational successions commonly exhibit base-level rise asym-
and aggradational units that conform to Walther’s metry, and less commonly with a thin base-level fall
law. These short-term cycles are the building blocks hemicycle. Asymmetry means that the time of in-
of the stratigraphic framework. Aggradation occurs creasing A/S is preserved as rock, whereas the time of
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 459

decreasing A/S is mainly represented by surfaces of in the interval. From deep to shallow, the swamp
stratigraphic discontinuity. In the Mirador Forma- facies succession is from massive mudstone (mMd),
tion, typical thickness of short-term cycles composed to laminated mudstone with starved ripples (lsrMd),
of channel facies successions is 3– 6 m (10 –20 ft), and to burrowed and irregularly laminated mudstone
the range is 2.1 – 9 m (7 – 30 ft). (blMd), to gray mudstones with roots to mottled
When channel successions exhibit enough facies mudstones. This transition from swamp (wet) to dry
diversity, the recognition of base-level rise deposits flood plain is interpreted as a rise in base level.
is based on a vertical change of facies representing If flood-plain facies successions occur without in-
reduction in flow concentration and/or flow strength. tervening swamp and lake beds, short-term cycles
A complete succession would include facies depos- are defined by the degree of soil development. A suc-
ited as hyperconcentrated flows, followed by facies cession from mtMd to brecciated mudstone (brMd)
deposited in the upper flow regime, followed by fa- facies indicates a base-level-rise hemicycle. In inter-
cies deposited in the lower flow regime. This suc- vals composed only of mtMd facies, short-term cycles
cession could be indicated by a transition from granule are defined by changes in the degree of mottling. An
sandstone with upper plane bed stratification (uppSs) increase in mottling (increased duration of soil-forming
facies at the channel base to trough cross-stratified processes) is interpreted as a base-level rise, and a
granule sandstone (txgSs) or trough cross-stratified sand- decrease in mottling is interpreted as a base-level fall
stone (txSs) facies, followed by trough cross-stratification as more sediment is aggraded over the flood plain.
with some burrowing (btxSs), ripple-laminated sand- Short-term stratigraphic cycles in the bay facies
stone (rpSs), or convolute-laminated sandstone (cvSs) tract were identified in channel, bay-head delta, and
facies (Figure 13). The base-level-fall hemicycle is bay-fill facies successions. A single short-term cycle
commonly represented by a scour surface. However, can be composed of only one or a combination of
if it is preserved as rock, it is recognized by facies these facies successions.
successions that indicate an increase in sediment Channel and bay-head delta facies successions com-
reworking and amalgamation. A complete succession monly coexist at each geographic location in a single
would be a transition from facies deposited in the short-term cycle. Cycles are base-level rise or fall asym-
lower flow regime to facies deposited in the upper metrical, or they are symmetrical. Channels are pre-
flow regime and even to facies deposited as hyper- served during both short-term base-level rise and fall.
concentrated flows. In contrast, bay-head deltas only occur during short-
Short-term stratigraphic cycles in crevasse splay term base-level-fall time. Bay-head deltas are probably
facies successions exhibit a base-level fall or base- replaced by bay-fill facies successions during short-
level rise asymmetry or are symmetric alternating rise term base-level-rise time. Cycle thickness is variable:
and fall hemicycles (Figure 15). Thickness of these cycles composed of channel and bay-head delta suc-
short-term cycles in the Mirador Formation ranges cessions range from 3.3 to 9 m (11 to 30 ft), but cycles
from 0.9 to 5.1 m (3 to 17 ft). composed of either one or the other facies succession
Definition of cycles in crevasse splay facies suc- range from 1.2 to 3.6 m (4 to 12 ft).
cessions is based on the recognition of shallowing- A short-term base-level-rise hemicycle in channel
upward intervals developed during the progradation facies successions is indicated when a channel scour
of a crevasse splay into permanent or ephemeral surface is overlain by trough cross-stratified sand-
flood-plain lakes. Sometimes, the crevasse facies be- stones (txSs), which constitutes the major part of the
come less frequent and thinner, and swamp and flood- channel deposit and is followed by burrowed sand-
plain facies becomes thicker and more frequent. stone with relict trough cross-stratification (btxSs) and,
This transition is interpreted as a rise base-level. in some cases, is capped by rippled sandstone (rpSs).
Short-term stratigraphic cycles in lake and flood- The increase in burrowing toward the top and the
plain facies successions are more symmetrical than decrease in flow velocity reflect increasing accom-
in any other facies succession of the coastal-plain modation and gradual drowning of the channel.
facies tract. Thickness of Mirador lake and flood- Short-term base-level-fall hemicycles in bay-head
plain short-term cycles ranges from 0.6 to 4.8 m (2 to delta successions are recognized from facies succes-
16 ft). The definition of short-term stratigraphic cy- sions and changes in cross-stratification set thick-
cles in intervals composed only of swamp and flood- ness. Bay-head delta successions exhibit both tran-
plain facies successions is based on the interpreted sitional and sharp lower contacts with underlying
wetness and dryness of the different facies that occur facies. A typical facies succession is a change from
460 / Ramon and Fajardo

FIGURE 20. (A) Channel and bay-head


delta facies successions, Cupiagua
A-1 well. Trough cross-stratified sand-
stone in the lower part of bay-head delta
succession is intensively burrowed, and
the burrowing decreases upward. The
core photograph (B) shows the same
interval. Note the better oil saturation
toward the top of the core (right), indi-
cating better reservoir quality. (See
Figure 9 for legend.) Depth in feet.

burrowed sandstone (bSs) at the base


to burrowed sandstone with relict
trough cross-stratification (btxSs) and
then to trough cross-stratified sand-
stone (txSs). Burrowing intensity di-
minishes toward the top, and trough
cross-stratification set thickness in
facies txSs thins upward. These facies
successions record an increase in the
hydromechanical energy from base
to top and increased cannibalization
of migrating dunes during progra-
dation, which are interpreted as an
upward decrease in accommodation
(Figure 20).
Short-term stratigraphic cycles in
bay-fill facies successions tend to
have base-level-fall asymmetry with
a thin base-level-rise cap. Typical
cycle thickness ranges from 1.2 to
3.6 m (4 to 12 ft), but in excep-
tional cases, cycles are as much as
5.4 m (18 ft) thick. Short-term cycles
commonly are recognized from fa-
cies successions of shallowing-up-
ward bay-fill deposits. In a typical
shallowing-upward bay-fill succes-
sion, there is a transition from bur-
rowed and irregularly laminated
mudstone (bMd) at the base to bur-
rowed and irregularly laminated
sandstone (blSs), followed by vertical
burrowed sandstone (vbSs) toward
the top. This succession represents a
continuous decrease in accommo-
dation-to-sediment supply ratio
caused by progradation. The bay-
head delta facies succession some-
times overlies bay-fill successions in
a single short-term cycle (Figure 20).
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 461

Thin base-level rise caps are recognized in bay-fill Cusiana field show that the proportion of swamp and
deposits. They exhibit a succession from shallower to flood-plain facies successions increases from south-
deeper facies, from vbSs to blSs facies, and eventually, east to northwest, and the proportion of crevasse
to bMd facies. The short-term cycles have a thicker splay facies successions decreases in this direction.
base-level-fall hemicycle with thin base-level-rise caps. The lower proportion of the channel facies tracts
with respect to the crevasse and flood-plain facies
Intermediate-term Stratigraphic Cycles tracts in the Cupiagua field area suggests that depo-
Six intermediate-term stratigraphic cycles are de- sitional dip orientation was toward the northwest.
fined in the Mirador Formation in the Cupiagua field This is consistent with the trend of paleovalleys
(Figures 10 – 12, 19). Intermediate-term stratigraphic mapped by Fajardo (1995).
cycles are defined from the stacking pattern of their
component short-term cycles, from the general trend Cycle II
of facies successions, and from surfaces of strati- The lower boundary of intermediate-term cycle II
graphic discontinuity or dislocation. These cycles are is a base-level-fall surface. This cycle exhibits a base-
designated with Roman numerals (I–VI). Facies as- level-rise asymmetry. Cycle II consists of channel and
semblages of the three basal cycles (I, II, and III) be- crevasse splay facies successions. The flood-plain fa-
long to the coastal-plain facies tract. Cycles IV, V, and cies succession is only found in a couple of wells.
VI are composed of bay facies tracts. Cycle II has a transition from channel to crevasse
splay facies successions (Figure 10), which, in some
Cycle I cases, are overlain by the swamp and flood-plain fa-
Cycle I is only cored in the Cupiagua A-1 well. This cies succession. This trend of facies successions de-
cycle exhibits mostly flood-plain and crevasse facies fines an intermediate-term base-level rise. Short-term
tracts. This core shows a transition from stacked cycles in channel and crevasse splay intervals have a
flood-plain facies of the Los Cuervos Formation into general landward-stepping pattern in the base-level-
crevasse channel and splay facies with some swamp rise hemicycle.
and wet flood-plain facies successions. This trend of Channel facies successions are dominant in the
facies successions suggests a base-level fall to rise lower half of cycle II. Facies that compose these cycles
progression (Figure 10). Well-log correlation indicates have a general decrease in amalgamation and an in-
that laterally, the crevasse sequences change into crease in bedform preservation from base to top. Cre-
channel-belt sandstones (Figure 21). These channel- vasse splay facies successions occur in the upper half
belt sandstones tend to be single story and laterally of the intermediate-term base-level cycle II (Figure 10).
discontinuous. These short-term cycles are successively composed
The transition from aggradational flood-plain de- of deeper or more distal crevasse splay facies; hence,
posits into crevasse sandstone bodies indicates the these short-term cycles have a landward-stepping stack-
progradation of the crevasses over a distal part of the ing pattern that defines an increasing accommoda-
flood plain. The crevasse splay facies succession at tion regime.
the upper half of this cycle occurs in six short-term Intermediate-term cycle II has a higher propor-
cycles (Figure 15). Short-term cycles cored in Cu- tion of channel facies tracts than cycle I. In addition,
piagua A-1 well consist of more distal crevasse splay the channel facies tracts are thicker and show a much
facies, so each of these cycles steps landward (away higher degree of amalgamation. Channel sandstone
from the channel belt) with respect to the underly- bodies are laterally more continuous (Figures 19, 21).
ing one defining an overall base-level rise in the These indicate that cycle II records a seaward step-
upper part of this intermediate cycle. ping of the depositional systems of the alluvial plain.
Fajardo (1995) showed that this lowermost inter- In Cupiagua, this cycle has the thickest proportion of
mediate cycle in the adjacent Cusiana field is com- granule facies. Also, channels at the base of this cycle
posed of channel to crevasse to flood-plain facies are more amalgamated than cycle I (Fajardo, 1995). This
tracts. His cycle starts with the erosion at the base of evidences also a seaward-stepping stacking pattern.
the channel belt and ends with the next channel belt
base. This is interpreted as a base-level-rise cycle. Cycle III
Fajardo (1995) made an isopach map of his cycle I, This intermediate-term cycle shows the highest
which shows that the cycle occurs only in the south- degree of facies variability. In the three wells (A-1, C-3,
ern part of the Cusiana field. Cross sections across the and H-11; Figures 10–12) that cored this interval, the
462 / Ramon and Fajardo

FIGURE 21. Well-log correlation of the A-1, Q-6, and U-23 wells of the Cupiagua field. Cycle I shows lateral facies changes. Crevasse and swamp and flood-plain
deposits cored in A-1 change laterally into channel-belt sandstones. Depth in feet.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 463

FIGURE 21. (cont.).


464 / Ramon and Fajardo

cycle has diverse proportions of channel, crevasse, Although facies changes from well to well, the
and flood-plain facies tracts. Cycle III is more sym- base-level-rise and base-level-fall hemicycles are pres-
metrical than cycles I and II, but symmetry changes ent in all wells. Well A-1 records the base-level rise as
from well to well. In the A-1 well, the cycle is com- the transition from channel to bay-fill and the base-
posed primarily of agradational flood-plain facies level fall with the change from bay-head delta back
tracts. The base-level-rise hemicycle is composed of into estuarine-channel facies tracts (Figure 10). Well
crevasse splay (20%) and swamp-flood-plain (80%) H-11 records the rise-to-fall cycle as the transition
facies successions (Figure 10). The base-level-fall hemi- from estuarine-channel to bay-head and back into
cycle is composed of swamp (65%), flood-plain (15%), channel facies tracts (Figure 11). Finally, well C-3
and channel (20%) facies successions. In the C-3 well, shows a transition from channel to bay-head facies
the cycle consists mostly of channel facies successions tracts (Figure 12). The presence of new more distal
(Figure 12). Minor intercalations of channel aban- brackish to shallow-marine facies tracts indicates
donment and crevasse facies successions occur. Well that cycle IV records a major landward stepping of
H-11 is composed of channel (60%), crevasse (20%), all facies tracts. The alluvial facies tracts dominant
and swamp (20%) facies tracts (Figure 11). The base- in the Lower Mirador (cycles I, II, and III) shifted
level-rise hemicycle is characterized by a transition toward the east at the time of cycle IV.
from channel to crevasse splay to swamp and flood-
plain facies successions (Figure 11). This trend of Cycle V
facies successions defines an overall increase in Intermediate-term cycle V records a continuous
accommodation space. The intermediate-term base- landward shift of facies tracts. This cycle is com-
level-fall hemicycle is characterized by a transition posed by bay-head delta (40 –70%) and bay-fill (30–
from flood-plain to crevasse to channel facies succes- 60%) facies successions. Channel facies tracts are
sions. This trend defines an overall decrease in accom- only locally present toward the top of the cycle. The
modation/sediment supply ratio. The turnaround point lower and upper boundaries are turnaround points
from intermediate-term base-level rise to fall is located (fall to rise) in conformable strata and are located at
within the massive mudstone facies of the swamp the position of maximum thickness of channel and
and flood plain. bay-head delta facies successions. Cycle V is gener-
Channel-belt sandstones of this intermediate-term ally symmetrical.
cycle III are laterally discontinuous and less amal- The base-level-rise hemicycle of cycle V is char-
gamated as compared to the ones in cycle II. The acterized by a transition from bay-head delta to bay-
highest proportion of flood-plain and crevasse facies fill facies successions. Short-term cycles show landward-
tracts and the single-story character of the channel- stepping stacking patterns. Progressively, bay-head
belt bodies indicate that cycle III represents a landward facies becomes thinner and more bioturbated, indi-
stepping into more distal, lower energy facies tracts. cating preferential preservation of the distal portions.
In the same direction, bay-fill facies tracts become
Cycle IV thicker. Bay-fill mudstones are better developed to
Intermediate-term cycle IV marks a major change the middle part of the cycle. This cycle has the highest
in the depositional setting in the Cupiagua field proportion of distal bay-fill facies tracts. This suggests
area. Cycle IV is composed of channel (30 –60%), a continuous landward-stepping pattern.
bay-fill (0 10%), and bay-head delta (30 –70%) facies The base-level-fall hemicycle of cycle V records the
successions. The lower boundary is generally a base- transition from bay-fill to bay-head delta facies suc-
level-fall surface, whereas the upper boundaries are cessions, which is sometimes capped by channel facies
base-level-rise flooding surfaces. This cycle commonly successions. This trend of facies successions indicates
exhibits base-level-rise and base-level-fall hemicycles decreasing A/S. Short-term cycles have a seaward-
of similar thickness. Mudstones of cycle IV are com- stepping stacking pattern. This transition represents a
monly black or dark gray (Figure 18D) and are bur- shallowing-upward profile produced by progradation.
rowed as compared to mudstones in cycle III, which
exhibit light- to medium-gray, brown, and purple Cycle VI
mottling (Figure 14J, K). The change from mottled This intermediate-term cycle is asymmetrical where
mudstones to dark-gray mudstones is inferred to re- mostly the base-level rise is preserved. This cycle is
flect a change from drier, soil-forming conditions to only cored at well A-1. The lower boundary is located
wet coastal-plain or bay conditions. within conformable strata at the turnaround point
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 465

from decreasing to increasing accommodation-to- stratigraphic boundaries commonly coincide with the
sediment supply ratio. Cycle VI is composed of bay- most laterally continuous shifts of facies, and conse-
fill (40– 70%) and bay-head delta (30 –60%) facies quently, the most pronounced changes in lithological
successions. properties commonly occur across those bounding sur-
In the rise hemicycle, the proportion of deeper facies faces. Thus, petrophysically significant boundaries of
progressively increases, and shallower facies decreases fluid-flow compartments commonly are coincident
within short-term cycles. This upward increase in the with time-significant stratigraphic bounding surfaces.
proportion of deeper facies means that short-term Numerous occurrences of natural stratigraphic bound-
cycles have a landward-stepping stacking pattern. aries are not associated with petrophysical boundaries;
In few wells, the base-level-fall hemicycle is made but occurrences of petrophysical boundaries that are
up of short-term cycles with a general seaward-stepping not associated with stratigraphic boundaries are far
stacking pattern. The upper half of cycle VI shows chan- less common (Cross et al., 1993).
nel facies successions directly over the bay-fill facies A second stratigraphic attribute that can contribute
tract. This would indicate an anomalous seaward shift to reservoir analysis is the relation between dynamic
in facies (downhill offset in facies). Above the channel changes in accommodation/sediment supply ratio
facies tract, a rapid landward shift in facies without and the resulting systematic changes in stratal ar-
stratigraphic discontinuity occurs. This landward off- chitecture and diversity, successions, and continuity
set places deeper bay-fill facies of the C8 member of the of facies preserved in the stratigraphic record. For
Carbonera Formation on top of channel deposits. example, Allen (1978, 1979) and Bridge and Leeder
The higher proportion of channel facies tracts in (1979) suggested that fluvial channel-belt sands de-
cycle VI relative to cycle V suggests a seaward-stepping posited during conditions of slow subsidence (equiv-
stacking pattern. Above the seaward facies offset, the alent to low accommodation) form vertically and
landward-stepping stacking pattern is maintained up laterally interconnected, blanketlike sandstone bod-
to the top of the Mirador Formation. Both the sea- ies that could function as single compartment res-
ward and landward shift in facies are interpreted as ervoirs in terms of fluid flow and pressure. By con-
anomalies in the stacking pattern and are not con- trast, channel-belt sandstones of identical depositional
sidered boundaries of intermediate-term stratigraphic systems may occur as isolated, stringerlike reservoir
cycles. These stratigraphic anomalies were identified sandstones if deposited during periods of faster sub-
in the cored well (A-1, Figure 10) and might be present sidence (equivalent to high accommodation). Each
in uncored wells. sandstone body would function as a separate fluid-
flow compartment.
The Lower Mirador cycles clearly show these dif-
RESERVOIR ZONATION OF THE ferences in channel-belt architecture. Intermediate
MIRADOR FORMATION cycles I and III deposited under relatively higher ac-
commodation conditions have thinner channel-belt
High-resolution stratigraphic correlation, defined sandstone bodies associated with thick crevasse and
as establishing temporal equivalency of rock units re- flood-plain facies tracts. Channel-belt sandstone bod-
gardless of rock type, mineralogy, texture, or environ- ies are laterally discontinuous between wells and show
ment of deposition is the base for a more accurate and strong thickness changes. Facies tract proportions
predictive reservoir characterization. Because the ba- change from well to well, but crevasse and flood-
sic lithological, petrophysical, geometrical, and res- plain facies tracts are generally thicker than in cycle II
ervoir continuity attributes of rocks originate during (Figures 10, 19). As observed in Figure 10, the Cu-
sediment accumulation, accurate correlation permits piagua A-1 well has only a thin channel sandstone in
accurate representation of these rock properties in a cycle I. Close wells show fining-upward, bell-shaped
four-dimensional (time-space) context. High-resolution gamma ray response that is interpreted as channel
correlation is the best means of identifying stratigraph- successions (Figure 21). Stratigraphic position and
ically controlled fluid-flow compartments, compart- thickness of the channel facies successions change be-
ment boundaries, and fluid-flow pathways in reser- tween wells. In contrast, channel-belt sandstone bod-
voirs. This is because the natural bounding surfaces of ies are thicker and laterally more continuous in cycle II
episodically accumulated stratigraphic successions (Figures 10, 11). The relative proportion of overbank
compartmentalize strata into time-bounded packages facies tracts in cycle II is lower as compared to cycles I
of varying spatial and temporal scales. These natural and III. Channel sandstone bodies are laterally more
466 / Ramon and Fajardo

continuous and are present in almost every single bay-fill facies successions. In single wells, the differ-
well in the Cupiagua field (Figures 19, 21, 22). ent facies tracts control the porosity and permeability
Finally, facies tracts control the quality of the fluid- vertical distribution. Well H-11 has massive sand-
flow properties of the sandstone bodies. Porosity stones with only minor variations in the gamma-
versus log permeability as a function of facies suc- ray response over most of the Mirador Formation
cessions were plotted for intermediate-term cycles I (Figure 25). Interpreted facies tracts correspond to
to III (Figure 23). This plot shows that channel facies major changes in the porosity and permeability val-
successions have higher porosity and permeability ues. These different facies tracts then define differ-
values than do crevasse splay facies successions in ent units with different fluid-flow capabilities. The
intermediate-term cycles I, II, and III of the Lower crevasse-splay facies successions typically have po-
Mirador. These figures show two distinctive popula- rosities in the 3–4% range and permeabilities lower
tions: one with higher porosity and permeability, than 0.1 md. Bay-head facies successions typically
which corresponds to the channel facies succession, exhibit porosities in the 5–7% range and permeabil-
and the other with lower porosity and permeability, ities of tens of milidarcys.
which corresponds to the crevasse splay facies succes- Water saturation data from porous plate and Dean
sion. Because both channel and crevasse splay facies Stark analysis performed in core plugs indicate the
succession intervals are laterally continuous through- existence of different rock types. Channel of the Lower
out the field, these differences in petrophysical prop- Mirador and channel and bay-head facies successions
erties become critical in the reservoir zonation of the of the upper Mirador show water saturation values
lower Mirador long-term cycle. ranging from 5 to 7%. Crevasse and bay-fill sand-
Similar differences are found in the bay facies tract stones show water saturation values higher than 17%.
of the upper Mirador (cycles IV– VI; Figure 24). Chan-
nel and bay-head facies successions have higher per- FLUID-FLOW UNITS AND STATIC MODEL
meability values for the same range in porosity as OF THE MIRADOR FORMATION
compared to the bay-fill facies successions. The plot
shows that for a given porosity, the channel and bay- Previous reservoir characterization studies identi-
head sandstones have as much as two orders of mag- fied four mega-rock-types for the three producing
nitude higher permeabilities than the equivalent reservoir sandstones (Guadalupe, Barco, and Mirador

FIGURE 22. Well-log correlation of the wells in the southern part of the Cupiagua field. Depth in feet.
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 467

FIGURE 23. Plot of porosity versus log permeability as a function of facies successions for the coastal-plain facies tract
(intermediate-term cycles I to III).

FIGURE 24. Plot of porosity versus log permeability as a function of facies successions for the bay facies tract
(intermediate-term cycles IV to VI).
468 / Ramon and Fajardo

FIGURE 25. Vertical variation of porosity and permeability in the Cupiagua H-11 well.

formations): mudstone (nonreservoir), lithoarenites, erate reservoir quality. For each of these reservoir
quartzarenites, and phosphatic lithoarenites. The rock types, a porosity-permeability transform was ob-
lithoarenites have poor reservoir quality; the quart- tained by comparing core with wire-line data. For static
zarenites and phosphatic lithoarenites have mod- modeling purposes, to capture the reservoir quality
Sedimentology, Stratigraphy, and Architecture of the Eocene Mirador Formation / 469

variability in the quartzarenites, a total of four facies REFERENCES CITED


associations and their four porosity-permeability trends
Allen, J. R. L., 1978, Studies in fluviatile sedimentation:
were used to cover the entire rock quality range of this
Bars, bar complexes and sandstone sheets (low sin-
mega-rock-type (Figures 23, 24). The facies associa-
uosity braided streams) in the Brownstones (L. Dev.),
tions identified in cored wells were extrapolated to Welsh borders: Sedimentary Geology, v. 26, p. 281 –293.
the rest of the wells using electrical log signatures Allen, J. R. L., 1979, Studies in fluviatile sedimentation:
(gamma ray, spectral gamma ray, neutron, density, An exploratory quantitative model for the architecture
and resistivity). The interwell static modeling was of avulsion-controlled alluvial suites: Sedimentary Geol-
generated using geostatistics techniques to populate ogy, v. 21, p. 129–147.
the facies, followed by the porosity and permeability Barrell, J., 1917, Rhythms and the measurement of geo-
logic time: Geological Society of America Bulletin, v. 28,
based on the facies tracts.
p. 745 – 904.
Net pay is defined at a permeability cutoff equal or Bridge, J. S., and M. R. Leeder, 1979, A simulation model of
higher than 0.1 md, which is a 3.5% porosity and alluvial stratigraphy: Sedimentology, v. 26, p. 617 – 644.
about 388 API gamma-ray cutoff for the Mirador For- Cooper, M. A., et al., 1995, Basin development and tec-
mation quartzarenites. These cutoffs have been eval- tonic history of the Llanos basin, eastern Cordillera,
uated at different scales from pore scale (mercury and Middle Magdalena Valley, Colombia: AAPG
injection data) to centimeter scale in Repeated For- Bulletin, v. 79, no. 10, p. 1421 – 1443.
mation Test flow response, to meter scale using core Coral, M., and W. Rathke, 1997, Cupiagua field, Colombia:
Interpretation case history of a large, complex thrust
fluorescence presence ending with unit flows, pro-
belt gas condensate field: Cartagena, Memorias del VI
ducing intervals confirmed by production logging Simposio Bolivariano ‘‘Exploración Petrolera en las
tools (PLTs) response. Cuencas SubAndinas,’’ Colombia, Tomo 1, p. 119 – 128.
As mentioned before, facies associations were pop- Cross, T. A., et al., 1993, Applications of high-resolution
ulated within the gross rock volume using stochastic sequence stratigraphy to reservoir analysis, in R. Eschard
methods. The 3-D stochastic reservoir model was built and B. Doligez, eds., Subsurface reservoir characteri-
using the statistics of rock type distribution gained zation from outcrop observations: Proceedings of the
7th Exploration and Production Research Conference:
from the Cupiagua core suite and semiregional depo-
Paris, Teichnip, p. 11 – 33.
sitional model. Porosity and water saturation distri-
Dengo, C., and M. Covey, 1993, Structure of the eastern
butions were assigned to each of the rock types in the Cordillera of Colombia: Implications for trap styles
gross rock volume. and regional tectonics: AAPG Bulletin, v. 77, p. 1315 –
Average porosity and permeability values in the 1337.
geocellular static model are consistent with the well Fajardo, A. A., 1995, 4-D stratigraphic architecture and 3-D
average values in each formation and during the up- reservoir fluid flow model of the Mirador Fm., Cusiana
scaling process to generate the reservoir-simulation field, foothills area in the Cordillera Oriental, Colom-
bia: M.Sc. thesis, Colorado School of Mines, Golden,
model; those averages were maintained representa-
Colorado, 171 p.
tive of the well average values. Martinez, J., 2003, Modelamiento estructural 3D y aplica-
The stock tank barrels of original oil initially in place ciones en la exploración y explotación de hidrocar-
(STOOIP) volumes were obtained from the initializa- buros en el Cinturón de Cabalgamiento del Pie-
tion of the dynamic reservoir model, and they are good demonte Llanero, Cordillera Oriental, Colombia: VIII
representations of the static stochastic 3-D model. Simposio Bolivariano — Exploración Petrolera en las
Cuencas Subandinas.
ACKNOWLEDGMENTS Villamil, T., 1999, Campanian – Miocene tectonostratig-
raphy, depocenter evolution and basin development
of Colombia and western Venezuela: Paleogeography,
We are grateful to BP for the support to this work Paleoclimatology, Paleoecology, v. 153, p. 239 – 275.
and permission to publish. The authors thank Steven Wheeler, H. O., 1964, Baselevel, lithosphere surface, and
Bachtel, Paul (Mitch) Harris, and an anonymous re- time-stratigraphy: Geological Society of America Bulletin,
viewer for their helpful suggestions. v. 75, p. 599 – 610.

You might also like